Cutaneous Melanoma 3030050688, 9783030050689

Written by the leading melanoma experts from the United States, Australia, and Europe, this new edition incorporates the

133 21 51MB

English Pages 1539 [1508] Year 2020

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface to the Sixth Edition
Acknowledgments
Contents
Editorial Board
Contributors
Part I: Biology and Immunology of Melanoma
Biology of Melanocytes and Primary Melanoma
Introduction
Development of Melanoblasts
Regulation of Specification
Regulation of Migration
Regulation of Survival and Proliferation
Differentiation of Melanocytes
Regulation of Differentiation
Regulation of Survival
Melanomagenesis
From Melanocyte to Melanoma: A Multistep Process
Molecular Genetics: Early Lessons from Familial and Sporadic Melanoma
Melanoma: A Consequence of Homeostatic Disruption
Melanoma: Cell of Origin
Melanoma and the Environment
Sun Exposure and Epidemiology
Photobiology and Melanoma
Conclusion
References
Immunology of Melanoma
Innate Immunity
Adaptive Immunity
Immune Regulation and Tolerance
Co-stimulatory/Co-inhibitory Receptors
B7-CD28 Family
TNFR Family
TIM Family
Regulatory T Cells
Cytokines
Animal Tumor Models
Antigens Recognized by Tumor-Reactive T Cells
Identification of Tumor Antigens Recognized by T Cells: General Principles
Cancer Germline Antigens
Tissue-Specific Differentiation Antigens
Overexpressed Gene Products
Melanoma Neoantigens
Conclusions: Implications for Immunotherapy
Cross-References
References
Biomarkers for Melanoma
Biomarker Definition and Use
Definition of Cancer Biomarkers
Differential Utilization of Cancer Biomarkers
Biomarker Use in Melanoma
Biomarker Discovery and Validation
Biomarker Discovery
Biomarker Validation
Tumor Tissue-Based Markers
Diagnostic Markers for Primary Melanoma
FISH-Based Diagnostic Assays for Melanoma
Gene Expression Profiling of Melanocytic Neoplasms
Immunohistochemical Analysis in the Diagnosis of Melanocytic Neoplasms
Prognostic Markers for Primary Melanoma
Gene Expression Profiling of Melanoma Prognostic Markers
IHC Analysis of Melanoma Prognostic Markers
Tumor Environment-Based Non-soluble Biomarkers
Tumor-Initiating Cells
Epithelial-Mesenchymal Transition
Epigenetic Changes
Immune Escape Mechanisms
Soluble Biomarkers
Lactate Dehydrogenase (LDH)
S100B
Other Serum Biomarkers (CRP, FGF, IL-8, MIA, SAA, VEGF, YKL-40)
Circulating Tumor Cells (CTCs)
Circulating Tumor DNA (ctDNA)
Circulating MicroRNA (miRNA)
Circulating Immune Cells
Treatment-Associated Biomarkers
Serum Lactate Dehydrogenase (LDH)
Burden of Metastatic Disease
Body Mass Index (BMI)
Molecular Features Associated with Outcome on BRAF Inhibitor-Based Therapy
References
Part II: Diagnosis and Staging
Clinical Presentations of Melanoma
Introduction
Patterns of Presentation
Clinical Assessment
Patient History
Personal History of Skin Cancer
Family History
Phototype and Sun Exposure
Signs and Symptoms
Physical Examination
Clinical Features
Growth Patterns
Aids to Diagnosis
Clinical Photography
Dermoscopy
Reflectance Confocal Scanning Laser Microscopy (RCM)
Image Analysis for Diagnosis
Other Techniques: Multispectral Imaging, Electrical Impedance Spectroscopy, Adhesive Patch Molecular Assays, Optical Coherence...
Multispectral Imaging
Electrical Impedance Spectroscopy
Adhesive Patch Molecular Assays
Optical Coherence Tomography
Ultrasound Imaging
Evolving Paradigms in the Visual Assessment of Skin Lesions
Cross-References
References
Dermoscopy/Confocal Microscopy for Melanoma Diagnosis
Introduction
Dermoscopy
Basics of Dermoscopy
Diagnostic Accuracy of Dermoscopy
Reflectance Confocal Microscopy
Basics of Reflectance Confocal Microscopy
Diagnostic Accuracy of Reflectance Confocal Microscopy
Clinical Scenarios for the Use of Reflectance Confocal Microscopy
Total Body Photography and Digital Dermoscopy
Other Noninvasive Detection Methods
Dermoscopic Features of Melanoma and Melanoma Simulants
Local Dermoscopic Features
Common Global Dermoscopic Patterns
Anatomical Site Considerations
Featureless Melanomas
Diagnostic Algorithms for Dermoscopy
Two-Step Algorithm
Three-Point Checklist
Revised Seven-Point Checklist
ABCD Rule of Dermoscopy and Clinical EFG Rule
Menzies Method
CASH Acronym
Chaos and Clues
TADA
Pattern Analysis
Reflectance Confocal Microscopy Features of Melanoma
Common Confocal Features of Melanoma
Diagnostic Algorithms for Reflectance Confocal Microscopy
Modena Algorithm
Barcelona Algorithm
Dermoscopic and Confocal Features of Non-superficial Spreading Melanoma Subtypes
Nodular Melanoma
Dermoscopic Features of Nodular Melanoma
Reflectance Confocal Microscopy Features of Nodular Melanoma
Lentigo Maligna and Lentigo Maligna Melanoma
Dermoscopic Features of Lentigo Maligna and Lentigo Maligna Melanoma
Reflectance Confocal Microscopy Features of Lentigo Maligna and Lentigo Maligna Melanoma
Spitzoid Melanomas
Dermoscopic Features of Spitzoid Melanomas
Reflectance Confocal Microscopy Features of Spitzoid Melanomas
Desmoplastic Melanomas
Dermoscopic Features of Desmoplastic Melanomas
Reflectance Confocal Microscopy Features of Desmoplastic Melanomas
Conclusion
Cross-References
References
Biopsy of Suspected Melanoma
Prebiopsy Lesion Assessment
Biopsy Techniques
Excisional Biopsy
Incisional Biopsy
Fine-Needle Aspiration and Core Biopsy
Frozen Sections
Biopsy of the Nail Unit
Biopsy of Mucosa
Conclusion
References
Lymphoscintigraphy in Patients with Melanoma
Introduction
The Definition of a Sentinel Node
Lymphatic Mapping of the Skin: Early Studies
Lymphoscintigraphy
The First Radiocolloid, and Early Experience with Lymphoscintigraphy
Radiopharmaceuticals
99m-Technetium Labeled Colloids
99mTc-Antimony Sulfide Colloid
99mTc-Rhenium Sulfide Colloid and 99mTc-Nanocolloid of Albumin
99mTc-Sulfur Colloid
99mTc-Human Serum Albumin
99mTc-Tilmanocept
What Is the Ideal Radiocolloid?
Lymphoscintigraphy for Sentinel Lymph Node Biopsy Procedures
The Technique of Lymphoscintigraphy
Injecting the Tracer
Immediate Dynamic Imaging
Delayed Static Imaging
Unexpected Lymphatic Drainage Pathways
SPECT/CT Imaging of the Sentinel Nodes
Marking the Surface Location of the Sentinel Node
Radiation Dosimetry: Risks to the Patient
Radiation Dosimetry: Risks to a Pregnant Patient
Radiation Dosimetry: Risks to Attending Staff
Patterns of Lymphatic Drainage From the Skin
Trunk (n = 4604 patients)
Posterior Trunk (Including Posterior Chest) (n = 3803 patients)
Anterior Trunk (Including Anterior Chest) (n = 801 patients)
Head and Neck (n = 2995 patients)
The Limbs
Upper Limb (n = 2738 patients)
Lower Limb (n = 2981 patients)
Interval Nodes (n = 1285 patients)
Drainage to Multiple Node Fields
Complex Lymphatic Drainage Patterns
Lymphoscintigraphy in Patients with Clinically Involved Node Fields
The Future
New Tracers for Lymphatic Mapping
Gamma Probes for Intraoperative Detection of Radioactivity
Gamma Cameras for Intraoperative Imaging
Alternative Methods of Lymphatic Mapping
Conclusions
References
Biopsy of the Sentinel Lymph Node
History and Conceptual Basis of Sentinel Lymph Node Biopsy
Rationales for SLN Biopsy
Rationale: Staging
Staging Value of SLN Biopsy: Relationship to Primary Tumor Thickness
Rationale: Regional Disease Control
Rationale: Survival Improvement
Selection for SLN Biopsy
Technical Details of Mapping
Special Situations: Difficult Sites
Special Situations: Patients Presenting After Wide Excision
Special Situations: Nonclassical Nodal Sites
Pathology of the SLN
False-Negative SLNs
Complications
Lymphatic Mapping and SLN Biopsy from Melanoma Metastases
Completion Lymph Node Dissection
SLN as an Experimental Model for Tumor-Host Interface
Considerations for the Future
Cross-References
References
Melanoma Prognosis and Staging
Overview of the Eighth Edition AJCC Melanoma Staging System
Prognostic Factors and Staging of Primary Melanoma (AJCC Stages I and II)
Primary (Breslow) Tumor Thickness
Primary Tumor Ulceration
Clark Level of Invasion
Primary Tumor Mitotic Rate
Patient Age
Tumor-Infiltrating Lymphocytes
Lymphovascular Invasion
Neurotropism
Regression
Other Variables
Staging Value of the Sentinel Lymph Node
Prognostic Factors and Staging in Regionally Metastatic Melanoma (Stage III): Lymph Node, In-Transit, Satellite, and Microsate...
Number of Involved Regional Lymph Nodes
Clinically Occult Versus Clinically Detected Regional Lymph Node Metastases
Sentinel Lymph Node Tumor Burden
Extranodal Extension
Non-Nodal Locoregional Metastases (Microsatellite, Satellite, and In-Transit Metastases)
Metastatic Melanoma to Lymph Node(s) from an Unknown Primary Site
Prognostic Factors and Staging of Patients with Distant Metastatic Melanoma (Stage IV)
Site of Distant Metastasis
LDH Level
Other Factors
Additional Staging Recommendations
Patients with Multiple Primary Melanomas
Staging Patients After Systemic or Radiation Therapy
Staging Patients at Recurrence
Conclusions
References
Models for Predicting Melanoma Outcome
Introduction
Prediction Tools and Statistical Models
Personalized Prognosis
Clinical Applications
The Link Between Prediction Tools and Staging Systems
The Relevance of Prediction Tools for Clinical Trials
Brief History of Melanoma Prediction Tools
Prediction Tools Developed from AJCC Databases
Other Prediction Tools
Planning to Build a Prediction Model
Reporting Prediction Models
Criteria for Building Prediction Models
Selection of a Patient Population
Selection of an Outcome to Predict
Time to Relapse After Initial Disease Management
Conditional Survival
Probability of Binary Outcome
Considering the Treatment Landscape
Selection of Relevant and Clearly Defined Predictors
Model Development
Cox Proportional Hazards Model
Hazard Function
Relative Risk
The Proportional Hazards Assumption
When the Proportional Hazards Assumption Is Violated
Model Validation and Performance
Internal Validation
External Validation
Putting Contemporary Models to the Test
Challenges and Opportunities
Sample Size
Selection Bias
Missing Data
Survival from Metastatic Melanoma
The Future
Cross-References
References
Part III: Pathology
Classification and Histopathology of Melanoma
Introduction
The Role and Challenges of Pathologic Assessment of Melanocytic Tumors
Biopsying Clinically Suspicious Pigmented Tumors
Pathologic Assessment of Primary Melanomas
Accuracy of Pathologic Assessment Is Enhanced by Clinical Correlation
The Role of Specimen Orientation
Melanoma Tumor Progression: The Concept of Radial and Vertical Growth Phases
Pathways of Melanoma Pathogenesis and Clinicopathologic Classification of Melanoma
Low Cumulative Sun Damage Melanoma/Superficial Spreading Melanoma
Pitfalls
High-Cumulative Sun Damage Melanoma/Lentigo Maligna Melanoma
Pitfalls
Assessment of Excision Margins in Lentigo Maligna
Acral Melanoma
Nodular Melanoma
Histologic Features of the Vertical Growth Phase of Melanoma
Predominantly Epithelioid Cell Vertical Growth Phase
Predominantly Spindle Cell Vertical Growth Phase
Pitfalls
Mixed Spindle Cell and Epithelioid Cell Vertical Growth Phase
Nevoid Vertical Growth Phase (``Nevoid Melanoma´´)
Other Melanoma Subtypes and Variants
Desmoplastic Melanoma
Differential Diagnosis
Desmoplastic Melanoma Versus Sclerosing Nevus (ScN)
Desmoplastic Melanoma Versus Scar or Fibroma
Desmoplastic Melanoma Versus Non-melanocytic Desmoplastic Malignant Spindle Cell Neoplasms
Nevoid Melanoma
Pigmented Epithelioid Melanocytoma
Melanoma Arising From a Blue Nevus
Mucosal Melanoma
Vulvar and Penile Melanoma
Conjunctival Melanoma
Prognosis
Management
Atypical Spitz Nevi/Atypical Spitzoid Tumors and Spitz Melanoma (Malignant Spitz Tumor)
Melanocytic Tumors of Uncertain Malignant Potential (MELTUMP), Intermediate Melanocytic Proliferations and Melanocytomas
Management
Spitz Melanoma/Malignant Spitz Tumor
Histopathologic Features of Prognostic Importance
Breslow Thickness
Clark Level of Invasion
Ulceration
Mitotic Rate
Inflammatory Host Response (Including TILs)
Regression
Microscopic Satellites
Blood Vessel and Lymphatic Invasion
Angiotropism
Neurotropism
Desmoplasia
Anatomic Site
Sex
Age
Lymph Node Metastasis
The Melanoma Pathology Report Including a Synoptic Format
The Synoptic Pathology Report
Melanoma in Children and Adolescents
Histopathology
Cutaneous Metastases of Melanoma
Regional Lymph Node Metastases of Melanoma
Laboratory Assessment of Regional Lymphadenectomy Specimens
Lymphatic Mapping and Sentinel Lymph Node Biopsy
Laboratory Assessment of Sentinel Lymph Nodes
Laboratory Confirmation That a Submitted Node Is Truly Sentinel
Intraoperative Evaluation of Sentinel Nodes
The Need to Evaluate Multiple Levels of the Sentinel Lymph Node
The Role of Immunohistochemistry in Evaluation of Sentinel Lymph Nodes
Identification of Nodal Nevi and Their Separation from Metastatic Melanoma
Sentinel Lymph Node Evaluation in Assessment of the Malignant Potential of Ambiguous (Nevoid) Melanocytic Lesions
Molecular Biology as a Supplement to Histological Evaluation of Sentinel Nodes
Measurement of the Amount of Tumor Present and Its Distribution in Sentinel Lymph Nodes
Melanoma Metastatic to Visceral Organs and Other Sites
Fine Needle Biopsy in Melanoma Patients
Clear Cell Sarcoma (Melanoma of Soft Parts)
Molecular Pathology of Melanoma
Therapeutic Targets
The Concept of the Cancer Stem Cell
The Metastatic Niche Concept
Closing Remarks
Cross-References
References
Molecular Pathology and Genomics of Melanoma
Introduction
The Beginnings of Cancer Genetics
Basic Principles and Terms in Cancer Genetics
Types of Genetic Aberrations
Recent Advances in Cancer Genetics
Pathogenesis: Acquisition of Mutations
Ultraviolet radiation (UV) Pathogenesis
Stepwise Progression of Melanocytic Tumors
Nevi
Intermediate Melanocytic Tumors
Melanoma
Genetic Aberrations in Melanocytic Tumors
MAPK-Activating Aberrations
BRAF
NRAS
NF1
MAP2K1/MAP2K2(MEK1/MEK2)
KIT
GNAQ/GNA11/CYSLTR2/PLCB4
Translocations
Loss-of-Function Mutations
CDKN2A
PTEN
Noncoding Genetic Aberrations
TERT Promoter Mutations
Mutations in Other Noncoding Regions of the DNA
Noncoding RNAs
Genes Associated with Increased Melanoma Susceptibility
Cell Cycle Genes: CDKN2A and CDK4
Telomere-Associated Mutations: TERT, POT1, ACH, and TERF2IP
BAP1
MITF
MC1R
Cancer Syndromes Associated with Increased Risk of Melanoma
Genetic Tests in the Diagnosis of Melanoma
Polymerase Chain Reaction-Based Methods
PCR
Quantitative PCR and Gene Expression-Based Assays
DNA Sequencing Methods
Sanger Sequencing
Next-Generation Sequencing
Methods to Detect DNA Copy Number Aberrations
Array Comparative Genomic Hybridization
SNP Arrays and Molecular Inversion Probes
Fluorescence In Situ Hybridization
Genetic Testing for Therapeutic Decision-Making
BRAF and MAPK Inhibitors
Immunotherapy
Outlook
References
Part IV: Epidemiology and Prevention of Melanoma
Clinical Epidemiology of Melanoma
Introduction
Worldwide Incidence and Mortality Patterns
Global Distribution
Highest Rates
Temporal Trends in Incidence
Mortality
Melanoma Risk Factors
Demographic
Age
Sex and Age by Sex
Socioeconomic Status
Ethnicity
Constitutional
Phenotypic
Anthropometric Measures
Environmental
Ambient UVR and Exposure to Sun
Other Environmental Exposures
Ionizing Radiation
Artificial Sources of UVR
Occupational
Reproductive Factors and Exogenous Hormone Use
Health History
Diet
Smoking
Alcohol
Trauma
Immunosuppression
Screening
Risk Prediction
Summary and Conclusions
Cross-References
References
Molecular Epidemiology of Melanoma
Molecular Characterization of Melanoma
Overview of Genomic Features
Cutaneous Melanoma (Non-desmoplastic and Non-acral Types)
Cutaneous Melanoma (Desmoplastic Type)
Acral and Mucosal Melanomas
Uveal Melanoma
Genomic Factors and Melanoma Risk
Genetic Susceptibility to Melanoma
Gene-Environment, Gene-Phenotype, and Gene-Gene Interactions
Gene-Gene Interactions and Melanoma Risk
Interactions Between Genes, Phenotype, and the Environment
Clinical and Public Health Applications
Future Studies
Cross-References
References
Clinical Genetics and Risk Assessment of Melanoma
Introduction
Risk Factors
All Types of UV Radiation Classified as ``Carcinogenic to Humans´´
Sun Exposure
Sunbed Use
Phenotypical Factors
Common Melanocytic and Atypical Nevi
Phenotypical Factors: Skin, Hair, and Eye Color and Freckles
Second Primary Malignancies Among Melanoma Patients and Risk of Second Primary Melanoma Among Other Primary Cancer Patients
Actinic Damage Indicators
Personal and Family History of Other Cancers in Melanoma Families
Melanoma and Neurological Disorders
Genetics
Rare High-Penetrance Melanoma Genes
Melanocortin 1 Receptor
GWAS Analyses for Melanoma and Nevi
Telomere Biology and Aging
Epigenetic and DNA Methylation
Gut Microbiome and Melanoma
Vitamin D
Integrating Risk Factors in the Clinical Setting
Historical Features
Personal History
Genetic Counseling
Childhood and Adolescent Melanoma
Risk Factors for Childhood Melanoma
Genetic Susceptibility of Childhood Melanoma
References
Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma
Introduction
Phenotypic Markers
Melanocytic Nevi as Risk Factors
Total Nevus Count
Clark Nevi (Synonym ``Large Acquired Nevi,´´ ``Atypical Nevi,´´ ``Dysplastic Nevi´´)
Congenital Melanocytic Nevi
Spitz Nevi
BAP1-Inactivated Melanocytic Tumors (``Bapomas´´)
Classic Genotype/Phenotype Risk Correlates
MC1R Polymorphisms (Variants), Skin Color, and Hair Color
Eye Color
Freckles and Lentigines
Precursor Lesions
Clark Nevus (Syn. Large Acquired Nevi, Atypical, Dysplastic Nevi)
Congenital Melanocytic Nevi
Solar Lentigo as Precursor in Xeroderma Pigmentosum Patients
Other Yet-to-Be-Defined Potential Intermediate Lesions
Management of High-Risk Patients
Individuals with Specific Subtypes of Nevi
Congenital Melanocytic Nevi
Spitz Nevi
Patients with BAP1-Inactivated Melanocytic Tumors
Individuals with Many Nevi and Clark Nevi
Patients with Single to a Few Clark Nevi
Patients with Many Nevi and Many Clark Nevi (So-Called Atypical Mole Syndrome)
Conclusion
References
Melanoma Prevention and Screening
Prevention of Melanoma
Trends in Melanoma Incidence and Mortality
Primary Prevention
Reducing Personal Exposure: Shade, Clothing, and Sunscreens
Shade and Clothing
Use of Sunscreens
Sunscreen Types and Controversies
Behavioral Change Programs for Reducing Personal Exposure
Multicomponent Community-Wide Interventions
Youth Education and Counseling Programs
US Preventive Services Task Force (USPSTF): Behavioral Counseling
The SunSmart Program
Intervention Trials of the Prevention of Nevi
Prevention of Nevi in Children
Controlling Exposure to Indoor Tanning Beds
Therapeutic Prevention of Melanoma and Populations to Target for Interventional Trials
Therapeutic Prevention of Melanoma
Candidate Agents
Pigmentation Enhancers
DNA Repair Enzymes
Vitamins and Minerals
Repurposed Therapeutic Agents
Aspirin and Nonsteroidal Anti-inflammatory Drugs (NSAIDs)
Statins
N-Acetylcysteine
Difluoromethylornithine
Phytochemicals (Plant-Derived Biologically Active Compounds)
Secondary Prevention
Early Detection and Screening
Potential Benefits of Screening
Uncertainties and Conflicts in Melanoma Screening
Screening-Related Harms
Prevalence of Screening
Evidence Relating to the Effectiveness of Screening
Randomized Trials
Case-Control Studies
Economic Assessments of Screening
Programs of Screening
Population Screening Programs
Germany
University of Pittsburgh Medical Center
Educational Campaigns to Promote Early Detection
Opportunistic Screening in Normal Medical Practice
Skin Self-Examination
Open-Access Skin Checks
American Academy of Dermatology (AAD)
Euromelanoma
Screening for Occupational Groups
Screening for Selected High-Risk Groups
Adjuncts to Clinical Examination
Challenges in the Detection of More Lethal Melanoma Subtypes
Advances in Screening Technologies and Community Outreach to Improve Early Detection
Conclusion
References
Part V: Management of Primary Melanoma and Locoregional Metastases
Treatment of Primary Melanomas
Historical Perspective and the Emergence of a Contemporary Paradigm
Wide Excision of Primary Melanomas: Fundamental Concepts
T0: Melanoma In Situ/Lentigo Maligna
Current Excision Margin Recommendations for T0 Melanomas
T1: Invasive Melanomas 1 mm in Thickness
Randomized Trials of Excision Margins for T1 Melanomas
Nonrandomized Studies of Excision Margins for T1 Melanomas
Current Excision Margin Recommendations for T1 Melanomas
T2: Invasive Melanomas >1-2 mm Thick
Randomized Trials of Excision Margins for T2 Melanomas
Nonrandomized Studies of Excision Margins for T2 Melanomas
Current Excision Margin Recommendations for T2 Melanomas
T3: Invasive Melanomas >2-4 mm in Thickness
Randomized Trials of Excision Margins for T3 Melanomas
Nonrandomized Studies of Excision Margins for T3 Melanomas (>2-4 mm)
Current Excision Margin Recommendations for T3 Melanomas
T4: Melanomas >4 mm in Thickness
Randomized Trials of Excision Margins for T4 Melanomas (>4 mm)
Nonrandomized Studies of Excision Margins for T4 Melanomas
Current Excision Margin Recommendations for T4 Melanomas
Excision Margins Summary
Techniques for Routine Wound Closure
Excisions for Melanomas in Unusual or Restrictive Locations
References
Reconstructive Options Following Surgery of Primary Melanoma
Introduction
Principles of Reconstructive Surgery
Assessment of the Acquired Defect
Patient Factors
Disease Factors
Local Tissue Factors
Reconstructive Options for Complex Wounds
Definitions
Grafts
Flaps
Undermining with Primary Closure
Skin Grafts
Composite Grafts
Local Flaps
Regional Flaps
Free Tissue Transfer
Head and Neck Reconstruction
Scalp
Forehead
External Ear
Nose
Cheek
Lip
Eyelid
Extremities
Volar Skin of Foot and Hand
Distal Phalanx
Digits
Dorsal Skin
Joints
Reconstruction of Large Wide Local Excision Limb Defects
Reconstruction of Defects of the Groin and Axilla
Approaches to Reconstruction of the Groin
Regional Flaps to the Groin
Anterolateral Thigh Flap
Rectus Abdominis Flap
Gracilis Flap and Perforator Flap Variations
Reconstruction of Defects of the Axilla
Skin Graft
Local Flaps
Latissimus Dorsi Flap
Pectoralis Major Flap
Reconstruction of Mucosal Melanoma Defects
References
Axillary and Epitrochlear Lymph Node Dissection for Melanoma
Introduction
Axillary Dissection
Anatomy
Surgical Technique
Preoperative and Perioperative Preparation
Incision
Skin Flaps
Dissection of Nodal Tissue
Closure
Operative Considerations for Recurrent or Bulky Axillary Metastases in the Upper Axilla
Postoperative Management
Complications
Epitrochlear Dissection
Rationale
Anatomy and Surgical Technique
References
Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma
Inguinofemoral Dissection
Indications
Technique
Modifications of the Classic Technique of Inguinofemoral Lymphadenectomy
Iliac/Obtuartor (Deep Pelvic) Lymph Node Dissection
Benefits
Indications
Operative Technique
Modifications of the Classic Technique of Iliac/Obturator Lymphadenectomy
Robotic-Assisted Transperitoneal Pelvic Lymphadenectomy
Postoperative Complications: Incidence and Risk Factors
Complications of Lymph Node Dissection
Management of Postoperative Complications
Popliteal Dissection
Indications
Operative Technique
Cross-References
References
Neck Dissection and Parotidectomy for Melanoma
Head and Neck Lymphatics and Their Impact on Melanoma Outcomes
Neck Dissection and Parotidectomy for Melanoma
Technique for Neck Dissection and Parotidectomy
Completion Lymph Node Dissection Utility in Head and Neck Melanoma
Conclusion
References
Local and Recurrent Regional Metastases of Melanoma
Introduction
Local and Regional Recurrence of Melanoma
Local Recurrence
In-transit Recurrence
Regional Nodal Recurrence
Hyperthermic Isolated Limb Perfusion
History and Early Clinical Studies
Patient Selection
Preoperative Evaluation
Equipment
Operation
Leak Monitoring
Agents
Hyperthermia
Results
Specific Toxicities and Management
Isolated Limb Infusion
Background
Patient Selection and Indications
Technique
Response to Therapy
Survival After ILI
Burden of Disease
Toxicity
Intralesional Therapies for Cutaneous Melanoma
Introduction
Bacille Calmette-Guerin (BCG)
Interleukin-2 (IL-2)
Granulocyte Macrophage-Colony Stimulating Factor (GM-CSF)
Velimogene Aliplasmid (Allovectin)
Talimogene Laherparepvec (T-VEC)
Rose Bengal (PV-10)
Daromun (L19IL2 + L19TNF)
Coxackievirus A21
Combination with Systemic Immune Therapies
Electrochemotherapy (ECT)
Conclusion
Neoadjuvant Therapy for Borderline Resectable Nodal Metastasis
References
Radiotherapy for Primary and Regional Melanoma
Introduction
Lentigo Maligna
Primary Melanoma
Adjuvant RT After Wide Local Excision of a Primary Melanoma
Adjuvant RT for Resected Stage III Melanoma
RT for Inoperable Regional Node Metastases and In-Transit Metastases
References
Adjuvant Systemic Therapy for High-Risk Melanoma Patients
Introduction
Who Should Be Considered for Adjuvant Therapy?
Prior Adjuvant Therapeutic Approaches (Interferon and Vaccines)
Interferons
Role of Dose and Duration of IFN-α Therapy in Melanoma
Ulcerated Primary Melanoma: A Potentially More IFN-Sensitive Population
Immune Checkpoint Blockade
Anti-CTLA-4
Ipilimumab
Anti-PD1
Nivolumab
Pembrolizumab
Targeted Therapy
Choice of Adjuvant Therapy in the BRAF V600E/K Mutation-Positive Patient
Contraindications to Immunotherapy
Conclusions
References
Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients
The Current Landscape of Systemic Therapy for Stage III and IV Melanoma
Patients with Clinical Stage III Melanoma Are Ideal Candidates for Neoadjuvant Treatment
Rationale for Neoadjuvant Therapy
The History of Neoadjuvant Therapy Use in Melanoma
Neoadjuvant Biochemotherapy
Neoadjuvant High Dose Interferon
The Current State of Neoadjuvant Therapy
Neoadjuvant Targeted Therapies
Dabrafenib Plus Trametinib
Neoadjuvant Immune Checkpoint Inhibitors
Pembrolizumab
Ipilimumab Plus Nivolumab
Neoadjuvant Checkpoint Inhibitor Therapy in Combination with Other Therapies
Neoadjuvant Local Therapies
Talimogene Laherparepvec
L19IL2 plus L19TNF
Neoadjuvant Therapy for Borderline Resectable or Unresectable Melanoma
Targeted Therapies for Borderline Resectable/Unresectable Melanoma
Immune Checkpoint Blockade for Borderline Resectable/Unresectable Melanoma
Predicting Response to Neoadjuvant Therapy
Prognostic Biomarkers
Predictive Biomarkers
Unmet Needs in Neoadjuvant Therapy for Melanoma
Therapeutic Goals and Considerations
Conclusion
References
Hyperthermic Regional Perfusion for Melanoma of the Limbs
Introduction
Epidemiology and Natural History of Extremity In-Transit Disease
Treatment of Recurrent Extremity Melanoma
Historical Perspective and Early Clinical Series
Technical Aspects of Isolated Limb Perfusion
Melphalan
Pharmacokinetics of Melphalan in Isolated Limb Perfusion
Other Chemotherapeutics
Hyperthermia
Tumor Necrosis Factor
Toxicity of Isolated Limb Perfusion
Adjuvant Isolated Limb Perfusion
Therapeutic Limb Perfusion
Melphalan Alone
Melphalan and TNF
Future Role of Limb Perfusion in Current Era of More Effective Systemic Therapy
Conclusions
References
Isolated Limb Infusion for Melanoma
Introduction and Historical Perspective
Patient Selection for Isolated Limb Infusion
Technical Details of the Isolated Limb Infusion Procedure
Preoperative Assessment and Management
Insertion and Positioning of Arterial and Venous Catheters
Procedure in the Operating Room
Postoperative Course and Care
Similarities and Differences Between Isolated Limb Infusion and Hyperthermic Isolated Limb Perfusion
Drugs Used in Isolated Limb Infusion
Pharmacokinetics of Melphalan During Isolated Limb Infusion
Melphalan Dosage and Ideal Body Weight
Use of Microdialysis During Isolated Limb Infusion
Toxicity and Side Effects Following Isolated Limb Infusion
Locoregional Side Effects of Isolated Limb Infusion
Limb Toxicity Following Isolated Limb Infusion
Systemic Toxicity and Complications of Isolated Limb Infusion
Clinical Results of Isolated Limb Infusion
Response Rates Following Isolated Limb Infusion
Limb Recurrence-Free Interval and Overall Survival Following Isolated Limb Infusion
Prognostic Factors for Outcome Following Isolated Limb Infusion
Special Isolated Limb Infusion Regimens and Indications
Novel Isolated Limb Infusion Regimens
Future of Isolated Limb Infusion
Conclusions
Cross-References
References
Surveillance and Follow-Up of Melanoma Patients
Goals of Surveillance
Patterns of Melanoma Recurrence
Risk of Local and Regional Recurrence
Regional Relapse
Time to Recurrence
Strategies for Active Follow-Up of Melanoma Patients
Detection of Recurrences
Role of Physical Examination
Patient Education
Patient Well-Being and Follow-Up
Follow-Up Schedules
Radiologic Studies and Laboratory Tests
Screening for Risk of New Primary Melanomas
Screening for Other Primary Cancers
Current Recommendations for Surveillance
References
Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes, and Selection of Treatment Options
Introduction
Etiology of Local and In-transit Metastases
Incidence of Locoregional Metastasis and Survival
Treatment Options for Local and In-transit Metastases: Overview
Surgical Excision
Cryotherapy
Topical Treatments
Diphencyprone Cream
Imiquimod Cream
Intralesional Therapies
BCG
DNCB
Interferon Alpha
Allovectin
Coxsackie Virus A-21
Interleukin-2 (IL-2)
Rose Bengal (PV-10)
Talimogene Laherparepvec (T-VEC)
Daromun (Combined IL-2 and TNF)
Intralesional Oncolytic Immunotherapy and the Abscopal Effect
Laser and Light-Based Therapies
Ablative Laser Therapy
Non-ablative Laser Therapy
Radio-Frequency Ablation
Photodynamic Therapy
Electrochemotherapy
Reported Results of ECT
ECT in Perspective
Regional Therapies
Isolated Limb Perfusion and Infusion
Radiation Therapy for Locally Recurrent Melanoma and In-transit Metastases
Amputation
Systemic Therapy
Adjuvant and Neoadjuvant Therapies
Adjuvant Therapies
Neoadjuvant Therapies
Conclusions
References
Part VI: Uncommon Presentations of Melanoma
Acral Lentiginous Melanoma
Epidemiology
Pathogenesis
Clinical Features
Dermoscopic Findings
Pathological Features
Immunohistochemistry of ALM
Molecular Feature
BRAF
NRAS
KIT
TERT
NF1
Mutation Burden
Diagnosis
Prognosis
Compared with Other Subtypes of CM
Comparison in the Localization of ALM
Compared in Racial Groups
Treatment
Surgery
Immunotherapy
Targeted Therapy
Conclusion
References
Lentigo Maligna Melanoma
Introduction
Epidemiology and Risk Factors
Natural Course
Clinical Features
Dermoscopic Features of Lentigo Maligna
Reflectance Confocal Microscopy Diagnostic Features of Lentigo Maligna
Diagnostic Biopsy Techniques
Histopathologic Diagnostic Challenges
Treatment
Surgical Modalities
Standard Wide Excision
Surgical Techniques with Complete Peripheral Margin Assessment
Staged Excision Techniques
Staged Excision with Radial Vertical Sections
Square Technique
Mohs Surgery
Histopathological Challenges Associated with Lentigo Maligna Surgical Margins
Nonsurgical Techniques
Topical Imiquimod
Neoadjuvant Use of Imiquimod Prior to Surgery
Use of Imiquimod as Primary Therapy for LM
Use of Imiquimod as Adjuvant Therapy for LM
Monitoring for Response and Recurrence During Imiquimod Treatment
Radiation Therapy (Radiotherapy) for LM/LMM
Long-Term Follow-Up
Potential Role of Reflectance Confocal Microscopy for LM Management
Quality of Life Considerations
Conclusion
References
Mucosal Melanoma
Introduction
Epidemiology
Pathological Features and Diagnosis
Staging and Prognosis
Mucosal Melanoma of the Head and Neck
Differential Diagnosis
Staging and Prognosis
Treatment
Treatment Overview
Female Genital Tract Mucosal Melanomas
Vulvar Melanoma
Vulvar Melanoma Summary
Vaginal Melanoma
Vaginal Melanoma Treatment Overview
Cervix and Urethra Melanoma
Mucosal Melanoma of the Penis and Scrotum
Anorectal Mucosal Melanoma
Anorectal Melanoma Treatment Overview
Gastrointestinal Tract Melanoma
Mucosal Melanoma Adjuvant Therapy
Mucosal Melanoma Systemic Therapy
References
Melanoma in Children and Teenagers
Introduction
Epidemiology
Pathology, Molecular Characteristics, and Differential Diagnosis
Congenital Melanoma
Giant Congenital Nevi
Neurocutaneous Melanosis
Small- and Medium-Sized Congenital Nevi
Etiology
Clinical Presentation and Risk Factors
Surgical Management
Medical Management
Interferon
Chemotherapy
Targeted Therapy
Immunotherapy
Outcome
Follow-Up and Surveillance
Conclusion
Cross-References
References
Pregnancy and the Use of Hormones in Melanoma Patients
Aspects of Female Reproduction Physiology Including Lactation Relevant to Melanoma
Do Melanomas Have Functional Estrogen Receptors?
Do Exogenous Hormones Impact the Risk of Developing Melanoma?
Does Pregnancy Impact the Risk of Developing Melanoma?
Does Pregnancy Impact the Prognosis of Melanoma?
Should a Patient with a Prior History of Melanoma Become Pregnant?
Should Exogenous Hormones Including Infertility Drugs Be Used in a Patient with History of Melanoma?
Should Hormonal Contraceptives Be Used in a Patient with History of Melanoma?
Should Hormonal Infertility Drugs Be Used in a Patient with History of Melanoma?
Should Postmenopausal Hormone Replacement Therapy Be Used in a Patient with History of Melanoma?
How Does Management of Melanoma Differ in the Pregnant Woman?
Surgery Including Sentinel Node Biopsy in the Pregnant or Lactating Patient
How Does Follow-Up of Melanoma Differ in the Pregnant Woman?
Transplacental Transmission and Neonatal Melanoma: Should Termination Ever Be Considered?
Which Antimelanoma Drugs Can Be Used During Pregnancy?
Does Puberty Influence the Outcome of Pediatric Melanoma?
Conclusions
References
Diagnosis of Stage IV Melanoma
Introduction
Timing of Distant Metastasis
Pattern of Metastasis
Prognosis and Prognostic Factors
Sites of Distant Metastases
Number of Metastatic Sites
Elevated Serum Lactate Dehydrogenase
Duration of Remission
Performance Status
Other Prognostic Factors
Clinical Evaluation of Metastasis
History and Physical Examination
Laboratory Tests/Biomarkers
Detection of Cells in Messenger RNA Using PCR
Radiologic Tests
CT Scans
PET Scans
The Impact of PET/CT Imaging on Clinical Management
Ultrasound
Brain MRI/Spinal MRI
Radionuclide Scans
Radiolabeled Monoclonal Antibodies
Pathologic Tests
Molecular Tests
Sites of Distant Metastases
Skin, Subcutaneous Tissues, and Distant Lymph Nodes
Lung, Pleura, and Mediastinum
Brain and Spinal Cord
Gastrointestinal Tract
Liver, Biliary Tract, and Spleen
Bone
Kidneys and Urinary Tract
Heart and Pericardium
Pancreas
Peritoneum and Mesentery
Endocrine Organs
Breast
Ovaries, Uterus, and Placenta
Testes and Penis
Oral Cavity, Pharynx, and Larynx
Eye and Orbit
Cutaneous Melanosis
Surveillance in Patients Who Have Localized Melanoma
Cross-References
References
Part VII: Management of Distant Metastases
Evolving Role of Chemotherapy-Based Treatment of Metastatic Melanoma
Introduction
Single-Agent Chemotherapy
Dacarbazine and Temozolomide
Nitrosoureas
Platinums
Microtubule Toxins
Combination Cytotoxic Chemotherapy
Nitrosourea Combinations
Taxane-Based Combination Therapy
Cisplatin-Based Combination Therapy
Biochemotherapy
Interferon Alfa-Based Biochemotherapy Regimens
Interleukin-2 and Dacarbazine or Temozolomide Combinations
Cisplatin and Interleukin-2-Based Biochemotherapy Regimens
Checkpoint Inhibitor-Based Biochemotherapy
Other Biochemotherapy Combinations
Chemotherapy and Antiangiogenic Therapy
Antiangiogenic Agents
Chemotherapy and Map-Kinase-Targeted Therapy
Conclusion
References
Targeted Therapies for BRAF-Mutant Metastatic Melanoma
Introduction
Overview of the MAPK Pathway
MAPK Pathway in Cancer
BRAF Inhibitors
BRAF Inhibition in Other Cancers
Toxicity of BRAF Inhibitors
MEK Inhibitors
Combined BRAF and MEK Inhibition
Resistance to MAPK Inhibition
Transcriptional Mediators of Resistance
Genetic Drivers of Resistance
Experimental Approaches and Future Directions
ERK Inhibitors
Dimer-Disrupting RAF Inhibitors
Targeted/Immune Combinations
Conclusions
Cross-References
References
Molecularly Targeted Therapy for Patients with BRAF Wild-Type Melanoma
Background
KIT Mutations and Genetic Alterations
NRAS Mutant Melanoma
Preclinical Strategies Targeting NRAS Melanoma
Aurora Kinase A and PPP6c
Polo-Like Kinase 1 (PLK1) and MEK Inhibition
Targeting SHP2 (PTPN11)
Autophagy
Epigenetic Approaches
IDH1
ARID 2
EZH2
BH3 Mimetics
Non-V600 BRAF Mutations (Alternate BRAF Mutations)
Fusion Receptor Kinases
NF1
RAC1
TP53
CDKN2A
Signal Inhibitors Can Modulate the Tumor Immune Microenvironment
CDK4/6 Inhibitors (Palbociclib, Ribociclib, and Abemaciclib)
MEK Inhibitors (Trametinib, Cobimetinib, and Binimetinib)
MEK Inhibition and Cancer Immunity in Melanoma
Discussion
References
Cytokines (IL-2, IFN, GM-CSF, etc.) Melanoma
Interleukin-2
Pharmacology of IL-2
Immunologic Activity of IL-2
Systemic Effects of IL-2
Clinical Efficacy of High-Dose IL-2 Alone
Clinical Efficacy of Alternate Doses, Routes and Schedules of IL-2 Alone
Clinical Efficacy of IL-2 Combined with Interferon Alpha
Clinical Efficacy of IL-2 Combined with Other Cytokines, Immune Modulators, Antibodies or Vaccines
Clinical Efficacy of IL-2 Combined with Targeted Therapy
Agents to Reduce the Toxicity of IL-2
Predictors of Clinical Response
Other Cytokines for Therapy of Metastatic Melanoma
Summary
References
Checkpoint Inhibitors in the Treatment of Metastatic Melanoma
Background
Immune Checkpoints
CTLA-4 Inhibitors
PD-1 Inhibitors
Other Immune Checkpoint Inhibitors
Combination Immune Checkpoint Therapy
CTLA-4 and PD-1
Other Immune Checkpoint Combinations
Primary Resistance to Immune Checkpoint Inhibitors
Antigen Presentation and T-Cell Activation
Tumor Immunogenicity
Antigen Presentation
Costimulatory Signals
T-Cell Migration and Tumor Infiltration
Tumor Infiltration by Activated Cytotoxic T-Cells
The Tumor Microenvironment
Indoleamine 2,3-Dioxygenase
Vascular Endothelial Growth Factor
Tumor Cell Intrinsic Mechanisms of Creating a Resistant Microenvironment
Regulatory T-Cells
Myeloid-Derived Suppressor Cells
Adenosine
IPRES Signature Expression
Gut Microbiome
Secondary Resistance to Immunotherapy
Immunoediting
Antigen Presentation
Alternative Checkpoint Proteins
The Future of Checkpoint Inhibition in Melanoma
Cross-References
References
Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma
Introduction
Immune Checkpoint History
The T-Cell-Inflamed Tumor Microenvironment
Novel Therapeutics to Enhance a T-Cell-Inflamed Tumor Microenvironment
The Non-T-Cell-Inflamed Tumor Microenvironment
Oncogene Pathways of Immune Exclusion
Direct Activation in the Non-T-Cell-Inflamed Tumor Microenvironment
Novel Checkpoint Therapies
4-1BB
GITR/GITRL
LAG3
TIGIT
Intra-tumoral Therapy
Oncolytic Virus Therapy
Non-oncolytic Virus Therapy
Toll-Like Receptor (TLR) Agonists
Stimulator of Interferon Gene (STING) Agonists
PV-10 (Rose Bengal)
Cytokine
IL-2
IL-15
IL-10
Tumor Antigens
Tumor Vaccines
Antibody Drug Conjugates
Bispecific Antibodies and Bifunctional Fusion Proteins
Adoptive Cell Therapy (ACT)
Chimeric Antigen Receptor (CAR) T-Cell Therapy
Conclusion
References
Managing Checkpoint Inhibitor Symptoms and Toxicity for Metastatic Melanoma
Introduction
Meta Analyses: Gastrointestinal irAEs
Colitis
Enteritis
Esophagitis/Gastritis
Pancreatitis
Meta-analyses: Neurologic irAEs
Encephalitis
Peripheral Neuropathies
Cranial Neuropathies
Guillain-Barré Syndrome
Myasthenia Gravis
Cerebellar Ataxia
Transverse Myelitis
Management and Work-Up of Neurologic irAEs
Pneumonitis
Sarcoidosis
Meta-analysis of Hematologic irAEs
Thrombocytopenia
Leukopenia
Anemia
Pan-cytopenia
Meta-analysis of Dermatologic Adverse Effects
Dermatitis
Toxic Epidermal Necrolysis (TEN)
Management and Work-Up of Dermatologic Adverse Events
Meta-analysis of Autoimmune Hepatitis
Hepatitis
Management and Work-Up of Autoimmune Hepatitis
Meta-analysis of Endocrinopathies
Thyroid Dysfunction
Hypophysitis/Adrenal Insufficiency
Type I Diabetes
Management and Work-Up of Endocrinopathies
Nephritis
Management and Work-Up of Nephritis
Myocarditis
Management and Work-Up of Myocarditis
Arthropathies
Management and Work-Up of Arthropathies
Conclusions
References
Sequencing and Combinations of Molecularly Targeted and Immunotherapy for BRAF-Mutant Melanoma
Introduction
Targeted Therapy
Vemurafenib and Cobimetinib
Dabrafenib Plus Trametinib
Encorafenib with Binimetinib
Activity Profiles for BRAF/MEK Inhibitor Combinations
Immunotherapy: Checkpoint Inhibitors
Ipilimumab
Anti-PD-1/PD-L1 Agents
Combined CTLA-4 and PD-1 Blockade
Characterization of Response Patterns to Immunotherapy
Differential Activity of Immunotherapy in Patients with BRAF-Mutant Versus BRAF Wild-Type Melanoma
Sequencing of BRAF/MEK Inhibition and Checkpoint Blockade
Combined Targeted Therapy and Immunotherapy
Combined BRAF/MEK Inhibition and Checkpoint Blockade
Conclusion
References
Melanoma Vaccines
Introduction
Melanoma Vaccine Studies Prior to Immune Checkpoint Blockade Treatments
Vaccines with Tumor-Associated Melanoma Antigens
Melanin Pathway Antigens
Cancer-Testis Antigens
Ganglioside Antigen Vaccines
Melanoma Vaccines in Development
Neoantigens from Mutated Proteins: The Basis for New Personalized Vaccines in Melanoma
Clinical Studies with Neoantigen Melanoma Vaccines
Basic Principles Involved in Melanoma Vaccine Design
Basics of Antigen Presentation
Getting the Right Effector Cells Against Melanoma
Properties of Vaccines Required for Effective Responses Against Melanoma?
The Size of the Inoculum and the Vaccine Platform Used
Broad Antigen and Cytokine Responses Are Better than Narrow Specific Responses
Boosting Responses with Co-stimulators
Overcoming T Cell Exhaustion
Vaccine Platforms and Adjuvants
Combining Cancer Vaccines with Immune Checkpoint Blockade
Cancer Vaccines Combined with Immune Checkpoint Blockade in Patients
Where Next for Melanoma Vaccines?
Selecting Patients Who May Benefit from Vaccine Treatment
Is the Type of Vaccine Important to Synergize with ICB Treatment?
Conclusion
References
Cellular Therapy for Melanoma
T Cells Mediate Recognition of Melanoma
Identifying Tumor-Associated Antigens and Stress Ligands Recognized by T Cells
Adoptive Cell Therapy Using Tumor Infiltrating Lymphocytes
ACT Using T Cells Transduced to Express Antitumor Receptors
The T-Cell Repertoire in TIL
The Future of ACT
References
Systemic Therapy for Mucosal, Acral, and Uveal Melanoma
Clinical Characteristics
Epidemiology
Natural History and Prognosis
Molecularly Targeted Therapy
Molecular Biology
Therapeutic Implications
Immunotherapy
Immunobiology
Therapeutic Implications
Other Therapeutic Options for Advanced Disease
Chemotherapy, Biotherapy, and Biochemotherapy
Other Management Strategies for Advanced Disease
Systemic Therapy in the Adjuvant Setting
Conclusions
References
Dermatological Complications of Systemic Therapies for Melanoma
Targeted Therapy with Kinase Inhibitors
BRAF Inhibitor Monotherapy (Vemurafenib, Dabrafenib, Encorafenib)
Photosensitivity
Keratinocytic Lesions
Verrucal Keratoses
Keratoacanthomas and Cutaneous Squamous Cell Carcinomas
Palmar-Plantar Erythrodysesthesia and Plantar Hyperkeratosis
Maculopapular Rash
Atypical Melanocytic Proliferations and New Primary Melanomas
Other Cutaneous AEs
Monotherapy with MEK Inhibitors (Cobimetinib, Trametinib, Binimetinib)
Papulopustular Eruption (``Acneiform Rash´´)
Xerosis Cutis and Pruritus
Other Cutaneous AEs
BRAF and MEK Combination Therapy (Vemurafenib and Cobimetinib, Dabrafenib and Trametinib, Encorafenib and Binimetinib)
Immunotherapy with Checkpoint Inhibitors
Anti-CTLA-4 Monotherapy
Rash
Pruritus
Anti-PD-1 Monotherapy
Rash
Pruritus
Lichenoid Reactions
Psoriasis
Vitiligo
Other Adverse Events
Combination of Anti-CTLA-4 and PD-1 Antibodies
Rash
Other Adverse Events
Conclusions
References
Surgical Management of Distant Melanoma Metastases
Introduction
Historical Perspective of the Role of Surgery in Patients with Stage IV Melanoma
Surgery with Curative Intent for Patients with Stage IV Melanoma
Indications for Surgical Resection of Distant Melanoma Metastases
Relationship Between Tumor Burden, Immune Function, and Surgical Treatment
Selection of Patients for Surgery and Prognostic Factors
Palliative Surgery for Distant Melanoma Metastases
Surgery for Localized Residual Disease After Systemic Therapy
The Value of Local Disease Control
Outcomes After Surgical Resection of Melanoma Metastases at Specific Sites
Skin, Subcutaneous Tissue, and Lymph Nodes
Lung, Trachea, and Bronchi
Brain and Spinal Cord
Gastrointestinal Tract
Liver
Biliary Tract
Pancreas
Spleen
Bone
Kidney and Urinary Tract
Heart and Pericardium
Endocrine Organs
Breast
Reproductive Organs
Oral Cavity and Pharynx
Eye and Orbit
Surgery in the Era of Effective Systemic Therapies
The Changing Role of Surgery in Patients with Stage IV Melanoma
Role of Imaging in Predicting Surgical Benefits
Conclusions
Cross-References
References
Radiotherapy for Distant Melanoma Metastases
Introduction
Soft Tissue Metastases
Bone Metastases
Brain Metastases
Single and Oligo-Brain Metastases
Adjuvant Whole Brain Radiation Therapy After Local Treatment of Single or Oligo Brain Metastases
The Role of Whole Brain Radiation Therapy in Multiple Metastases and Leptomeningeal Disease
Stereotactic Body Radiation Therapy for Extracranial Metastases
Combination of Systemic Drug Therapy with Radiation Therapy
References
Melanoma Brain Metastases: Unique Biology and Implications for Systemic Therapy
Introduction
Biology and Immunology of MBM
The Structure and Function of the Blood-Brain Barrier
Molecular Biology of Melanoma Brain Metastases
Immunity of the Central Nervous System
The Immune Microenvironment in the Setting of Brain Metastasis
Systemic Therapy for MBM
Targeted Therapy
Immunotherapy
Radiation Therapy for MBM
Whole Brain Radiation Therapy (WBRT)
Stereotactic Radiotherapies/Radiosurgery
Combinatorial Approaches
Combining Immunotherapy and Targeted Therapy
Targeted Therapy in Combination with Stereotactic Radiosurgery
Immunotherapy in Combination with Stereotactic Radiosurgery
Key Challenges and Opportunities
Assessment of Clinical Responses in MBM Patients
Pseudoprogression and Radiation Necrosis
Corticosteroids
Considerations for Future Clinical Trial Design
Leptomeningeal Disease (LMD)
Multidisciplinary Approaches to Management of CNS Metastases
Conclusion
References
Part VIII: History of Melanoma
A History of Melanoma: From Hunter to Morton
Introduction
John Hunter (1728-1793)
Rene Laennec (1781-1826)
William Norris (1792-1877)
Other Early Nineteenth Century Descriptions
Contemporary Histories of Melanoma
Excision of Lymph Node Metastases
Oliver Pemberton (1825-1897)
James Paget (1814-1899)
Sir Jonathan Hutchinson (1828-1913)
Other Late Nineteenth Century Descriptions of Melanoma
Early Twentieth Century Reports
William Sampson Handley (1872-1962)
Alexander Breslow (1928-1980)
Vincent J. McGovern (1915-1983)
Sophie Spitz (1910-1956) And Arthur Allen (1910-1994)
Wallace H. Clark, Jr. (1924-1997)
Thomas B. Fitzpatrick, Jr. (1919-2003)
Gerald W. Milton (1924-2007)
Seng-Jaw Soong (1943-2012)
Donald L. Morton (1934-2012)
References
Index
Recommend Papers

File loading please wait...
Citation preview

Charles M. Balch Michael B. Atkins Claus Garbe Jeffrey E. Gershenwald Allan C. Halpern John M. Kirkwood Grant A. McArthur John F. Thompson Arthur J. Sober  Editors

Cutaneous Melanoma Sixth Edition

Cutaneous Melanoma

Charles M. Balch • Michael B. Atkins Claus Garbe • Jeffrey E. Gershenwald Allan C. Halpern • John M. Kirkwood Grant A. McArthur • John F. Thompson Arthur J. Sober Editors

Cutaneous Melanoma Sixth Edition

With 371 Figures and 115 Tables

Editors Charles M. Balch Department of Surgical Oncology The University of Texas MD Anderson Cancer Center Houston, TX, USA

Michael B. Atkins Department of Oncology Georgetown University Medical Center Lombardi Comprehensive Cancer Center Washington, DC, USA

Claus Garbe Centre for Dermatooncology Department of Dermatology Eberhard Karls University Tuebingen, Germany

Jeffrey E. Gershenwald Departments of Surgical Oncology and Cancer Biology Melanoma and Skin Center The University of Texas MD Anderson Cancer Center Houston, TX, USA

Allan C. Halpern Dermatology Service, Department of Medicine Memorial Sloan Kettering Cancer Center New York, NY, USA

John M. Kirkwood Departments of Medicine, Dermatology, and Translational Science University of Pittsburgh and UPMC Hillman Cancer Center Pittsburgh, PA, USA

Grant A. McArthur Divisions of Cancer Medicine and Research Peter MacCallum Cancer Centre Melbourne, VIC, Australia

John F. Thompson Melanoma Institute Australia Faculty of Medicine and Health The University of Sydney Sydney, NSW, Australia

Arthur J. Sober Department of Dermatology Massachusetts General Hospital Harvard Medical School Boston, MA, USA

ISBN 978-3-030-05068-9 ISBN 978-3-030-05070-2 (eBook) ISBN 978-3-030-05069-6 (print and electronic bundle) https://doi.org/10.1007/978-3-030-05070-2 1st and 2nd edition: © J. B. Lippincott Company, Philadelphia 1985, 1991 3rd and 4th edition: © Quality Medical Publishing Inc., St. Louis, MO 1997, 2003 5th edition: © Quality Medical Publishing Inc., St. Louis, MO. Transfer: copyright transferred from the editors and individual contributors by written agreement 2009 6th edition: © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

To my wife, Carol, and our children, Glen, Alan, Laura, and Mark C.M.B. To my wife, Susan Crockin, and our children, Ben, Melea, and Jon M.B.A. To my wife, Eva, and my daughters, Anna, Lisa, and Sophie C.G. To my wife, Donna, and my children, Sophie and Ellie J.E.G. To my wife, Judy, and our children, Josh, Orly, and Deb A.C.H. To my wife, Gayle, and my children, Gerrit and Gordon J.M.K. To my wife, Maree, and our children, Sarah, Ethan, and Chloe G.A.M. To my children, Sally, James, Julia, and Anna J.F.T. To my wife, Cheryl, and our children, Felicia and Stephanie A.J.S.

To the Memory of: Seng-jaw Soong Ph.D. (1942–2012) Statistician extraordinaire and Editor of the first five editions of Cutaneous Melanoma

Seng-jaw Soong with Charles Balch (circa 1985) Donald L. Morton, M.D. (1934–2014) Consummate melanoma investigator and pioneer of the technique of lymphatic mapping and sentinel node biopsy for melanoma

Preface to the Sixth Edition

For over 34 years, we have been privileged to write six editions of a comprehensive text describing the biology and clinical management of melanoma. Cutaneous Melanoma has now become the authoritative and comprehensive treatise on this subject, which is used worldwide. It has a special emphasis on data presentation and collectively incorporates the clinical outcomes of more than 50,000 patients treated at major melanoma centers throughout the world. This book spans the entire spectrum of the disease, from precursors of melanoma to advanced stages of metastatic disease, from melanoma genes to population-based epidemiology, and from prevention of melanoma to all forms of multidisciplinary treatments. It incorporates the basic principles of diagnosis and pathologic examination with treatment approaches for the disease’s many clinical presentations. Each facet of clinical management is supported by statistical data about natural history, prognosis, and treatment results. During the past decade, there have been major advances in the clinical management of melanoma, most notably in the effectiveness of new immunotherapy and targeted therapies. As a result, this sixth edition has been completely rewritten by acknowledged experts in a wide variety of disciplines. Cutaneous Melanoma has grown considerably larger since the last edition, reflecting the vast amount of new information available on the topic and the increased incidence of melanoma throughout the world. The book contains 52 chapters and is organized into 7 major parts focusing on biology, diagnosis and staging, pathology, epidemiology and prevention, management of localized melanoma, management of regional metastases, diagnosis and treatment of distant metastases, systemic treatment of metastatic disease, and history of melanoma. The chapters are either completely new or extensively revised and updated. These chapters provide the latest data on melanoma staging and prognosis as well as on randomized prospective clinical trials involving surgical treatment and systemic treatments of melanoma. There are new chapters on clinical genetics, sentinel lymph node and other regional metastases, adjuvant systemic therapy, immune modulators, melanoma-specific targeted therapies, management of central nervous systemic metastases, and biomarkers. Chapters are also devoted to special types of melanomas, such as childhood melanomas, mucosal melanomas, and subungual melanomas. Finally, every clinical variation of recurrent and metastatic melanoma is addressed, with treatment options included for each clinical entity. These vii

viii

Preface to the Sixth Edition

topics include detailed descriptions about local recurrences and their management, regional recurrences, in-transit metastases, and specific sites of distant metastases (e.g., lung, gastrointestinal tract). There is also an emphasis on prevention, screening, and early detection of melanoma. Cutaneous Melanoma is oriented to all health-care providers and researchers involved in the care of patients with melanoma, including surgeons, dermatologists, medical and surgical oncologists, radiation oncologists, and other physicians and nurses who diagnose and treat patients with melanoma. There are valuable sections for geneticists, cell biologists, and immunologists as well as epidemiologists and public health workers to provide them with current reviews by the experts in those fields. We have attempted to present a balanced perspective of the risks and benefits involved in each treatment modality and how these can be optimized in various clinical settings. This book also provides a reliable reference for general physicians, internists, and other health-care workers to help them recognize cutaneous melanomas at the earliest possible stage, learn how to biopsy them properly for both diagnosis and micro-staging, and clinically manage these patients during and after treatment. Cutaneous Melanoma represents the accumulated wisdom of our colleagues throughout the world who have contributed significantly to the amazing progress that is being made in the study and treatment of melanoma. This progress is built on the solid foundation that was laid by the pioneers in the field, whose interest in this topic led to many of the discoveries that have been made to date. It is our hope that this book will help to further the understanding of melanoma, its biologic and clinical characteristics, and lead to further advances in treatment. We have been sustained throughout our professional careers by the courage and warmth of the large number of patients for whom it has been our privilege to care. Personal rewards from these relationships and a desire to help improve the care of other patients are the main reasons we continue to update this book. Houston, USA Washington, USA Tuebingen, Germany Houston, USA New York, USA Pittsburgh, USA Melbourne, Australia Sydney, Australia Boston, USA December 2019

Charles M. Balch Michael B. Atkins Claus Garbe Jeffrey E. Gershenwald Allan C. Halpern John M. Kirkwood Grant A. McArthur John F. Thompson Arthur J. Sober

Acknowledgments

So many colleagues and coworkers have contributed to this book that it is difficult to select only a few for specific recognition. This sixth edition of Cutaneous Melanoma was edited to a greater degree than any other edition. For this we thank all the authors and coauthors for their cooperation and willingness to make this edition even more consistent and integrated in its contents and treatment recommendations. Since publication of the fifth edition in 2009, there have been revolutionary changes in melanoma diagnosis and therapy, making this sixth edition the most extensively updated of all editions (since 1985) and a must read for everyone interested in current information on melanoma management. We also want to thank those in our academic offices who helped us in so many ways to facilitate the manuscript flow and editorial review process, as well as typing our manuscripts. These include Ms. Brigitte Nelson (for Dr. Balch), Ms. Kaye Oakley and Dr. Gabrielle Williams (for Dr. Thompson), Clara Florentino (for Dr. Sober), Ms. Julia Tijerina (for Dr. Atkins), Peg Grossman (for Dr. Halpern), Catherine Ringin (for Dr. McArthur), Lisa Huntley and Victoria Alisasis (for Dr. Kirkwood), and Irma Wintle (for Dr. Gershenwald). We are privileged to work with a dynamic and talented publishing team at Springer Publishing, led by Andrew Spencer, Senior Editor, who coordinated all aspects of this book, and Purva Ashok Kumar who was the point person on the production team. The Project Manager for this book was Kate Emery, who worked diligently and patiently with the authors on their chapter submissions. She was particularly valuable in coordinating the flow of all the manuscripts among the multiple authors and editors and functioning as the primary interface with the publisher. The research and data management represented in this book would not have been possible without the financial support and generous assistance of numerous institutions and individuals: NCI Center Core Grants and (P30 CA051008) to Georgetown Lombardi Comprehensive Cancer Center, The Marion Gardner Jackson Trust, and the Jason Sabbag Melanoma Research Fund (Massachusetts General Hospital), and support from the NCI Cancer Center Core Grant and the NIH SPORE grants in melanoma at The University of Texas MD Anderson Cancer Center (MD Anderson), Houston, Texas (1P50CA221703), and the University of Pittsburgh SPORE in Melanoma and Skin Cancer (CA121973); the Robert and Lynne Grossman Family Foundation and the Michael and Patricia Booker Melanoma Research Endowment (MD ix

x

Anderson), the Sandra and Thomas Usher/Frances and James McGlothlin endowments (University of Pittsburgh Cancer Institute at UPMC Hillman); and grants from the National Health and Medical Research Council of Australia and the Victorian Cancer Agency. The Melanoma Institute Australia database and its data managers have been supported by the Melanoma Foundation of the University of Sydney. Our families have provided vital emotional support and graciously forfeited some of their time with us to the additional effort required to complete this book. As only they know, considerable personal sacrifices were required that they willingly made. For this reason, we have chosen to dedicate this book to them. This edition is also dedicated to the memory of Drs. Donald L. Morton and Seng-jaw Soong. Dr. Morton was a visionary leader as a clinical investigator in the field of melanoma, a consummate clinician, and a dedicated teacher. He was a mentor and a friend to many of us. He contributed enormously to the field over the years (see chapter “A History of Melanoma: From Hunter to Morton” for more details). Dr. Soong brought a unique statistical dimension to the entire field of melanoma research and was an editor of in all five previous editions of Cutaneous Melanoma. He provided a valuable, meticulously assembled statistical database that facilitated our understanding of the biology and natural history of melanoma as well as treatment outcomes. Their legacy lives on through the pages of this melanoma book, and we will miss them both greatly. Their impact will last for many years to come.

Acknowledgments

Contents

Volume 1 Part I

Biology and Immunology of Melanoma . . . . . . . . . . . . . . .

1

..............

3

Immunology of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Paul F. Robbins and Yong-Chen Lu

41

Biomarkers for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dirk Schadendorf, Keith T. Flaherty, Lyn M. Duncan, Mohammed Kashani-Sabet, and Selma Ugurel

73

Biology of Melanocytes and Primary Melanoma M. Raza Zaidi, David E. Fisher, and Helen Rizos

Part II

Diagnosis and Staging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

105

Clinical Presentations of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . 107 Allan C. Halpern, Ashfaq A. Marghoob, Arthur J. Sober, Victoria Mar, and Michael A. Marchetti Dermoscopy/Confocal Microscopy for Melanoma Diagnosis . . . . . 145 Katie J. Lee, Nicola di Meo, Oriol Yélamos, Josep Malvehy, Iris Zalaudek, and H. Peter Soyer Biopsy of Suspected Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 Noah Smith, Timothy M. Johnson, John W. Kelly, Arthur J. Sober, and Christopher Bichakjian Lymphoscintigraphy in Patients with Melanoma . . . . . . . . . . . . . . 205 Roger F. Uren, Omgo E. Nieweg, and John F. Thompson Biopsy of the Sentinel Lymph Node . . . . . . . . . . . . . . . . . . . . . . . . . 239 Mark B. Faries, Alistair J. Cochran, Michael McLemore, Vernon K. Sondak, Sandra L. Wong, and John F. Thompson Melanoma Prognosis and Staging . . . . . . . . . . . . . . . . . . . . . . . . . . 271 Emily Z. Keung, Charles M. Balch, John F. Thompson, John M. Kirkwood, Richard A. Scolyer, Vernon K. Sondak, and Jeffrey E. Gershenwald xi

xii

Contents

Models for Predicting Melanoma Outcome . . . . . . . . . . . . . . . . . . . 299 Lauren E. Haydu, Phyllis A. Gimotty, Daniel G. Coit, John F. Thompson, and Jeffrey E. Gershenwald Part III Pathology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

315

Classification and Histopathology of Melanoma . . . . . . . . . . . . . . . 317 Richard A. Scolyer, Victor G. Prieto, David E. Elder, Alistair J. Cochran, and Martin C. Mihm Jr. Molecular Pathology and Genomics of Melanoma . . . . . . . . . . . . . 381 Klaus Georg Griewank, Rajmohan Murali, and Thomas Wiesner Part IV

Epidemiology and Prevention of Melanoma . . . . . . . . .

423

Clinical Epidemiology of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . 425 Catherine M. Olsen and David C. Whiteman Molecular Epidemiology of Melanoma . . . . . . . . . . . . . . . . . . . . . . 451 Anne E. Cust, Hensin Tsao, Marianne Berwick, Graham J. Mann, and Mark M. Iles Clinical Genetics and Risk Assessment of Melanoma . . . . . . . . . . . 471 V. Bataille, Hensin Tsao, S. Raimondi, and S. Gandini Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma . . . . . . . . . . . . . . . . . . 501 Cristian Navarrete-Dechent, Alon Scope, Hensin Tsao, Nadeem G. Marghoob, Arthur J. Sober, and Ashfaq A. Marghoob Melanoma Prevention and Screening . . . . . . . . . . . . . . . . . . . . . . . . 525 Susan M. Swetter, Alan C. Geller, Sancy A. Leachman, John M. Kirkwood, Alexander Katalinic, and Jeffrey E. Gershenwald Part V Management of Primary Melanoma and Locoregional Metastases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

571

Treatment of Primary Melanomas . . . . . . . . . . . . . . . . . . . . . . . . . . 573 John F. Thompson, Michael A. Henderson, Gabrielle Williams, and Merrick I. Ross Reconstructive Options Following Surgery of Primary Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 595 Marc Moncrieff, Brian Gastman, Rogerio Izar Neves, Howard Peach, and Anthony P. Tufaro Axillary and Epitrochlear Lymph Node Dissection for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657 Sandra L. Wong, Douglas S. Tyler, Charles M. Balch, John F. Thompson, and Kelly M. McMasters

Contents

xiii

Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . 669 Keith A. Delman, Lesly A. Dossett, Clara R. Farley, Kelly M. McMasters, and Omgo E. Nieweg Neck Dissection and Parotidectomy for Melanoma . . . . . . . . . . . . . 689 Brian Gastman, Rebecca Knackstedt, Ryan P. Goepfert, Baran Sumer, Ashok Shaha, and Michael E. Kupferman Local and Recurrent Regional Metastases of Melanoma . . . . . . . . 705 Matthew C. Perez, Kenneth K. Tanabe, Charlotte E. Ariyan, John T. Miura, Dorotea Mutabdzic, Jeffrey M. Farma, and Jonathan S. Zager Radiotherapy for Primary and Regional Melanoma Angela M. Hong and Graham Stevens

. . . . . . . . . . . 739

Adjuvant Systemic Therapy for High-Risk Melanoma Patients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 747 Yana G. Najjar, Ryan Massa, Vernon K. Sondak, Alexander M. M. Eggermont, Helen Gogas, and John M. Kirkwood Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 767 Emily Z. Keung, Rodabe N. Amaria, Vernon K. Sondak, Merrick I. Ross, John M. Kirkwood, and Jennifer A. Wargo Hyperthermic Regional Perfusion for Melanoma of the Limbs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 795 Douglas S. Tyler, Douglas L. Fraker, Harald J. Hoekstra, and H. Richard Alexander Jr. Isolated Limb Infusion for Melanoma . . . . . . . . . . . . . . . . . . . . . . . 827 Georgia Marie Beasley, John T. Miura, Jonathan S. Zager, Douglas S. Tyler, John F. Thompson, and Hidde M. Kroon Surveillance and Follow-Up of Melanoma Patients . . . . . . . . . . . . 851 Rachael L. Morton, Anne Brecht Francken, and Mbathio Dieng Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes, and Selection of Treatment Options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 867 John F. Thompson, Nicola Mozzillo, and Merrick I. Ross

Volume 2 Part VI

Uncommon Presentations of Melanoma . . . . . . . . . . . . .

895

Acral Lentiginous Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897 Yukiko Teramoto, Hector Martinez-Said, Jun Guo, and Claus Garbe

xiv

Contents

Lentigo Maligna Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 925 Cristian Navarrete-Dechent, Kelly C. Nelson, Anthony M. Rossi, Erica H. Lee, Christopher A. Barker, Kishwer S. Nehal, and Susan M. Swetter Mucosal Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 953 Michael A. Henderson, Charles M. Balch, Claus Garbe, Alexander N. Shoushtari, Bin Lian, Chuanliang Cui, and Jun Guo Melanoma in Children and Teenagers . . . . . . . . . . . . . . . . . . . . . . . 969 Ines B. Brecht, Ira J. Dunkel, and Claus Garbe Pregnancy and the Use of Hormones in Melanoma Patients . . . . . 983 Alexandra Gangi, Robyn Saw, and Vernon K. Sondak Diagnosis of Stage IV Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . 997 Ahmad A. Tarhini, Sanjiv S. Agarwala, Arjun Khunger, Richard L. Wahl, and Charles M. Balch Part VII

Management of Distant Metastases

. . . . . . . . . . . . . . . . 1045

Evolving Role of Chemotherapy-Based Treatment of Metastatic Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1047 Sanjiv S. Agarwala, Mark R. Middleton, and Michael B. Atkins Targeted Therapies for BRAF-Mutant Metastatic Melanoma . . . . 1067 Douglas B. Johnson, Reinhard Dummer, Keith T. Flaherty, and Keiran S. Smalley Molecularly Targeted Therapy for Patients with BRAF Wild-Type Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1087 Sunandana Chandra, Grant A. McArthur, and Jeffrey Sosman Cytokines (IL-2, IFN, GM-CSF, etc.) Melanoma . . . . . . . . . . . . . . 1109 John B. A. G. Haanen, Ryan J. Sullivan, John M. Kirkwood, Michael B. Atkins, and Douglas J. Schwartzentruber Checkpoint Inhibitors in the Treatment of Metastatic Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1141 Alison Weppler, Peter Lau, and Grant A. McArthur Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma . . . . . . . . . . . . . . . . . . . 1165 Daniel J. Olson, Rodolfo Gutierrez, Salah Eddine Bentebibel, Randy F. Sweis, Omid Hamid, Adi Diab, Douglas B. Johnson, and Jason J. Luke Managing Checkpoint Inhibitor Symptoms and Toxicity for Metastatic Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187 Anna Pavlick and Jeffrey Weber Sequencing and Combinations of Molecularly Targeted and Immunotherapy for BRAF-Mutant Melanoma . . . . . . . . . . . . . . . . 1215 Paolo A. Ascierto and Michael B. Atkins

Contents

xv

Melanoma Vaccines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1243 Peter Hersey, Stuart J. Gallagher, John M. Kirkwood, and Jonathan Cebon Cellular Therapy for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1267 Udai S. Kammula and Michael T. Lotze Systemic Therapy for Mucosal, Acral, and Uveal Melanoma . . . . 1301 Suthee Rapisuwon, Yong Qin, Jason Roszik, Fernando Carapeto, Sapna Patel, and Richard D. Carvajal Dermatological Complications of Systemic Therapies for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337 Egle Ramelyte, Reinhard Dummer, Cristina Libenciuc, Gregory S. Phillips, Mario E. Lacouture, and Caroline Robert Surgical Management of Distant Melanoma Metastases John F. Thompson, Mark B. Faries, Erica B. Friedman, Jeffrey E. Lee, and Charles M. Balch Radiotherapy for Distant Melanoma Metastases Angela M. Hong and Christopher A. Barker

. . . . . . . . 1359

. . . . . . . . . . . . . . 1403

Melanoma Brain Metastases: Unique Biology and Implications for Systemic Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1421 Kim Margolin, Michael Davies, Harriet Kluger, and Hussein Tawbi Part VIII

History of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1455

A History of Melanoma: From Hunter to Morton . . . . . . . . . . . . . 1457 Arthur J. Sober, Charles M. Balch, John F. Thompson, and John M. Kirkwood Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1477

Editorial Board

Charles M. Balch Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Michael B. Atkins Georgetown Lombardi Comprehensive Cancer Center, Washington, DC, USA Department of Oncology, Georgetown University Medical Center, Washington, DC, USA Claus Garbe Centre for Dermatooncology, Department of Dermatology, Eberhard Karls University, Tuebingen, Germany Jeffrey E. Gershenwald Departments of Surgical Oncology and Cancer Biology, Melanoma and Skin Center, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Allan C. Halpern Dermatology Service, Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY, USA John M. Kirkwood Departments of Medicine, Dermatology, and Translational Science, University of Pittsburgh and UPMC Hillman Cancer Center, Pittsburgh, PA, USA Grant A. McArthur Divisions of Cancer Medicine and Research, Peter MacCallum Cancer Centre, Melbourne, VIC, Australia University of Melbourne, Parkville, VIC, Australia John F. Thompson Melanoma Institute Australia, Faculty of Medicine and Health, The University of Sydney, Sydney, NSW, Australia Department of Melanoma and Surgical Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia Arthur J. Sober Department of Dermatology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA

xvii

Contributors

Sanjiv S. Agarwala Department of Medical Oncology, St Luke’s University Hospital and Health Network, Easton, PA, USA Temple University School of Medicine, St. Luke’s University Hospital, Bethlehem, PA, USA H. Richard Alexander Jr. Departments of Surgery, Biochemistry and Molecular Biology and Pathology, Rutgers Cancer Institute of New Jersey, New Brunswick, NJ, USA Rodabe N. Amaria Department of Melanoma Medical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Charlotte E. Ariyan Memorial Sloan-Kettering Cancer Center, New York, NY, USA Paolo A. Ascierto Unit of Melanoma, Cancer Immunotherapy and Development Therapeutics, Istituto Nazionale Tumori IRCCS Fondazione G. Pascale, Naples, Italy Michael B. Atkins Georgetown Lombardi Comprehensive Cancer Center, Washington, DC, USA Department of Oncology, Georgetown University Medical Center, Washington, DC, USA Charles M. Balch Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Christopher A. Barker Department of Radiation Oncology, Memorial Sloan Kettering Cancer Center, New York, NY, USA V. Bataille Department of Twin Research and Genetic Epidemiology, King’s College London | KCL, London, UK Georgia Marie Beasley Department of Surgery, Duke University Medical Center, Durham, NC, USA Salah Eddine Bentebibel Department of Melanoma Medical Oncology, Division of Cancer Medicine, MD Anderson Cancer Center, The University of Texas, Houston, TX, USA

xix

xx

Marianne Berwick Department of Internal Medicine, University of New Mexico Cancer Center, University of New Mexico, Albuquerque, NM, USA Christopher Bichakjian Department of Dermatology, University of Michigan, Ann Arbor, MI, USA Ines B. Brecht Department of Pediatric Oncology and Hematology, Eberhard Karls University, Tuebingen, Germany Fernando Carapeto Department of Melanoma Medical Oncology, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Richard D. Carvajal Department of Medicine, Division of Hematology/ Oncology, Columbia University Medical Center; Herbert Irving Comprehensive Cancer Center, New York, NY, USA Jonathan Cebon Cancer and Neurosciences, Clinical Services Unit, Austin Health, Melbourne, VIC, Australia Cancer Immunobiology Program, Olivia Newton-John Cancer Research Institute, LaTrobe University, Heidelberg, VIC, Australia Sunandana Chandra Division of Hematology Oncology, Northwestern University Feinberg School of Medicine, Chicago, IL, USA Alistair J. Cochran Departments of Pathology, Laboratory Medicine and Surgery, David Geffen School of Medicine at UCLA and Jonsson Comprehensive Cancer Center, UCLA, Los Angeles, CA, USA Daniel G. Coit Department of Surgery, Memorial Sloan Kettering Cancer Center, New York, NY, USA Chuanliang Cui Department of Renal Cancer and Melanoma, Key Laboratory of Carcinogenesis and Translational Research, Ministry of Education, Peking University Cancer Hospital and Institute, Beijing, China Anne E. Cust Cancer Epidemiology and Prevention Research, Sydney School of Public Health, The University of Sydney, Sydney, NSW, Australia Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia Michael Davies Department of Melanoma Medical Oncology, Department of Systems Biology, Department of Translational Molecular Pathology, Department of Investigational Cancer Therapeutics, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Keith A. Delman Division of Surgical Oncology, Department of Surgery, Emory University School of Medicine, Atlanta, GA, USA Nicola di Meo Department of Dermatology and Venereology, University of Trieste, Trieste, Italy

Contributors

Contributors

xxi

Adi Diab Department of Melanoma Medical Oncology, Division of Cancer Medicine, MD Anderson Cancer Center, The University of Texas, Houston, TX, USA Mbathio Dieng NHMRC Clinical Trials Centre, The University of Sydney, Camperdown, NSW, Australia Lesly A. Dossett Division of Surgical Oncology, Department of Surgery, University of Michigan, Ann Arbor, MI, USA Reinhard Dummer Department of Dermatology, University Hospital Zurich, Zurich, Switzerland Department of Dermatology, University of Zurich, Zurich, Switzerland Lyn M. Duncan Department of Pathology, Massachusetts General Hospital and Harvard Medical School, Boston, MA, USA Ira J. Dunkel Department of Pediatrics, Memorial Sloan-Kettering Cancer Center, New York, NY, USA Alexander M. M. Eggermont Gustave Roussy Cancer Campus Grand Paris, Villejuif, France David E. Elder Department of Pathology and Laboratory Medicine, Hospital of the University of Pennsylvania, Philadelphia, PA, USA Mark B. Faries Cedars Sinai Medical Center, The Angeles Clinic and Research Institute, Los Angeles, CA, USA Clara R. Farley Division of Surgical Oncology, Department of Surgery, Emory University School of Medicine, Atlanta, GA, USA Jeffrey M. Farma Department of Surgical Oncology, Fox Chase Cancer Center, Philadelphia, PA, USA David E. Fisher Department of Dermatology, Harvard/MGH Cutaneous Biology Research Center, and Melanoma Program, MGH Cancer Center, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA Keith T. Flaherty Department of Medicine, Massachusetts General Hospital, Boston, MA, USA Cancer Center, Massachusetts General Hospital, Boston, MA, USA Douglas L. Fraker Department of Surgery, Division of Endocrine and Oncologic Surgery, Hospital of the University of Pennsylvania, Philadelphia, PA, USA Anne Brecht Francken Centre of Oncology, Isala, Zwolle, Netherlands Erica B. Friedman Division of Surgical Oncology, New York University, School of Medicine, NY, USA Stuart J. Gallagher Melanoma Research, Central Clinical School, Centenary Institute, University of Sydney, Camperdown, Sydney, NSW, Australia Melanoma Institute Australia, Crows Nest, NSW, Australia

xxii

S. Gandini Department of Experimental Oncology, European Institute of Oncology, Milan, Italy Alexandra Gangi Samuel Oschin Comprehensive Cancer Center, CedarsSinai Medical Center, Los Angeles, CA, USA Claus Garbe Centre for Dermatooncology, Department of Dermatology, Eberhard Karls University, Tuebingen, Germany Brian Gastman Department of Plastic Surgery, Cleveland Clinic, Lerner Research Institute, Cleveland, OH, USA Alan C. Geller Department of Social and Behavioral Sciences, Harvard TH Chan School of Public Health, Boston, MA, USA Jeffrey E. Gershenwald Departments of Surgical Oncology and Cancer Biology, Melanoma and Skin Center, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Phyllis A. Gimotty Department of Biostatistics, Epidemiology and Informatics, Perelman School of Medicine, The University of Pennsylvania, Philadelphia, PA, USA Ryan P. Goepfert Department of Head and Neck Surgery, UT MD Anderson Cancer Center, Houston, TX, USA Helen Gogas First Department of Medicine, Laiko Hospital, University of Athens, Athens, Greece Klaus Georg Griewank Department of Dermatology – University Hospital, University Duisburg-Essen, Essen, Germany Dermatopathologie bei Mainz, Nieder-Olm, Germany Jun Guo Department of Renal Cancer and Melanoma, Key Laboratory of Carcinogenesis and Translational Research, Ministry of Education, Peking University Cancer Hospital and Institute, Beijing, China Rodolfo Gutierrez Hematology and Oncology, Translational Research and Immunotherapy, Melanoma Therapeutics, The Angeles Clinic and Research Institute, Los Angeles, CA, USA John B. A. G. Haanen Antoni van Leeuwenhoek Hospital, Division of Medical Oncology, The Netherlands Cancer Institute, Amsterdam, The Netherlands Allan C. Halpern Dermatology Service, Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY, USA Omid Hamid Translational Research and Immunotherapy, Melanoma Therapeutics, The Angeles Clinic and Research Institute, Los Angeles, CA, USA Lauren E. Haydu Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Michael A. Henderson Division of Cancer Surgery, Peter MacCallum Cancer Centre, Melbourne, VIC, Australia

Contributors

Contributors

xxiii

Peter Hersey Melanoma Research, Central Clinical School, Centenary Institute, University of Sydney, Camperdown, Sydney, NSW, Australia Melanoma Institute Australia, Crows Nest, NSW, Australia Harald J. Hoekstra Surgical Oncology Department of Surgery, University of Groningen, Groningen, The Netherlands Angela M. Hong Radiation Oncology, Melanoma Institute Australia, The University of Sydney, North Sydney, NSW, Australia Mark M. Iles Leeds Institute for Data Analytics, University of Leeds, Leeds, UK Douglas B. Johnson Division of Hematology/Oncology Department of Medicine, Vanderbilt University Medical Center and Vanderbilt Ingram Cancer Center, Nashville, TN, USA Timothy M. Johnson Department of Dermatology, University of Michigan, Ann Arbor, MI, USA Udai S. Kammula Solid Tumor Cell Therapy Program, Division of Surgical Oncology, University of Pittsburgh, UPMC Hillman Cancer Center, Pittsburgh, PA, USA Mohammed Kashani-Sabet Center for Melanoma Research and Treatment, California Pacific Medical Center Research Institute, San Francisco, CA, USA Alexander Katalinic Department of Medicine, University Lübeck, Institute for Social Medicine and Epidemiology, Lübeck, Schleswig-Holstein, Germany John W. Kelly Victorian Melanoma Service, The Alfred Hospital, Melbourne, VIC, Australia Emily Z. Keung Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Arjun Khunger Department of Hematology and Oncology, Taussig Cancer Center, Cleveland Clinic, Cleveland, OH, USA John M. Kirkwood Departments of Medicine, Dermatology, and Translational Science, University of Pittsburgh and UPMC Hillman Cancer Center, Pittsburgh, PA, USA Harriet Kluger Internal Medicine, Medical Oncology; Skin Diseases Research Center; Dermatology, SPORE in Skin Cancer; Urology, Urologic Oncology; Yale Cancer Center, Yale Medical School, New Haven, CT, USA Rebecca Knackstedt Department of Plastic Surgery, Cleveland Clinic, Cleveland, OH, USA Hidde M. Kroon Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia Department of General Surgery, The University of Adelaide, Royal Adelaide Hospital, Adelaide, SA, Australia

xxiv

Michael E. Kupferman Department of Head and Neck Surgery, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Mario E. Lacouture Department of Medicine, Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA Peter Lau Peter MacCallum Cancer Centre, Melbourne, VIC, Australia Sancy A. Leachman Department of Dermatology, Oregon Health and Science University, Portland, OR, USA Erica H. Lee Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA Jeffrey E. Lee Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Katie J. Lee Dermatology Research Centre, The University of Queensland, The University of Queensland Diamantina Institute, Brisbane, QLD, Australia Bin Lian Department of Renal Cancer and Melanoma, Key Laboratory of Carcinogenesis and Translational Research, Ministry of Education, Peking University Cancer Hospital and Institute, Beijing, China Cristina Libenciuc Oncology Department, Dermatology Service, Gustave Roussy Cancer Campus, Villejuif-Grand Paris, France Michael T. Lotze Department of Surgery, Immunology, and Bioengineering, University of Pittsburgh, UPMC Hillman Cancer Center, Pittsburgh, PA, USA Yong-Chen Lu Surgery Branch, National Cancer Institute, National Institutes of Health, Bethesda, MD, USA Jason J. Luke Department of Medicine, University of Pittsburgh and UPMC Hillman Cancer Center, Pittsburgh, PA, USA Josep Malvehy Department of Dermatology, Melanoma Unit, Hospital Clínic de Barcelona, IDIBAPS, Universitat de Barcelona, Barcelona, Spain Graham J. Mann Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia Centre for Cancer Research, Westmead Institute for Medical Research, The University of Sydney, Sydney, NSW, Australia Victoria Mar Victorian Melanoma Service, Alfred Hospital, Melbourne, VIC, Australia Michael A. Marchetti Dermatology Service, Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY, USA Ashfaq A. Marghoob Dermatology Service, Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY, USA Nadeem G. Marghoob New York Institute of Technology College of Osteopathic Medicine, OMSIII, Old Westbury, New York, NY, USA

Contributors

Contributors

xxv

Kim Margolin Departments of Medical Oncology and Therapeutics Research, City of Hope National Medical Center, Duarte, CA, USA Hector Martinez-Said Melanoma Clinic, Surgical Oncology, Instituto Nacional de Cancerología, México City, Mexico Ryan Massa Hematology/Oncology, University of Pittsburgh Medical Center, Pittsburgh, PA, USA Grant A. McArthur Divisions of Cancer Medicine and Research, Peter MacCallum Cancer Centre, Melbourne, VIC, Australia University of Melbourne, Parkville, VIC, Australia Michael McLemore Department of Pathology and Laboratory Medicine, David Geffen School of Medicine at UCLA, Los Angeles, CA, USA Kelly M. McMasters Department of Surgery, University of Louisville School of Medicine, Louisville, KY, USA Mark R. Middleton Department of Oncology, University of Oxford, Oxford, UK Martin C. Mihm Jr. Dermatology Department, Brigham and Women’s Hospital, Harvard Medical School, Boston, MA, USA John T. Miura Department of Cutaneous Oncology, Moffitt Cancer Center, Tampa, FL, USA Marc Moncrieff Department of Plastic and Reconstructive Surgery, Norfolk & Norwich University Hospital, Norwich, UK Rachael L. Morton NHMRC Clinical Trials Centre, The University of Sydney, Camperdown, NSW, Australia Nicola Mozzillo Department Melanoma and Soft Tissue, Istituto Nazionale dei Tumori, Napoli, Italy Rajmohan Murali Department of Pathology, Memorial Sloan Kettering Cancer Center, New York, NY, USA Dorotea Mutabdzic Department of Surgical Oncology, Fox Chase Cancer Center, Philadelphia, PA, USA Yana G. Najjar Division of Hematology-Oncology, UPMC Hillman Cancer Center, Pittsburgh, PA, USA Cristian Navarrete-Dechent Melanoma and Skin Cancer Unit, Department of Dermatology, Facultad de Medicina, Pontificia Universidad Catolica de Chile, Santiago, Chile Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA Kishwer S. Nehal Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA

xxvi

Kelly C. Nelson Department of Dermatology, MD Anderson Cancer Center, The University of Texas, Houston, TX, USA Rogerio Izar Neves Department of Plastic Surgery, College of Medicine, Penn State University, Hershey, PA, USA Omgo E. Nieweg Melanoma Institute Australia, Department of Surgery, The University of Sydney Central Clinical School, Royal Prince Alfred Hospital, Department of Melanoma and Surgical Oncology, Sydney, NSW, Australia Catherine M. Olsen Cancer Control Group, QIMR Berghofer Medical Research Institute, Brisbane, QLD, Australia Faculty of Medicine, The University of Queensland, Brisbane, QLD, Australia Daniel J. Olson Department of Medicine, Section of Hematology/Oncology, The University of Chicago, Chicago, IL, USA Sapna Patel Department of Melanoma Medical Oncology, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Anna Pavlick Laura and Isaac Perlmutter Cancer Center, NYU Langone Health, New York, NY, USA Howard Peach Department of Plastic and Reconstructive Surgery, Leeds Teaching Hospitals NHS Trust, Leeds, UK Matthew C. Perez Department of Cutaneous Oncology, Moffitt Cancer Center, Tampa, FL, USA Gregory S. Phillips Department of Medicine, Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA Victor G. Prieto Departments of Pathology and Dermatology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Yong Qin Department of Melanoma Medical Oncology, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA S. Raimondi Department of Experimental Oncology, European Institute of Oncology, Milan, Italy Egle Ramelyte Dermatology Department, University Hospital Zurich, Zurich, Switzerland Suthee Rapisuwon Department of Medicine, Division of Hematology/ Oncology, Georgetown University; Lombardi Comprehensive Cancer Center, Washington, DC, USA Helen Rizos Department of Biomedical Sciences, Faculty of Medicine and Health Sciences, Macquarie University, and Melanoma Institute Australia, Sydney, NSW, Australia

Contributors

Contributors

xxvii

Paul F. Robbins Surgery Branch, National Cancer Institute, National Institutes of Health, Bethesda, MD, USA Caroline Robert Oncology Department, Dermatology Service, Gustave Roussy Cancer Campus, Villejuif-Grand Paris, France Université Paris-Sud, Kremlin Bicêtre, France Merrick I. Ross Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Anthony M. Rossi Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA Jason Roszik Department of Melanoma Medical Oncology, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Robyn Saw Melanoma Institute Australia, University of Sydney, Royal Prince Alfred Hospital, Sydney, NSW, Australia Dirk Schadendorf Department of Dermatology, University Hospital Essen, Essen, Germany German Cancer Consortium, Heidelberg, Germany Douglas J. Schwartzentruber Indiana University Health, Bloomington, IN, USA Richard A. Scolyer Tissue Pathology and Diagnostic Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia Melanoma Institute Australia, Central Clinical School, The University of Sydney, Sydney, NSW, Australia Alon Scope Medical Screening Institute, Sheba Medical Center, Ramat Gan, Israel Sackler Faculty of Medicine, Tel Aviv University, Tel Aviv, Israel Ashok Shaha Department of Otolaryngology, Memorial Sloan Kettering Cancer Institute, New York, NY, USA Alexander N. Shoushtari Memorial Sloan Kettering Cancer Center, Weill Cornell Medical College, New York, NY, USA Keiran S. Smalley Tumor Biology, H. Lee Moffitt Cancer Center and Research Institute, Tampa, FL, USA Noah Smith Department of Dermatology, University of Michigan, Ann Arbor, MI, USA Arthur J. Sober Department of Dermatology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA Vernon K. Sondak Department of Cutaneous Oncology, H. Lee Moffitt Cancer Center, Tampa, FL, USA

xxviii

Jeffrey Sosman Division of Hematology Oncology, Northwestern University Feinberg School of Medicine, Chicago, IL, USA H. Peter Soyer Dermatology Research Centre, The University of Queensland, The University of Queensland Diamantina Institute, Brisbane, QLD, Australia Dermatology Department, Princess Alexandra Hospital, Brisbane, QLD, Australia Graham Stevens Radiation Oncology, Orange General Hospital, Orange, NSW, Australia Bathurst Rural Clinical School, School of Medicine, Western Sydney University, Bathurst, NSW, Australia Ryan J. Sullivan Massachusetts General Hospital, Boston, MA, USA Baran Sumer Department of Otolaryngology, UT Southwestern Medical Center, Dallas, TX, USA Randy F. Sweis The University of Chicago, Department of Medicine, Section of Hematology / Oncology, Chicago, IL, USA Susan M. Swetter Department of Dermatology, Pigmented Lesion and Melanoma Program, Stanford University Medical Center and Cancer Institute, Stanford, CA, USA Dermatology Service, Veterans Affairs Palo Alto Health Care System, Palo Alto, CA, USA Kenneth K. Tanabe Division of Surgical Oncology, Department of Surgery, Massachusetts General Hospital Cancer Center, Boston, MA, USA Ahmad A. Tarhini Department of Cutaneous Oncology, H. Lee Moffitt Cancer Center and Research Institute, Tampa, FL, USA Hussein Tawbi Department of Melanoma Medical Oncology, Department of Systems Biology, Department of Translational Molecular Pathology, Department of Investigational Cancer Therapeutics, Division of Cancer Medicine, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Yukiko Teramoto Department of Skin Oncology/Dermatology, Saitama Medical University International Medical Center, Saitama, Japan John F. Thompson Melanoma Institute Australia, Faculty of Medicine and Health, The University of Sydney, Sydney, NSW, Australia Department of Melanoma and Surgical Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia Hensin Tsao Skin Cancer Genetics Laboratory/Wellman Center for Photomedicine, Melanoma and Pigmented Lesion Center/Department of Dermatology/Melanoma Genetics Program, Massachusetts General Hospital, Cancer Center, Harvard Medical School, Boston, MA, USA

Contributors

Contributors

xxix

Anthony P. Tufaro Department of Surgery, Division of Plastic Surgery and Surgical Oncology, The University of Oklahoma Health Science Center, Oklahoma City, OK, USA Douglas S. Tyler Department of Surgery, University of Texas Medical Branch, Galveston, TX, USA Department of Surgery, Duke University Medical Center, Durham, NC, USA Selma Ugurel Department of Dermatology, University Hospital Essen, Essen, Germany Roger F. Uren Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia Alfred Nuclear Medicine and Ultrasound, RPAH Medical Centre, Newtown, NSW, Australia Central Clinical Schools, The University of Sydney, Sydney, NSW, Australia Richard L. Wahl Mallinckrodt Institute of Radiology, Washington University School of Medicine, St. Louis, MO, USA Jennifer A. Wargo Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA Jeffrey Weber Laura and Isaac Perlmutter Cancer Center, NYU Langone Health, New York, NY, USA Alison Weppler Peter MacCallum Cancer Centre, Melbourne, VIC, Australia David C. Whiteman Cancer Control Group, QIMR Berghofer Medical Research Institute, Brisbane, QLD, Australia Thomas Wiesner Department of Dermatology, Medical University of Vienna, Vienna, Austria Gabrielle Williams Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia Sandra L. Wong Department of Surgery, Geisel School of Medicine at Dartmouth-Hitchcock Medical Center, Lebanon, NH, USA Oriol Yélamos Department of Dermatology, Melanoma Unit, Hospital Clínic de Barcelona, IDIBAPS, Universitat de Barcelona, Barcelona, Spain Jonathan S. Zager Department of Cutaneous Oncology, Moffitt Cancer Center, Tampa, FL, USA M. Raza Zaidi Fels Institute for Cancer Research and Molecular Biology, Lewis Katz School of Medicine at Temple University, Philadelphia, PA, USA Iris Zalaudek Department of Dermatology and Venereology, University of Trieste, Trieste, Italy

Part I Biology and Immunology of Melanoma

Biology of Melanocytes and Primary Melanoma M. Raza Zaidi, David E. Fisher, and Helen Rizos

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4

Development of Melanoblasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Regulation of Specification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Regulation of Migration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 Regulation of Survival and Proliferation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 Differentiation of Melanocytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 Regulation of Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 Regulation of Survival . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 Melanomagenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . From Melanocyte to Melanoma: A Multistep Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Molecular Genetics: Early Lessons from Familial and Sporadic Melanoma . . . . . . . . . . . . Melanoma: A Consequence of Homeostatic Disruption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Melanoma: Cell of Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16 16 17 21 26

Melanoma and the Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Sun Exposure and Epidemiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

M. R. Zaidi (*) Fels Institute for Cancer Research and Molecular Biology, Lewis Katz School of Medicine at Temple University, Philadelphia, PA, USA e-mail: [email protected] D. E. Fisher Department of Dermatology, Harvard/MGH Cutaneous Biology Research Center, and Melanoma Program, MGH Cancer Center, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA e-mail: dfi[email protected] H. Rizos Department of Biomedical Sciences, Faculty of Medicine and Health Sciences, Macquarie University, and Melanoma Institute Australia, Sydney, NSW, Australia e-mail: [email protected] © Crown 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_42

3

4

M. R. Zaidi et al. Photobiology and Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

Abstract

Melanocytes make up only a tiny proportion of the skin cellular milieu but have a major impact on skin appearance as well as skin cancer risk by modulating skin pigmentation. Cutaneous melanocytes are derived from precursor cells called melanoblasts that originate from the neural crest of the developing embryo and migrate for long distances to their niches in the epidermis and hair follicles, where they differentiate into melanin pigment-producing mini-factories. Melanin is synthesized and packaged within melanosomes, which are lysosome-related organelles responsible for melanin trafficking through dendrites to interacting keratinocytes. Melanocyte development, migration, proliferation, and differentiation are regulated by a complex network of extrinsic and intrinsic signaling pathways, which are responsive to key signals, such as ultraviolet radiation that stimulates melanin production (tanning). Alterations in components of these pathways may lead to melanomagenesis. Here we present an overview of the genes and pathways that regulate different aspects of the biology of melanocytes as well as their transformation to melanoma. Keywords

Development of melanoblasts · Regulation of specification · Regulation of migration · Regulation of survival and proliferation · Differentiation of melanocytes · Regulation of differentiation · Regulation of survival · Melanomagenesis · From melanocyte to melanoma: a multistep process · Molecular genetics: early lessons from familial and sporadic melanoma · Melanoma: a consequence of homeostatic disruption · Melanoma: cell of origin · Melanoma and the environment · Sun exposure and epidemiology · Photobiology and melanoma

Introduction Melanocytes are highly specialized cells that produce and distribute melanins, which are high molecular weight pigmented biopolymers responsible for pigmentation in the skin, hair, eyes, and inner ear. Although melanocytes represent a small proportion of cells in these pigmented tissues, the pigments they produce have several diverse and important functions (Hearing and Leong 2006; Nordlund 2006; Prota 1992). Melanocytes destined for the skin, hair, and choroid of the eye originate from pluripotent neural crest cells during embryonic development. In contrast, melanocytes populating the retinal pigment epithelium are derived from the optic cup (Bharti et al. 2006) and are not considered further in this chapter. Epidermal melanocytes reside among the basal keratinocytes in an approximate ratio of 1:10 and transfer melanin via elongated dendrites that contact up to 40 overlying suprabasal keratinocytes. The melanin-carrying keratinocytes form a vital barrier against environmental assaults. Consequently, skin color is not defined solely by the melanocytes that produce the pigment but occurs in partnership with the pigment-carrying keratinocytes in the superficial skin layers. Given the intimate relationship between melanocytes and other cell types present in the surrounding microenvironment, the importance of microenvironmental cues in melanocyte development, function, and transformation to melanoma (melanomagenesis) is well established. In many respects, the process of melanomagenesis is integrally but inversely related to the process of melanocyte development, representing a return to a proliferative, dedifferentiated, and migratory phenotype. In fact, the same essential cascades of genes that regulate development and differentiation of melanocytes are involved in the growth and eventual metastasis of melanoma cells. This chapter provides an

Biology of Melanocytes and Primary Melanoma

overview of the processes critical to melanocyte development and pigmentation in mammals and describes the disruption of these processes during the transformation of melanocytes into melanoma cells.

Development of Melanoblasts A recent curation effort by the Loftus group at the National Human Genome Research Institute (USA) has compiled a list of 650 genes that are currently known to be involved in integumentary pigmentation in humans, mice, and/or zebrafish (Baxter et al. 2018). Among these, 128 genes have been validated to have associations with human pigmentary phenotypes, while others have only been shown to exhibit phenotypes in mice and zebrafish to date. Clearly, the mechanistic network involved in the process of pigmentation is highly complex, which is exemplified by the wide spectrum of protein complexes, signaling pathways, and functions represented by the 26 different protein classes featured among the products of the pigmentary genes (Baxter et al. 2018). Disruption of the functions of these genes leads to a variety of pigmentary disorders, which can result from perturbations in melanocyte development, differentiation, and regulation. Although the process of melanocyte development is a continuous one, it is generally considered in four discrete stages: specification, migration, survival, and proliferation.

Regulation of Specification Specification describes the process early in development when pluripotent neural crest cells, which exhibit all the key features of stem cells, begin to commit to becoming pigment cells (melanocytes) in adult tissues. Before their final specification, neural crest cells (which are specified at an earlier stage of development and originate from the dorsal edge of the neural tube) have the ability to develop into a number of distinct types of cells, including neurons, glial cells, cardiac cells, and pigment cells (Bronner and LeDouarin 2012). As one might expect, transcription factors and

5

intracellular signaling factors play major roles in determining which pathway of development a given neural crest cell will take. Regarding those destined to become melanocytes, the precursor cells, once specified, are called melanoblasts. Early during gestation (approximately 1 month in humans), neural crest cells destined to become melanoblasts begin to express several genes encoding transcription factors (Table 1), including PAX3, LEFl, and FOXD3 (Blake and Ziman 2005; Ignatius et al. 2008; Kos et al. 2001; Silver et al. 2006; Yasumoto et al. 2002). These transcription factors regulate downstream target genes that participate in specifying further development of melanoblasts; some of those downstream genes have been identified (Blake and Ziman 2005; Hornyak et al. 2001; Ignatius et al. 2008; Kos et al. 2001). Concurrently, factors that regulate the WNT signaling pathway (WNT1 and WNT3) are also expressed and thought to trigger melanocytic differentiation through direct transcriptional targeting of the MITF gene (Dorsky et al. 2000). The canonical WNT signaling pathway is activated by extracellular WNT proteins binding to the surface receptor complex which then regulates the redistribution of β-catenin to the nucleus, where it acts to modulate transcription through activities that involve the transcription factors LEF1 and MITF (Dunn et al. 2005; Garcia-Castro et al. 2002; Hari et al. 2012; Lewis et al. 2004; Liu et al. 2014; Takeda et al. 2003). The multiple roles of MITF and WNT signaling, as well as many other intracellular signaling pathways, at virtually all phases of melanocyte development and differentiation will be a recurring theme in the discussion to follow (Kawakami and Fisher 2017; Liu et al. 2014). At least two other transcription factors are thought to play roles in the early developmental process of specification, SOX9 (Meulemans and Bronner-Fraser 2004) and SNAI2/SLUG (LaBonne and Bronner-Fraser 1998). SOX9 has also been shown to regulate MITF expression (Passeron et al. 2007). While the transcription factor SOX10 is required for the maintenance and differentiation of neural crest cells as they migrate, SOX9 plays an essential role in neural crest generation (Guth and Wegner 2008). SNAI2

Bone morphogenetic protein 4

N-cadherin P-cadherin OB-cadherin

BMP2

BMP4

CDH2 CDH3 CDH11 Integrins, laminin, lectin fibronectin ADAMTS20

KIT ligand; stem cell factor Endothelin receptor type B

Endothelin 3 Dopachrome tautomerase

KITLG EDNRB

EDN3 DCT

KIT

SOX10 MITF

Bone morphogenetic protein 2

FOXD3 SNAI2 SOX9 WNT1 WNT3A CTNNB1

ADAM metallopeptidase with thrombospondin type I motif 20 SRY-box 10 Microphthalmia-associated transcription factor Kit proto-oncogene

Protein name Paired box 3 Lymphoid enhancer-binding factor 1 Forkhead box D3 Slug SRY-box 9 Wnt family member 1 Wnt family member 3A Catenin Beta 1

Gene PAX3 LEF1

Receptor tyrosine kinase Ligand for c-KIT G protein-coupled receptor Ligand for EDNRB Enzyme

Transcription factor Transcription factor

Metalloproteinase

Transcription factor Transcription factor Transcription factor Wnt signaling Wnt signaling Intracellular signaling, regulation of transcription Secreted ligand to TGF-β receptors Secreted ligand to TGF-β receptors Surface binding protein Surface binding protein Surface binding protein Extracellular matrix proteins

Function Transcription factor Transcription factor

X

X

X

X X X X X X

Specification X X

X X

X X

X

X X

X

X X X

X X

Migration

Table 1 Partial list of genes involved in determination of melanocyte development, differentiation, and malignancy

X X

X X

X

X X

X

X X

Proliferation X

X X

X X

X

X X

Survival X

X X

X X

X

X X

X X X X

X

Differentiation

X

X X

X

X X

X X X X

X

X

X

X

Melanoma X X

6 M. R. Zaidi et al.

Transcription factor AP-2 alpha Dickkopf 1

Fibroblast growth factor receptor 3 Basic fibroblast growth factor Melanocortin 1 receptor

Proopiomelanocortin B-cell CLL/lymphoma 2 Baculoviral IAP repeat containing 5

MET proto-oncogene

Hepatocyte growth factor

G protein subunit alpha Q

MYC proto-oncogene NRAS proto-oncogene

Checkpoint kinase 1 G protein subunit alpha 11

Tyrosinase Tyrosinase related protein 1

Premelanosome protein

Melan-A

Solute carrier family 45 member 2

G protein-coupled receptor 143

Agouti signaling protein

TFAP2A DKK1

FGFR3 FGF2 MC1R

POMC BCL2 BIRC5

MET

HGF

GNAQ

MYC NRAS

CHK1 GNA11

TYR TYRP1

PMEL

MLANA

SLC45A2

GPR143

ASIP

Transcription factor Wnt signaling antagonist Receptor for FGF Ligand for FGFR G protein-coupled receptor Ligand for MC1R Antiapoptotic regulator Antiapoptotic protein survivin Receptor tyrosine kinase, HGF receptor Scatter factor, ligand for c-MET Guanine nucleotide binding protein Transcription factor GTPase, intracellular signaling Serine/threonine kinase Guanine nucleotide binding protein Enzyme Enzyme, DHI/DHICA oxidase Melanosomal structural protein gp100 Chaperone-like for PMEL trafficking Solute carrier, transporter, cargo unknown G protein-coupled receptor for L-DOPA Antagonist for MC1R X X X

X

X

X

X

X X X

X X

X X

X X

X

X

X

X X

X

X

X

X

X

X X

X

X

X

X

X X X

X X

X (continued)

X X

X

X X

X

X

X

X X X

X X X

X X

Biology of Melanocytes and Primary Melanoma 7

Solute carrier family 24 member 5 Solute carrier family 7 member 11

RAS-associated protein 7A

RAS-associated protein 27

RAS-associated protein 38

Adaptor protein 3

Myosin VA

ATPase copper transporting alpha

Interferon regulatory factor 4 Transient receptor potential M7 Yin and Yang 1 Activating transcription factor 2 PIP3 dependent Rac exchange factor 1 Neural precursor cell expressed developmentally Down-regulated protein

SLC24A5 SLC7A11

RAB7A

RAB27

RAB38

AP3

MYO5A

ATP7A

IRF4 TRPM7 YY1 ATF2 PREX1

Function Solute carrier, transporter, cargo unknown Na-Ca exchange Cysteine/glutamate transporter GTPase, regulates melanosome transport GTPase, regulates melanosome transport GTPase, regulates TYRP1 trafficking Regulates TYR trafficking Regulates melanosome transport Copper transport, required for TYR activity Transcription factor Cation channel Transcription factor Transcription factor Guanine nucleotide exchange factor Focal adhesion protein

Specification

Migration

The order of the genes is based on their predominant function in the different stages of melanocyte development

NEDD9

Protein name Oculocutaneous albinism II

Gene OCA2

Table 1 (continued)

X X

Proliferation

X X X

Survival

X

X

X

X

X

X

X

X X

Differentiation X

X

X X X X

Melanoma

8 M. R. Zaidi et al.

Biology of Melanocytes and Primary Melanoma

has been shown, at least in lower organisms, to regulate the expression and function of SOX10 (Jiang et al. 1998; Sanchez-Martin et al. 2002). Bone morphogenetic proteins play critical roles in determining patterning during development, and these are closely regulated by the WNT signaling pathway; in that regard, BMP2 and BMP4 have been shown to be important to melanoblast specification (Jin et al. 2001). The expression of BMP2/4 is not specific to cells destined to become melanoblasts, and other factors that are specific to melanoblasts and regulate their specification will undoubtedly be identified in the future. The majority of studies defining the genes and developmental pathways involved in melanoblast specification have been performed in lower species, including mice, chicks, zebrafish, and amphibians, although results to date suggest that these pathways and genes are highly conserved in humans (as are the defects involved in pigmentary disorders) (Baxter and Pavan 2013; Liu et al. 2014; Pavan and Raible 2012). Cells expressing these developmental markers then undergo an epithelial-mesenchymal transition and begin to migrate dorsolaterally below the ectoderm (called early migration). Expression of these same markers (and others, as described in the “Differentiation” section) is found in melanocyte stem cells in adult tissues that serve as reservoirs of cells that can be recruited to become differentiated melanocytes by appropriate environmental cues (Buac and Pavan 2007; Nishimura 2011; Osawa et al. 2005).

Regulation of Migration As the precursor cells migrate (Fig. 1), they remain multipotent (i.e., they retain the capacity to become pigment cells, smooth muscle cells, neurons, and so on). Those destined to become melanoblasts continue to express SOX9 but now begin to express other genes that regulate their further development, including SOX10, MITF, KIT, and, a bit later, a melanogenic (pigmentation-inducing) enzyme called dopachrome tautomerase (DCT), although

9

expression of DCT is thought to result simply from the increased function of SOX10 and MITF and is not relevant to developmental processes (Guyonneau et al. 2004). Of the newly expressed genes, SOX10 and MITF are transcription factors, whereas KIT is a surface receptor that plays a critical role in regulating intracellular signaling in melanocytes at many levels of their development and differentiation. Although these transcription factors are not specific to the pigment cell precursors, MITF is differentially expressed in different cell lineages using alternative promoters and initial exons, which are alternatively spliced to perform distinct functions and activate distinct genes, with the form of MITF specific to pigment cells designated MITF-M (Harris et al. 2010; Kawakami and Fisher 2017) (Fig. 2), although other spliced isoforms (non-tissue-specific) are typically expressed in melanocytes as well as other cell types. MITF-M (hereafter, MITF) is often referred to as the master regulator of melanocytes, because it is a central focal point through which environmental signaling modulates the behavior of melanocytes at many levels (see the “Differentiation” section). Melanoblasts move in either a ventral or a dorsolateral direction, giving rise to the peripheral nervous system and endocrine cells or to melanocytes, respectively. The full complement of signals that regulate these processes not only involves genes expressed by melanoblasts but also factors produced in the local environment in the developing embryo. For example, KIT is a tyrosine kinase receptor on the plasma membrane that is regulated by the environmentally derived stem cell factor (SCF, encoded by the KITLG gene) ligand, which works through the MAP kinase signaling pathway to regulate the expression of genes critical for melanoblast migration and survival (WehrleHaller 2003). Most notable among KIT-regulated genes is MITF. Yet another receptor, the endothelin receptor (EDNRB), which works through the phospholipase C (PLC) signaling cascade, is also critical for regulating melanoblast migration and proliferation, and although that receptor is expressed on the surface of melanoblasts, it is activated by endothelins (EDN) produced in the local environment (Lee et al. 2003;

10

M. R. Zaidi et al.

Fig. 1 Stages of melanoblast development, melanocyte differentiation, and transformation to melanoma. Genes important to various stages are identified in the boxes, as

discussed in the text. (Adapted from Yamaguchi et al. 2007; Brenner and Hearing 2008)

Saldana-Caboverde and Kos 2010). One must always bear in mind that for appropriate pigmentation patterns to occur, the signaling for melanoblast migration to begin, and later to stop, must be carefully controlled. Some melanoblasts move only short distances in the developing embryo, whereas others need to migrate long distances to reach the extremities. Defects in many of the genes involved in migration and specification are associated with developmental pigmentary diseases, often eliciting hypopigmented areas in ventral regions. For example, deficiency of the Rac-specific guanine nucleotide exchange factor (GEF) P-Rex1, encoded by the PREX1 gene, causes a white belly and feet phenotype in Prex1-knockout mice, characterized by defective melanoblast migration (Lindsay et al. 2011).

Cell-cell interactions are critically important during migration; thus, it is no surprise that the expression of various proteins regulating such interactions (e.g., cadherins and components of the extracellular matrix [ECM], such as integrins, laminin, lectins, and fibronectin (Bonaventure et al. 2013; Haass et al. 2005; Herlyn et al. 2000; Jouneau et al. 2000; Pinon and Wehrle-Haller 2011)) is also closely involved in the regulation of melanoblast migration. Similarly, proteases associated with penetration of biologic membranes are also important to cell migration, and ADAMTS20 (a metalloproteinase) has been closely associated with melanoblast survival and migration during development, with mutations affecting this gene resulting in abnormal patterning in the skin (Rao et al. 2003; Silver et al. 2008). Studies in mice suggest that melanoblasts undergo

Biology of Melanocytes and Primary Melanoma

11

Fig. 2 Factors involved in interactions between melanocytes and neighboring keratinocytes and fibroblasts. (Ligands and receptors are discussed in the text)

proliferation before their initial migration and then cease proliferating while they migrate, after which they may undergo extensive proliferation once they have reached their final destination (Thomas and Erickson 2008; Wilkie et al. 2002). With respect to invasion and metastasis, migrating melanoblasts have the ability to cross the basement membrane and other biologic barriers to tissue penetration during development.

Regulation of Survival and Proliferation Once melanoblasts reach their intended destination, they must cease migrating, then survive and proliferate, and eventually differentiate. Because survival and proliferation occur concurrently at this time during development, most genes known to be involved play roles in both processes

(or at least, it is very difficult to distinguish their roles in those two processes), and they will be considered together in this section. The survival of differentiated melanocytes in the skin is discussed further on. Many of the genes mentioned above are also involved in regulating survival and/or proliferation (i.e., genes encoding MITF, SOX10, and PAX3), but additional genes act as “supporting actors,” including the transcription factor AP2, as well as additional signaling factors (basic fibroblast growth factor [bFGF, encoded by the FGF2 gene] and hepatocyte growth factor [HGF]) that join the previous list (EDN3 and KITLG). The KIT receptor and its ligand KITLG play important roles in regulating survival and proliferation, and mutations in either of these two genes are associated with pigmentary disorders of the skin, which can result in hypopigmentation (e.g., vitiligo and piebaldism) as well as in hyperpigmentation

12

(e.g., UVR melanosis and lentigo senilis), depending on whether their function is decreased or increased, respectively (Diwakar et al. 2008; Imokawa 2004). It is an interesting paradox that melanocytes are relatively fragile cells because of the inherent stresses they undergo, primarily as a result of their production of highly toxic intermediates during melanogenesis, yet they typically have extremely long biologic lives and seldom proliferate in adult tissues outside of the hair follicle. In fact, stimulation of their proliferation is a key to their malignant transformation, as discussed in the “Melanomagenesis” section later. Melanoblasts and melanocytes typically express AP2, BCL2, and survivin (BIRC5) (Braeuer et al. 2011; McGill et al. 2002; Nishimura et al. 2005; Raj et al. 2008), all of which are antiapoptotic, and thus prevent melanocytic cells from succumbing to the cytotoxicity of their microenvironment. Three new sets of ligands and receptors come into play at this point. HGF and its receptor MET also work via the MAP kinase pathway (with MITF contributing as a downstream effector) to influence melanoblast survival and proliferation (Hirobe et al. 2004; Kos et al. 1999). The FGF receptor and its ligand FGF2 also work through the MAP kinase pathway to regulate proliferation and survival (Beauvais-Jouneau et al. 1999; Ito et al. 1999). Similarly, DKK1, an inhibitor of WNT signaling, also participates in regulating the survival and proliferation of melanoblasts (and later, their differentiation) (Lin and Fisher 2007; Yamaguchi et al. 2007; Yamaguchi et al. 2008).

Differentiation of Melanocytes Regulation of Differentiation Human melanocytes reside at the dermal-epidermal junction of the skin and are interspersed among basal keratinocytes on the basement membrane. Differentiated melanocytes produce specialized lysosome-related organelles, termed melanosomes, which contain the enzymatic and structural components required for the synthesis of melanins (Dell’Angelica 2003; Park et al. 2009;

M. R. Zaidi et al.

Schiaffino 2010). Melanocytes then distribute those melanosomes to the overlying suprabasal keratinocytes via their elaborate network of dendrites and cell-cell contacts (see Fig. 2). Melanocytes and keratinocytes in the basal layer of the epidermis are highly stable cell populations that proliferate extremely slowly under normal circumstances. In contrast, keratinocytes in the upper layers of the epidermis proliferate relatively rapidly and differentiate further into highly keratinized layers that eventually form the stratum corneum at the surface of the skin. Their upward proliferation carries them toward the surface of the skin along with their ingested melanin where they form a critical barrier against the environment (including ultraviolet radiation [UVR]). In humans, keratinocytes move from the basal layer to the surface of the skin and are lost by desquamation in approximately 4–5 weeks, so it is a relatively rapid process. Because of that, transformation events in proliferating keratinocytes are not thought to pose a long-term risk for skin cancers, but rather it is the transformation of melanocytes or keratinocytes in the basal layer that have long-term consequences concerning skin cancer. Although the various types of melanins produced by melanocytes (eumelanins and/or pheomelanins) are important for the eventual characteristic color of the skin, it is also the distribution of those melanins in the more superficial layers of the skin that has major effects. Although melanocytes in lower species often produce only one type of melanin at a given time, in humans different mixtures of eumelanins and pheomelanins are typically produced on a constitutive basis (Ito and IFPCS 2003; Micillo et al. 2016). Melanocytes in other tissues (e.g., hair follicles, eyes, and inner ears) interact with their surrounding cells in distinct manners, yet the basic processes of melanin production and melanosome biogenesis are comparable, as are some of the factors that regulate melanogenesis. Factors that influence pigmentation at sites other than the skin are covered in several reviews (Bharti et al. 2006; Hubbard et al. 2010; Lavado and Montoliu 2006; Slominski et al. 2005; Sturm and Larsson 2009; Tobin 2008) and books (Hearing and Leong 2006; Nordlund 2006).

Biology of Melanocytes and Primary Melanoma

As noted earlier, 650 distinct loci are currently known to be involved in regulating pigmentation either directly or indirectly (Baxter et al. 2018). Many of those loci/genes affect developmental processes, but others regulate either differentiation/survival of melanocytes once in situ or regulate physiological factors that affect pigmentation. The remainder of this section will provide an overview of the latter two categories. Many of the pigment regulatory genes affect the biogenesis and/or function of melanosomes, the discrete membrane-bound organelles within which melanins are synthesized. Melanosomes, which are closely related to lysosomes, are within the family of lysosome-related organelles (LROs) and require a number of specific enzymatic and structural proteins to mature and become competent to produce melanin (Chi et al. 2006; Hu et al. 2007; Yamaguchi et al. 2007). Several reviews have elaborated on processes involved in melanosome biogenesis and on the specific functions of melanosomal proteins (Barral and Seabra 2004; Schiaffino 2010). Briefly, the critical enzymes include tyrosinase (TYR), TYR-related protein 1 (TYRP1), and dopachrome tautomerase (DCT), and mutations in their encoding genes dramatically affect the quantity and quality of melanins synthesized. Important structural proteins of melanosomes include PMEL (also known as gp100) and MLANA (MART-1), both of which are required for the structural maturation of melanosomes (Berson et al. 2001; Hoashi et al. 2005). Interestingly, most (perhaps all) of those melanosomal proteins have been identified as targets of the immune system in patients with melanoma (Kawakami et al. 2000; Sakai et al. 1997). Because melanosomes are LROs, there are many proteins involved in the sorting and trafficking of proteins to melanosomes, some of which affect all LROs (such as AP3 and BLOC components), whereas others (such as OCA2 and SLC45A2) affect only melanosomes, and mutations in any of those typically lead to inherited hypopigmentary disorders (Costin and Hearing 2007; Costin et al. 2003; Nordlund 2006; Toyofuku et al. 2002). As melanosomes mature and their constituent proteins are delivered, they become cargos themselves, carried by various molecular motors

13

(actinbased and myosin-based) from the perinuclear area to the cell periphery (Coudrier 2007; Wu and Hammer 2000), after which they are transferred to neighboring keratinocytes (Scott et al. 2003; Wu and Hammer 2014). Several critical components are involved in linking melanosomes to molecular motors responsible for their transport (e.g., RAB27A, melanophilin, and myosin Va), and mutations in those genes result in a pigmentary disease known as the Griscelli syndrome. Constitutive skin pigmentation Human skin color ranges from extremely fair and light to extremely dark, depending on racial or ethnic background, among other factors, but the density of melanocytes in a given geographic area of the body is virtually identical in all skin types (Tadokoro et al. 2003; Yamaguchi et al. 2004). Keratinocytes in lighter skin tend to cluster lesspigmented melanosomes above the nuclei (called capping), while in darker skin, greater numbers of pigmented melanosomes are distributed individually in keratinocytes, thus maximizing their absorption of light and their photoprotective value. Constitutive melanocyte density in the skin can be affected by the environment, such as by chronic UV radiation (which can increase melanocyte density threefold or fourfold), by secreted paracrine factors in the skin (Hirobe 2011; Scott et al. 2002) (see Fig. 2), and by toxic compounds such as hydroquinone (which antagonizes the melanogenesis enzyme tyrosinase but can also permanently destroy melanocytes in the skin.) Several types of inherited pigmentary disorders also affect melanocyte density (e.g., increased numbers in freckles or decreased numbers in vitiligo (Thong et al. 2003; Whiteman et al. 1999; Yoshida et al. 2007).) Epidermal melanocytes proliferate slowly, if at all, under normal circumstances, and they are quite resistant to apoptosis because of their expression of BCL2 and survivin (BIRC5) (McGill et al. 2002; Raj et al. 2008). Melanocyte density and differentiation are influenced by the environment, including UVR and factors secreted by neighboring keratinocytes and fibroblasts (see Fig. 2). For example, fibroblasts in the dermis of the palms/soles secrete high levels of DKK1,

14

which suppresses melanocyte growth and function by inhibiting the WNT/β-catenin signaling pathway (Yamaguchi et al. 2004; Yamaguchi et al. 2008). DKK1 inhibition of WNT signaling in melanocytes dramatically inhibits the melanogenic pathway, ranging from effects on transcription factors (e.g., MITF, PAX3, SOX9, and SOX10) to downstream melanogenic proteins. DKK1 also affects keratinocytes in overlying epidermis, reducing their uptake of melanin and inducing a thicker less-pigmented skin phenotype (Yamaguchi et al. 2008). One major determinant of pigment phenotype of the skin is the melanocortin 1 receptor (MC1R), a G protein-coupled receptor which regulates virtually all functional aspects of melanocytes, including dendricity, melanogenesis, and proliferation, among other things (Herraiz et al. 2017; Wolf Horrell et al. 2016). MC1R function (and dysfunction) has critical implications for the risk for skin cancer and is considered in detail in the “Melanomagenesis” section. MC1R function is regulated by peptide agonists encoded by the POMC gene (αMSH and ACTH) and by an antagonist, agouti signaling protein (ASIP). Activation of MC1R stimulates the melanogenic cascade leading to the synthesis of eumelanin, while ASIP can reverse those effects and favor the production of pheomelanin. αMSH and ACTH also upregulate expression of the MC1R gene, thus acting in a positive feedback loop. Although MC1R is closely involved with regulating skin pigmentation, it is not the only melanocytic receptor that functions in pigmentation. In fact, virtually all of the receptor/ligand pairs discussed above that participate in developmental processes also regulate mammalian pigmentation in adult tissues, including KITLG:KIT, CSF2:CSF2RA, HGF:MET, FGF2:FGFR, and EDN:EDNRB. Those ligands can originate from neighboring cells in the epidermis (keratinocytes) and the dermis (fibroblasts) and even from more distant locations, induced by environmental factors (e.g., UVR). As seen in the schematic outline in Fig. 2, virtually all known intracellular signaling pathways are involved in responses to these factors: receptors, the common effect of many of them being to regulate the function of MITF.

M. R. Zaidi et al.

In addition to regulation by a wide variety of receptor:ligands, melanin production is also modulated by multiple other factors within melanocytes. The SLC45A2 and OCA2 proteins are melanocyte-specific proteins with 12 transmembrane domains that play critical roles in the sorting/trafficking of TYR to melanosomes. If TYR is not successfully delivered to melanosomes, no melanin will be produced, even if the TYR is a functional enzyme (TYR can be abnormally degraded by proteasomes or secreted from melanocytes). SLC45A2 is an intracellular pump/ transporter that regulates ion transport across intracellular membranes which regulate the sorting of TYR and thus the biosynthesis of melanin. OCA2 acts on multiple levels to influence the melanosome, the melanogenic enzymes, pH, and glutathione metabolism (Ainger et al. 2017). Mutations in any of these proteins result in the abnormal secretion of TYR from melanocytes rather than normal sorting of TYR to melanosomes. Population studies have shown that polymorphisms of the SLC45A2 or OCA2 genes (together with polymorphisms of the MC1R gene) play major roles in determining the normal range of pigmentation in the hair, skin, and eyes (Ainger et al. 2017; Graf et al. 2005; Shriver et al. 2003). A close association between a polymorphism in a sodium:calcium exchanger (SLC24A5) and melanin content of the skin has also been described (Ginger et al. 2008; Lamason et al. 2005). Facultative skin pigmentation Facultative skin pigmentation is the term coined for increased skin color due to some type of physiologic regulation, the most obvious being UVR in what is commonly termed the tanning reaction (Eller and Gilchrest 2000; Tadokoro et al. 2005). Several studies have detailed the complex kinetics of responses of the skin to UVR, which result in tanning over the course of several weeks (Chen et al. 2014; Coelho et al. 2009). UVR is the most significant factor that influences human skin pigmentation and elicits several stages of increased skin color. Immediate pigment darkening occurs within minutes and persists for several hours, followed by persistent pigment darkening, which occurs within several hours and lasts for several

Biology of Melanocytes and Primary Melanoma

days (Young 2006). These rapid increases in skin color do not result from increased melanin synthesis but rather are due to the oxidation and polymerization of existing melanin and the redistribution of existing melanosomes. A slower process, termed delayed tanning, occurs several days after UVR exposure but requires more time since it involves the activation of melanocyte function and depends on increased melanin production. As might be expected, UVR induces skin pigmentation due to its effects on melanocytes and also via indirect effects on keratinocytes. Regarding effects on melanocytes, ultimately UVR elicits increased expression of MITF and the downstream melanogenic cascade (Miyamura et al. 2007; Yamaguchi and Hearing 2006). Keratinocytes in the skin respond to UVR by increasing their secretion of αMSH and ACTH as a result of the stimulation of POMC expression, which was shown to result from increased TP53 expression (Cui et al. 2007). Interleukin-1 (IL1) secretion by keratinocytes is also elicited by UVR and stimulates the autocrine secretion of ACTH, αMSH, EDN1, and FGF2 (Hirobe and Ootaka 2007). Each of those factors functions via receptors expressed on melanocytes, as discussed earlier. UVR can also affect fibroblasts in the dermis, and growth factors secreted from those cells in response to UVR include HGF, FGF2, and KITLG, each of which stimulates pigmentation via distinct receptors expressed by melanocytes (Imokawa 2004). The tanning response also relies on stimulation of secretion of nerve growth factor (NGF) by keratinocytes, which prevents apoptosis following UVR exposure (Botchkarev et al. 2006; Kadekaro et al. 2003). In humans, the red-hair/light-skinned phenotype is associated with an inability to tan after UVR exposure. The gene most commonly implicated in this phenotype is MC1R. The so-called “red-hair” alleles of POMC are notable for their inability to signal to adenylate-cyclase upon exposure to the ligand (αMSH). These MC1R variants thus provide genetic evidence of the rate-limiting role for MC1R signaling in the UVR-induced tanning pathway. Correspondingly, it has been demonstrated in animal models of red hair (carrying MC1R loss-of-function) that topical

15

administration of cAMP-inducing compounds can rescue eumelanin (dark) skin pigmentation (D’Orazio et al. 2006). In sum, skin pigmentation is determined at multiple levels, by (1) the migration of melanoblasts to various tissues during development, (2) the melanoblast survival and differentiation to melanocytes, (3) the density of melanocytes, (4) the expression/function of enzymatic and structural constituents of melanosomes, (5) the synthesis of different types of melanin (eumelanin and pheomelanin), (6) the transport of melanosomes to dendrites, (7) the transfer of melanosomes to keratinocytes, and (8) the distribution of melanin in suprabasal layers of the skin. Skin pigmentation plays a critical role in protecting the underlying skin from UVR damage that can ultimately result in various types of skin cancers, including melanoma.

Regulation of Survival Survival is not only a critical issue for developing melanoblasts but also for melanocytes in the skin and in hair bulbs. As noted above, this is especially important to melanocytes because of their extremely low rates of proliferation, combined with the high stresses imposed by their environment and also by the potentially toxic by-products of melanin synthesis within them. In this regard, the functions of antiapoptotic proteins, such as BCL2 and survivin, are essential for melanocyte survival, as are transcription factors involved in regulating their expression (e.g., MITF and PAX3) and upstream signaling events (e.g., mediated by the actions of KITLG:KIT, FGF2:FGFR, HGF:MET, and αMSH:MC1R). Dysfunctions of many of those genes/proteins have been associated with premature hypopigmentation of the skin and hair graying, as discussed and cited previously. Melanocyte stem cells A closely related issue is the presence and maintenance of melanocyte stem cells, i.e., cells that retain the potential to replenish melanocytes in the skin and hair as needed yet remain undifferentiated. This population has potential significance regarding the

16

process of melanomagenesis, as discussed later in this chapter. Melanocyte stem cells typically reside in what is termed the bulge region, a hair follicle stem cell niche located at the insertion site of the arrector pili muscle (which produces a small “bulge”) along each hair shaft. During the hair growth cycle, active melanocytes are normally present exclusively in the hair bulb and function by synthesizing melanin granules and transferring them to keratinocytes within the bulb, which incorporates the melanin into the growing keratin-containing hair matrix. Hair follicles undergo cyclic patterns of growth, involution, and rest. The growth phase (anagen) has a finite lifetime (measured in years in adult humans), which is followed by the catagen stage, at which time all active keratinocytes and melanocytes within the hair bulb die. New hair bulbs require a fresh supply of functional melanocytes; else the hair will be partially or totally unpigmented. That supply of active melanocytes comes from a pool of melanocyte stem cells that remain localized in the bulge region throughout the follicle stages. Telogen is the resting phase that follows the catagen (involution) phase. During telogen, melanocyte stem cells are in a largely undifferentiated state, but they do not express significant levels of TYR and other proteins required to produce melanin; hence they are unpigmented in the hair bulge region. Although there is some argument about the exact patterns of marker expression, the melanocyte stem cells do express a number of markers that identify them. Typically they are positive for PAX3 and SOX10 and weakly positive for MITF and DCT (which is likely expressed due to PAX3/ SOX10 activation); they are negative for other melanosomal proteins (enzymatic and structural) such as TYR, TYRP1, and PMEL (Buac and Pavan 2007; Lang et al. 2005; Nishikawa and Osawa 2007; Nishimura et al. 2005; Osawa et al. 2005). They retain expression of surface receptors such as KIT, MET, and EDNRB, which allows them to respond to environmental cues and start the differentiation/proliferation phase (anagen) to become functional melanocytes that incorporate into the newly formed hair bulb. Although these melanocyte stem cells are typically thought of as reservoirs required to repopulate melanocytes in

M. R. Zaidi et al.

hair bulbs, there is evidence to suggest that they also can be recruited to replace melanocytes at the dermal-epidermal border. The self-renewing capacity of bulge melanocyte stem cells is thought to be imperfect as depletion of these hair follicle melanocyte stem cells has been seen to correlate with age-dependent hair graying in both mice and humans (Nishimura et al. 2005).

Melanomagenesis From Melanocyte to Melanoma: A Multistep Process Over 20 years ago, Clark et al. (1984) devised a model for the stepwise development of melanoma from normal melanocytes based on both clinical and histopathologic features. According to this model, melanoma progression includes the following stages: (1) common acquired (benign) and congenital nevi consisting of nests of cytologically normal melanocytes, thought to exist in a senescent state; (2) dysplastic nevi characterized by structurally and architecturally atypical features and posing a risk factor for melanoma (Goldstein and Tucker 2013); (3) radial growth phase (RGP) or microinvasive malignant melanoma characterized by intraepidermal proliferation and “pagetoid spreading”; (4) vertical growth phase (VGP) or invasive malignant melanoma associated with the ability to penetrate through the basement membrane into the underlying dermis and metastasize and with a significantly poorer prognosis; and (5) metastatic melanoma, which has spread to other areas of the skin and other more distant sites. However, this sequence of progression can be histologically documented in only about a third of all melanoma cases, and malignant melanoma may also be able to arise in the absence of visually obvious pathologic progenitor lesions. How does this aggressive, often fatal disease arise from melanocytes? What goes wrong? Based originally on extensive cytogenetic analyses of a variety of melanocytic lesions, and later confirmed by high-resolution genomic analyses (e.g., array-comparative genomic

Biology of Melanocytes and Primary Melanoma

hybridization [CGH]), the degree of detectable genetic alterations increases rather dramatically as it progresses from nevus to primary melanoma to metastatic melanoma (Chin et al. 2006). There are few consistent changes associated with each of these stages, but years of such analyses have identified a number of genes and pathways that have proved critical to melanoma development and provided important hints about the transformation process (Shain et al. 2015). While numerous reports demonstrate gene expression changes in melanoma (Riker et al. 2008), we begin by examining some of the major early insights that came from extensive classic cytogenetic studies and array-CGH-based genome-wide analyses of nonrandom chromosomal alterations in melanomas (Bauer and Bastian 2006; Jonsson et al. 2007; Stark and Hayward 2007).

Molecular Genetics: Early Lessons from Familial and Sporadic Melanoma The CDKN2A locus It has long been recognized that melanoma susceptibility can be inherited, as in familial atypical multiple mole melanoma (FAMMM) syndrome (Soura et al. 2016). Subsequent genetic analysis of large melanoma-prone families led to the identification of the melanoma susceptibility locus CDKN2A at chromosome band 9p21 (Aoude et al. 2015; Chin et al. 2006; Kamb et al. 1994). The CDKN2A locus shows loss of heterozygosity (LOH) and loss of function mutations in 25–40% of melanoma kindreds and less frequently in sporadic melanoma. Subsequent experimental studies using mice genetically engineered to carry inactivated CDKN2A have corroborated the importance of this locus in cutaneous malignant melanoma (Perez-Guijarro et al. 2017). The CDKN2A locus is quite unique, encoding two distinct proteins within overlapping DNA sequences: p16INK4A and p14ARF (p19Arf in mice). This is accomplished through the initiation of transcription from two distinct promoters upstream of two distinct first exons (lα for p16INK4A and 1β for p14ARF). Thus, the genes encoding p16INK4A and p14ARF share common second and third exons, translated in

17

different reading frames. Although genetic deletions in this region typically inactivate both p16INK4A and p14ARF in melanoma, many cases have been documented in which expression of only one is affected, suggesting that each has a key role in melanomagenesis (Chin et al. 2006). Remarkably, both function as tumor suppressor genes, with p16INK4A functioning as a cyclindependent kinase inhibitor (CKI) to help regulate transit through the cell cycle and p14ARF stabilizing TP53 by binding and inhibiting the TP53 E3 ubiquitin ligase, HDM2 (Zhang et al. 1998). p16INK4A binds to cyclin-dependent kinases (CDK) 4 and 6, preventing its association with D-type cyclins and inhibiting its ability to phosphorylate and inactivate RB, which in turn initiates S phase associated DNA synthesis (Fig. 3). This is a critical pathway in melanomagenesis, as evidenced by the fact that in melanoma p16INK4A expression can be suppressed by mutation, methylation, and/or polymorphic variation at both the 50 and 30 untranslated regions (Chin et al. 2006). Moreover, germline mutations in CDK4 have been documented in melanoma patients that prevent binding of CDK4 to p16INK4A, as have amplifications in CCND1 (encoding Cyclin D1) and occasional mutations in RB1 (Tsao et al. 2012). Notably, the various p16INK4A/CDK4/RB mutations were largely found to be mutually exclusive and can be considered redundant. p14ARF acts as a positive regulator of TP53. It targets the degradation of HDM2 (Mdm2 in mice), which destabilizes TP53. TP53 is a sequence-specific transcription factor that triggers either cell cycle arrest or programmed cell death (apoptosis) on DNA damage or other stress (Liebermann et al. 2007), depending on the cell type and the circumstances (see Fig. 3). TP53 is the most frequently mutated gene in human cancer but is a relatively infrequent target in melanoma, estimated to be mutated in fewer than 10% of melanomas (Bennett 2008; Hussein et al. 2003b; Petitjean et al. 2007). As mentioned above, p14ARF is frequently inactivated in melanoma, and HDM2 amplification has been documented as well. Genetically engineered mouse models (GEMMs) of human melanoma

18

M. R. Zaidi et al.

Fig. 3 Critical pathways in melanomagenesis. The pathways are simplified, and arrows do not necessarily represent direct interactions (see text for more details.) Factors highlighted in red have been associated with activating

mutations and/or increased copy numbers in melanoma, whereas those highlighted in green have been inactivated and/or deleted. The sun symbols represent pathways thought to be directly activated by UV radiation

have corroborated the separate roles of p19ARF and p16INK4A in melanomagenesis (Ha et al. 2007; Perez-Guijarro et al. 2017). Moreover, recent evidence indicates that p19ARF has melanoma tumor suppressor functions that are independent of TP53, including a role in melanocyte senescence (Ha et al. 2007). Cyclin-dependent kinase 4 Germline mutations at chromosome 12q14 in the CDK4 locus are oncogenic due to their effects on cell cycle control via the same pathway as p16INK4A. Two different mutations have been identified in codon 24 of exon 2 (R24C and R24H) at the CDK4 locus, both of which result in CDK4 protein acting as a dominant oncoprotein due to loss of binding to p16INK4A, which is its negative regulator (Puntervoll et al. 2013). CDK4 locus is mutated or amplified in only 5% human melanomas, albeit at a higher frequency in patients with BRAF, NRAS, and NF1 triple wild-type tumors

(Cancer Genome Atlas 2015). GEMM studies have verified the oncogenic role of the CDKR24C mutant protein in melanomagenesis (Gaffal et al. 2011; Tormo et al. 2006). BRCA1-associated protein BAP1 is a tumor suppressor gene located on chromosome 3p21. Germline inactivating mutations at this locus were initially identified in two distinct syndromes; one was characterized by familial mesothelioma and uveal melanoma (Testa et al. 2011) and the other by cutaneous melanocytic neoplasia and uveal melanoma (Wiesner et al. 2011). Subsequently, cutaneous melanoma was included as part of the familial aggregation of cancers associated with the BAP1 syndrome (Carbone et al. 2013; Wadt et al. 2012; Wiesner et al. 2012). Approximately 5% of sporadic cutaneous melanoma have been proposed to have functional inactivation of BAP1 characterized by absence of BAP1 expression on immunohistochemistry

Biology of Melanocytes and Primary Melanoma

analysis, suggesting its contribution to melanomagenesis in at least a subset of cases (Murali et al. 2013). The PTEN-AKT pathway PTEN is localized on chromosome band 10q24, a region long associated with deletions and LOH in melanoma (Chin et al. 2006). PTEN is a dual phosphatase, containing both lipid phosphatase and protein phosphatase activities (Lee et al. 2018). Extracellular growth signals are often mediated through the intracellular second messenger lipid phosphatidylinositol-3,4,5-triphosphate (PIP3), levels of which increase upon signaling through PBK, resulting in the phosphorylation/activation of AKT (also known as protein kinase B) (see Fig. 3). The PTEN lipid phosphatase antagonizes PI3K by dephosphorylating PIP3 and therefore negatively regulates AKT activity, helping to control both cell proliferation and survival. PTEN deletions or inactivating mutations can be found in approximately 28% of melanoma cell lines, 7% of primary melanomas, and 15% of metastatic melanoma (Aguissa-Toure and Li 2012), causing sustained phosphorylation/activation of AKT; moreover, AKT3 overexpression has been reported to be associated with an increased DNA copy number in some melanomas (Robertson 2005). Despite intense study, the regulation and complex roles of PTEN are still incompletely understood. For example, PTEN has been found to be secreted into the extracellular space to be taken up by recipient cells, intriguingly suggesting a function as a cell nonautonomous tumor suppressor (Hopkins et al. 2013; Putz et al. 2012). The WNT/β-catenin pathway β-catenin, encoded by the gene CTNNB1, is a key regulator of the WNT signaling pathway, well-documented for its involvement in the progression of many cancer types (Kaur et al. 2016; Zhan et al. 2017) and also in the physiologic differentiation of melanocytes from neural crest progenitors. In the absence of WNT signaling, β-catenin is targeted for degradation in association with a complex that includes adenomatous polyposis coli (APC) and glycogen synthase kinase 3β (GSK3β) (see Fig. 3). Upon WNT signaling, β-catenin is stabilized, accumulates, and is transported to the nucleus

19

where it binds to LEF1 and acts as a transcriptional coactivator of LEF1 target genes. Downstream targets include the proto-oncogenes MYC and CCND1 and the gene encoding MITF (Zhan et al. 2017). Mutations have been detected in CTNNB1 that confer a resistance to β-catenin degradation in 2–23% of melanoma tissues and cell lines (Dahl and Guldberg 2007; Larue and Delmas 2006), with a higher frequency in vitro. Although infrequently mutated in melanoma, APC is more frequently characterized by decreased expression. β-catenin has also been shown to repress p16INK4A transcription (Delmas et al. 2007). NRAS and BRAF In 2002 Davies et al. (2002) published a remarkable paper in which a genomewide cancer sequencing program was used to discover that BRAF was mutated in 66% of melanoma samples, a finding that had significant implications for the treatment of melanoma. This finding was later confirmed and extended to melanocytic nevi, where approximately 80% harbored mutant BRAF, including benign nevi (Pollock et al. 2003). Oncogenic BRAF mutations in melanoma appear to be somatic, as germline mutations are not associated with enhanced cancer risk (Niihori et al. 2006; Rodriguez-Viciana et al. 2006). The most common BRAF mutation (V600E, found in more than 90% of melanoma mutations) introduces a conformational alteration in the BRAF activation domain, causing constitutive kinase activation (Wan et al. 2004). BRAF, a member of the RAF family of genes (ARAF, BRAF, CRAF), is a serine-/threonine-specific protein kinase that activates sequentially MEK1/2 and then ERK1/2 (Garnett and Marais 2004). RAF is a downstream target of the RAS family of small guanine-nucleotide-binding proteins (NRAS, HRAS, KRAS). It was therefore not surprising to find that a RAS, almost always NRAS, is mutated in 15–25% of melanoma cell lines and tumors but rarely in melanomas harboring mutations in BRAF (Bennett 2008; Davies et al. 2002; Jonsson et al. 2007). These results strongly endorse the importance of the NRAS-BRAFMEK-ERK (MAP kinase) pathway in melanomagenesis (see Fig. 3). It has been suggested that cancers may become overly dependent on, or “addicted” to, one or a few specific

20

genes for maintenance of the malignant phenotype and survival (Weinstein and Joe 2008); in melanoma this may apply to oncogenic members of this pathway. In vivo mouse experiments have confirmed the key role of this pathway in melanoma, as well as the ability of members of this pathway to collaborate with loss of the CDKN2A locus in provoking melanoma development (Ackermann et al. 2005; Burd et al. 2014; Chin et al. 1997; Damsky et al. 2015; Dhomen et al. 2009; Goel et al. 2009; Kwong et al. 2012; PerezGuijarro et al. 2017; Sharpless et al. 2003). Receptor tyrosine kinases (RTKs) RTKs play crucial roles in regulating virtually all basic cell processes under normal physiological conditions and have been implicated in the development of most cancers. The exact phenotypic response to RTK signaling is extremely complex, depending on the RTK and its signal, the cell type, and the circumstances under which it is stimulated. In melanoma, examples of DNA copy number gains and point mutations have been documented that involve several RTKs. For example, increased copies of regions of chromosome 7, which encodes the epidermal growth factor receptor (EGFR) at 7p12, have been detected in late-stage melanomas, along with enhanced EGFR expression (Chin et al. 2006). The EGFR is an intriguing candidate because melanomas develop in Xiphophorus fish in association with aberrant expression of activated Xmrk, an EGFR-related gene. Another region of chromosome 7 (7q33-qter) that exhibits increased copy number in late-stage melanoma harbors the MET gene, encoding the HGF receptor; moreover, increased MET expression has been detected in metastatic melanoma (Natali et al. 1993). Activating point mutations in EGFR or MET have not been detected to date in melanomas. However, a recurrent L576P mutation in KIT has been reported for a small number of melanomas, with amplification and/or selective loss of the normal allele (Curtin et al. 2006; Willmore-Payne et al. 2005). Interestingly, melanomas with KIT mutations do not carry BRAF mutations and arise only from chronically sun-damaged skin (see the “Melanoma and the Environment” section). L576P is a GIST-associated, activating mutation, and as such

M. R. Zaidi et al.

represents a promising target for the RTK inhibitor imatinib in this subset of melanoma patients. However, clinical trials of the RTK inhibitors imatinib and sunitinib for KIT-mutated melanomas showed only modest responses (Buchbinder et al. 2015; Carvajal et al. 2011; Guo et al. 2011; Hodi et al. 2013). Some enzymes with appropriately targeted phosphatase activity counter the RTKs and may act as tumor suppressors. For example, the gene encoding protein tyrosine phosphatase receptor type D (PTPRD) was found to be homozygously deleted in 9% of melanoma cell lines (Stark and Hayward 2007). Moreover, upregulation of RTKs, e.g., PDGFR and IGF-1R, has been shown to play a key role in determining resistance of melanoma to BRAF inhibitor drugs (Nazarian et al. 2010; Villanueva et al. 2011). Transcription factors High-density single nucleotide polymorphism (SNP) arrays were used to detect recurrent focal amplifications of the transcription factor MITF in 10% of primary melanomas and up to 20% of metastatic melanomas (Garraway et al. 2005). A recurrent activating mutation in MITF has also been identified in the germline of kindred from multiple continents with familial melanoma (Yokoyama et al. 2011). The recurrent mutation stimulates MITF’s transcriptional activity by disrupting a SUMO-modification consensus sequence that otherwise suppresses MITF function. As discussed earlier, MITF is a melanocyte lineage-survival factor required for the commitment of immature cells to the melanocyte lineage during development and heavily involved in melanocyte survival, growth, and differentiation (Kawakami and Fisher 2017) (see Fig. 3). Melanomas resistant to BRAF inhibitors have been shown to overexpress MITF, and knocking down MITF expression reversed this resistance (Smith et al. 2013). The reliance of melanocytes on MITF makes it an intriguing candidate target for melanoma patients (Merlino 2005). However, akin to most transcription factors, the lack of a catalytic domain in MITF protein makes it a problematic therapeutic target. Consequently, a number of strategies are being employed to inhibit MITF indirectly; for example, inhibitors of HDAC, AMP-activated kinase

Biology of Melanocytes and Primary Melanoma

(AMPK), and HIV1-protease have been shown to suppress MITF expression (Borgdorff et al. 2014; Smith et al. 2016; Yokoyama et al. 2008). A number of additional transcription factors have been implicated in melanoma progression (Poser and Bosserhoff 2004). The transcription factor TBX2, which can repress expression of p14ARF and p21CIP1 (product of CDKN1A gene), exhibits an increased gene copy number and overexpression in melanoma (Jacobs et al. 2000; Jonsson et al. 2007; Vance et al. 2005). Gene copy number increases up to 40% have also been reported in melanomas for the transcription factor MYC (Bennett 2008), but conflicting reports (Stark and Hayward 2007) have indicated a complex and yet unresolved role of MYC in melanoma. Recently, MYC overexpression in melanoma has been suggested to be correlated with accelerated tumor metastasis in vivo (Lin et al. 2017). The factors TP53 and β-catenin have already been discussed above. The pro-metastasis gene NEDD9 Over the years many genes have been implicated in melanoma metastasis (Chin et al. 2006; Crowson et al. 2007; Qiu et al. 2015; Xu et al. 2008). Proceeding on the assumption that overlapping genomic changes in human and mouse melanomas arise through shared selective pressure and represent critical events in melanomagenesis, Kim et al. used genome-wide high-resolution array-CGH to identify the scaffold protein NEDD9 as a prometastasis factor (Kim et al. 2006). Genetically engineered mouse models of human melanoma were employed to narrow down candidates in an amplified syntenic region in the human genome to NEDD9. Loss-of-function and gain-of-function experiments were employed to confirm that NEDD9 confers enhanced invasiveness and metastatic potential to melanocytic cells. NEDD9 was found to facilitate metastasis through interaction with focal adhesion kinase (Kim et al. 2006).

Melanoma: A Consequence of Homeostatic Disruption No cell is an island, and melanocytic cells live in dynamic and harmonic equilibrium with

21

their microenvironment. An intricate network of interactions ensures a state of homeostasis both during development and under normal adult physiological conditions. As discussed earlier, the melanocyte microenvironment includes cellular neighbors such as keratinocytes, fibroblasts, endothelial and immune cells, as well as components of a complex ECM. In particular, melanocytes and keratinocytes (the so-called melanin unit) are intimately associated with each other in the basal layer of the epidermis of the human skin. Each melanocyte makes contact with, and sends pigment-containing melanosomes to, dozens of keratinocytes, a process that can be dramatically altered by exposure to UVR. Keratinocytes are thought to employ the pigment as protection from subsequent UVR damage. Keratinocytes regulate melanocyte growth and behavior through the activity of cell-cell adhesion molecules and paracrine growth factors (Haass et al. 2005). The initiation and progression of melanoma marks a significant disruption in this tranquil homeostasis, involving key changes within the melanocyte genome and epigenome, and in the way these melanocytes grow, survive, and interact with their microenvironment (Haass and Herlyn 2005). Moreover, it is clear that exposure to UVR, known to initiate melanomagenesis, is a powerful homeostatic disruptor. Dysregulation of melanocyte gene expression Gene expression profiling using multiple technologies has permitted the simultaneous analysis of the entire melanoma genome. Studies have identified multiple significant changes at the level of gene expression as melanocytes are transformed into melanoma cells and further into metastatic melanomas (Haqq et al. 2005; Hoek et al. 2004; Hoek 2007; Kaufmann et al. 2008; Talantov et al. 2005). Moreover, exposure to UVR has been shown to induce significant alterations in gene expression patterns of melanocytes (Yang et al. 2006; Zaidi et al. 2011). Several mechanisms undoubtedly contribute to the observed global changes in expression. First, the expression or activity of many important transcription factors can be altered in melanoma, including MITF, LEF1/β-catenin, TBX2, MYC, PAX3, SOX10, and SLUG (Poser

22

and Bosserhoff 2004) (see Table 1). Another interesting transcriptional regulator, inhibitor of differentiation (ID1), has been shown to repress p16INK4A transcription and delay melanocyte senescence (Cummings et al. 2008; Polsky et al. 2001) and may play an early role in melanomagenesis. These transacting factors not only help regulate the expression of a plethora of critical downstream targets but can influence each other’s expression/activity as well, further demonstrating the complexity of this regulatory network. In addition, the modification of chromatin is another level of gene regulation that plays a major role in the development of tumors, including melanoma (Moran et al. 2018). Epigenetic changes include DNA methylation and numerous covalent modifications of histones, including methylation, acetylation, phosphorylation, sumoylation, and ubiquitination. In melanoma, promoter methylation has been documented in suppressing the expression of genes such as CDKN2A, PTEN, RASSF1A, MGMT, DAPK, RARB2, and APC. Finally, a relatively newly recognized category of powerful gene regulators is the microRNAs (miRNAs), which has been shown to participate in tumorigenesis, including metastasis (Di Leva et al. 2014). miRNAs are short molecules of RNA of about 22 nucleotides in length and are the most expressed class of noncoding RNAs in eukaryotic cells. Their main gene regulatory function is through silencing or degrading mRNAs, and a single miRNA can potentially regulate several target mRNAs simultaneously. Consequently, miRNAs can have oncogenic or tumor suppressor effects depending on the types of genes they regulate. miRNAs have been extensively studied in melanoma cells and clinical samples, and dysregulation of a relatively large number of melanoma-specific miRNAs has been identified, with regulatory effects on gene expression affecting a variety of pathways (Fattore et al. 2017). Nextgeneration sequencing studies have identified another category of regulatory RNAs called long noncoding RNAs (lncRNAs), which are defined as transcripts >200 nucleotides in length. Several lncRNAs have been characterized to play key roles in cellular proliferation, survival, migration, and genome stability, and their aberrant

M. R. Zaidi et al.

expression is associated with many cancer types, including melanoma (Huarte 2015; Leucci et al. 2016). Together, genetic and epigenetic alterations can cooperate to disrupt normal homeostasis as melanocytes move along the path to malignancy. Dysregulation of melanocyte proliferation/ differentiation Adult melanocytes are built to manufacture pigment, to respond to environmental challenges through enhanced differentiation but controlled growth, and to survive. In contrast, melanoma cells proliferate autonomously and resist differentiation and growth inhibition signals. How does this happen? It is important to appreciate that these critical melanocytic functions are regulated more through highly interactive webs or networks and not simple linear pathways (Palmieri et al. 2015) (see Fig. 3). What sits at the center of this web is the melanocyte master regulator MITF (Kawakami and Fisher 2017; Palmieri et al. 2015). MITF can regulate virtually all known melanocyte-specific genes, including TYR, DCT, and TYRP1. MITF is absolutely required for melanocyte survival, partly by regulating expression of survival factors such as BCL2, and a number of MITF-mutant mouse strains have been described that are devoid of melanocytes. Both a positive and a negative influence, MITF is also required for melanocyte proliferation but depending on its level of expression or posttranslational modification can also be growth inhibitory (Kawakami and Fisher 2017). For example, MITF can transactivate stimulators of growth such as CDK2 and TBX2 but can also directly transactivate growth inhibitors such as p16INK4A and p21CIP1 and indirectly activate the CDK inhibitor p27KIP1 (product of CDKN1B gene). The expression or activity of MITF can in turn be regulated by numerous key melanocytic factors, including the transcription factors PAX3, SOX9, SOX10, CREB, and LEF1; the cofactors CBP, RB1, p300, SWI/SNF, and β-catenin; and the MAPK pathway through EDN and EDNRB. The network centering on MITF radiates out to myriad pathways regulating virtually every aspect of melanocyte function. Notably, MITF amplification or mutation can occur within melanomas, suggesting that under the appropriate

Biology of Melanocytes and Primary Melanoma

circumstances, MITF can function as a lineagesurvival oncogene (Garraway and Sellers 2006). The ability of melanocytes to produce pigment is also heavily regulated through MC1R and its ligands, POMC-derived αMSH, and ACTH (Rouzaud et al. 2005). Polymorphic variations and mutations in MC1R in the general population determine the type of pigment that is produced (eumelanin or pheomelanin) and therefore the color of the skin (dark skinned or fair skinned). Notably, fair-skinned individuals have a greater risk for melanoma, linking MC1R to melanomagenesis. A significant association was detected between germline MCIR variants, BRAF mutations, and a subset of melanomas (Landi et al. 2006). As in other types of cells, the key to controlling proliferation is through regulation of transit through the cell cycle. Entry into the cell cycle is normally promoted by growth factors, which in melanocytes include FGF2, HGF, KITLG, EDN3, IGF-1, and the WNTs; these factors can of course regulate other melanocyte behaviors as well, such as survival, differentiation, and migration (see Table 1). By triggering cascading signaling pathways leading to the activation of downstream cell cycle-associated transcription factors (e.g., MYC, FOS, JUN, and ATF2), these growth factors help overcome the G1-S cell cycle checkpoint. These factors stimulate CDK4/6-cyclin D phosphorylation/inhibition of RB and thereby the release of E2F to transactivate critical cell cycle control genes, such as those encoding the E and A cyclins, MYC, and CDK1. CKIs normally serve as key regulators of G1 cell cycle progression and include p16INK4A, p15INK4B, p21CIP1, and p27KIP1. An important part of the melanoma story is written in the deregulation of these pathways. We have already noted above that mutations or altered expression characterizes virtually every key cell cycle regulator. Moreover, melanoma cells often become growth factor-independent, constitutively initiating these important pathways through, for example, activating mutations in RTKs such as KIT. Another mechanism of deregulated growth frequently described in melanoma is through the creation of autocrine signaling loops, wherein melanoma cells express a

23

growth factor or cytokine (e.g., HGF, FGF2, TGFα, VEGF, MGSA/GRO, IL-8/CXCL8), as well as its cognate receptor (e.g., MET, FGFR1, EGFR, VEGFR1/2, CXCR1/2). A critical intermediary between the initial phosphorylation event triggered by the binding of a growth factor to its receptor and the activation of downstream cell cycle-associated transcription factors is the MAPK pathway. RAS gene products are 21 kD GTPase proteins that serve as molecular switches converting the kinase activation events at the cell membrane to nuclear events and ultimately to altered melanocyte behavior (Simanshu et al. 2017). There are three different RAS genes, NRAS, HRAS, and KRAS. Despite their structural similarity, these three RAS genes have distinct functions. In fact, genetically engineered mouse models have confirmed that NRAS is significantly more effective than HRAS at transforming melanocytes in vivo (Perez-Guijarro et al. 2017). Based on the strong preference for NRAS mutations in melanomas, it has been shown that NRAS is the key regulatory G-protein in melanocytes (Burd et al. 2014). RAS activating point mutations cause impaired GTPase activity resulting in a constitutively activated G-protein and deregulated growth control. Similarly, the most common BRAF mutation in melanomas, V600E, is also forced into a state of perpetual activation. The obligatory nature of the deregulated MAPK pathway in melanomagenesis is evident from the high percentage of melanomas with either NRAS or BRAF mutations, which together range from 80% to 90%. Because mutant BRAF and NRAS can be detected in nevi (Damsky and Bosenberg 2017), it is assumed that these mutations are early events in melanomagenesis and are important in clonal expansion. Although the MAPK pathway is most central to melanocyte growth control, a second key RAS effector, PBK, also helps regulate proliferation. Although not as frequently mutated in melanomas, the PTEN-PBK-AKT pathway also plays a critical role in other melanocyte behaviors, including regulation of migration and apoptosis. A mouse model employing grafted human skin tissue containing melanocytes genetically engineered to express specific mutations has confirmed the importance of activated PBK

24

and NRAS, but notably not that of BRAF in melanomagenesis (Chudnovsky et al. 2005). In an extreme example of growth control, cells can enter an irreversible state of growth arrest called senescence. This phenomenon was described in cultured cells many years ago and was initially argued to be an in vitro artifact; however, a number of impressive examples have been reported demonstrating the relevance of in vivo senescence in various types of cancers, including melanoma (Liu et al. 2018; Mooi and Peeper 2006). Early benign pigmented nevi are now thought to represent collections of mostly senescent melanocytes, and senescence is considered a significant barrier that melanoma cells must overcome to become proliferative melanomas (Bennett 2008). Cellular senescence can be induced through two distinct mechanisms. Replicative senescence occurs after extensive proliferation and is associated with telomere exhaustion. The shortening of telomeres is a natural by-product of replication in normal cells with limited telomerase, a heterodimer consisting of the protein TERT and the RNA TERC. After enough replication cycles, chromosome ends become so short that their protective cap structure is compromised, resulting in the activation of TP53-mediated DNA damage signaling, chromosome fusion, and mitotic disruption, termed crisis (Shay 2016). Cancer cells cannot therefore grow indefinitely without employing a mechanism to overcome telomere shortening, as they would become senescent or die. A mechanism that is often exploited by cancer cells, including melanoma cells, to gain the ability to grow indefinitely, or become immortalized, is reactivation of telomerase expression. Senescence can also be induced by oncogene activation or an equally dramatic cellular stress (Ben-Porath and Weinberg 2005; Lowe et al. 2004; Mooi and Peeper 2006). Ironically, despite the fact that mutant BRAF and NRAS appear to be nearly obligatory for the development of human melanomas, these same activated oncogenes can induce a senescence response in melanocytes (Michaloglou et al. 2005). A number of factors have been implicated in overcoming early oncogene-induced senescence, most notably the

M. R. Zaidi et al.

p16INK4A-CDK4/6-RB and ARF-HDM2-TP53 pathways, and the CDKN2A locus in particular. p16INK4A has long been considered a key melanocyte senescence gene; BRAFV600E induces p16INK4A expression, and p16Ink4a deficiency in genetically engineered mice confers immortality to melanocytes and induces melanomagenesis (Liu and Sharpless 2012; Perez-Guijarro et al. 2017). Evidence in mice indicates that the other member encoded by the CDKN2A locus, p19ARF, is also very important in melanocyte senescence, highlighting the importance of the simultaneous loss of p16INK4A and p19ARF function when CDKN2A experiences deletions. Notably, p19ARF was found to induce senescence independently of TP53 in melanocytes (Ha et al. 2007). However, the view that human nevi undergo oncogene-induced senescence has been challenged in the light of the observations that the biomarkers widely used to define senescence are not exclusive to the senescence program (e.g., p16INK4A and SA-β-gal). p16INK4A was found to be abundantly expressed in both nevi and primary melanomas, and SA-β-gal, the most widely accepted senescence marker, was detected in both nevi and metastatic melanoma (Tran and Rizos 2013). These observations cast doubt on human nevi being defined as truly senescent (Tran and Rizos 2013). Dysregulation of melanocyte survival The ability of cancer cells to overcome built-in programming designed to trigger cell death (apoptosis) in response to DNA damage, depletion of survival factors, or disruption of interactions with the microenvironment is a hallmark of successful progression (Hanahan and Weinberg 2011). Apoptosis is activated when a sensor, usually TP53, detects a deathinducing signal, which in turn transmits an execution order to the cell death machinery, often associated with mitochondria. Pro-apoptotic signals, including TP53, NOXA, PUMA, BAX, BAD, tumor necrosis factor (TNF), TNF-related apoptosis-inducing ligand (TRAIL/TNFSF10), and the Fas ligand (FASLG), induce the release of cytochrome c, which in concert with apoptotic protease-activating factor 1 (APAF1) activates caspase-9 and the resulting effector protease cascade (Lowe and Lin 2000).

Biology of Melanocytes and Primary Melanoma

To achieve balance, these pathways are countered by pro-survival factors, such as BCL2, BCL2L1, nuclear factor-κB (NFKB), survivin (BIRC5), and livin (BIRC7) (Hussein et al. 2003a). Melanocytes have evolved to survive the adverse environment created from the production of melanin and exposure to UVR. Therefore, it is not surprising that melanoma cells utilize several effective means to avoid apoptotic destruction (see Fig. 3) (Hussein et al. 2003a). Although TP53 is infrequently mutated in melanomas, loss of its positive regulator, p14ARF, indirectly hampers the TP53mediated DNA damage response. Melanoma cells characteristically produce elevated levels of BCL2, limiting cytochrome c-mediated caspase9 activation. Caspase-9 activity is also inhibited through the influence of the kinase AKT, often highly active in melanomas. Advanced melanomas have been found to overexpress survivin, which binds to and inhibits effector caspase-3 and caspase-7, acting downstream of caspase-9 (Grossman et al. 1999). Similarly, livin is overexpressed in melanomas and can inactivate caspase-3, caspase-7, and caspase-9 (Kasof and Gomes 2001). Loss of APAF1 expression has been associated with melanoma cell survival and chemoresistance (Soengas et al. 2001). Caspases are also triggered by activation of the death receptors, through binding of TNF, FASLG, and TRAIL (TNFSF10) (Ivanov et al. 2003). For example, TRAIL can induce apoptosis through interactions with the death receptors TRAIL-R1 (TNFRSF10A) and TRAIL-R2 (TNFRSF10B), which are expressed in most normal tissues as well as in melanomas; normal tissues survive by expressing the inhibitory decoy receptors TRAILR3/DcR1 and TRAIL-R4/DcR2 (Lowe and Lin 2000). However, these survival pathways are complex, and death receptor activation also triggers NF-κB, which is frequently upregulated in melanoma and is an essential survival-promoting factor in melanoma. Countering the ability of melanoma cells to evade apoptotic destruction is an attractive therapeutic approach and is currently being actively pursued (Fulda 2015). Dysregulation of cell-cell and cell-ECM interactions Melanocytic cells, particularly melanoblasts, are highly migratory by nature. It has

25

therefore been suggested that melanoma cells become metastatic by coopting and enhancing this innate migratory tendency (Gupta et al. 2005). In the initial phase of this cellular relocation, melanocytes must first sever their molecular ties with keratinocytes, ties consisting of cell surface molecules that promote interactions with keratinocytes (Haass et al. 2005). Alterations in expression of cell adhesion molecules such as cadherins, integrins, and immunoglobulin superfamily members can significantly disrupt the homeostatic balance between melanocytes and keratinocytes, facilitating melanoma progression. Cadherins are multifunctional transmembrane proteins that act as both receptor and ligand to maintain appropriate cell-cell contacts. The switch in expression from E-cadherin (encoded by CDH1) to N-cadherin (CDH2), first observed in RGP melanoma, redirects gap junction formation from E-cadherin-expressing keratinocytes, which inhibit melanocyte proliferation and maintain a differentiated morphology, to N-cadherinexpressing fibroblasts and vascular endothelial cells, encouraging migration from the epidermis to the dermis (McGary et al. 2002). N-cadherin also enhances melanoma survival by stimulating β-catenin signaling, LEF1 activity, and subsequent activation of MITF and CCND1. E-cadherin is downregulated in part by the transcription factor Snail (SNAI1) (Tsutsumida et al. 2004), whereas N-cadherin expression is upregulated by the Notch signaling pathway (Liu et al. 2006), already implicated in melanomagenesis (Nickoloff et al. 2005). Loss of E-cadherin is often accompanied by changes in other cell adhesion molecules, including integrins. Integrins are heterodimeric cell surface receptors, consisting of 1 of 20 different α chains and 1 of 9 different β chains. The anchorage of melanocytes to the dermal-epidermal basement membrane can be weakened through changes in melanocytic integrins. Melanomas express specific integrins (e.g., αvβ3 and α4β1) that have been implicated in adhesion to ECM, promoting motility, invasion, and metastasis, which correlates closely with poor clinical outcome in melanoma patients (Kuphal et al. 2005). For example, integrin αvβ3, the vitronectin receptor,

26

can facilitate melanoma cell growth, survival, and matrix invasion and marks the progression of RGP to VGP melanoma. Expression of integrin αvβ3 has also been shown to correlate with the activity of the matrix metalloproteinases (MMPs) MMP-1 and MMP-2 (Hofmann et al. 2000). Melanomas also overexpress immunoglobulin superfamily members, including MCAM (MUC-18/CD146), ICAM-1 (CD54), and ALCAM (CD166) (Crowson et al. 2007; McGary et al. 2002). MCAM is overexpressed in more than 80% of melanomas and mediates interactions between melanoma cells and both the matrix and other cell types such as endothelial cells. In a three-dimensional epidermal skin model, melanoma cells expressing MCAM exhibited enhanced separation from the epidermis and invasion through the basement membrane and into the dermis (Satyamoorthy et al. 2001). Melanoma cell invasiveness can be accounted for in part by their ability to induce basement membrane breakdown and other forms of remodeling through secretion of proteolytic enzymes such as hyaluronidase, heparanase, and MMPs (Edward and MacKie 1993). Melanoma cells can also recruit surrounding fibroblasts to aid and abet invasion by encouraging secretion of their proteolytic enzymes. The degraded ECM in turn releases additional factors that can induce melanoma cell and endothelial cell growth, stimulating angiogenesis. The MMPs are broadly acting, zinc-dependent enzymes capable of functioning as collagenases, gelatinases, or stromelysins and are classified accordingly. MMP activity is carefully regulated, partly through their tissue inhibitors (TIMPs). Increased expression of a number of MMPs has been documented in invasive melanomas, including MMP-1, MMP-2, and MMP-9 and MT1-MMP (Crowson et al. 2007). One focal point in melanoma research has been MMP-2, which localizes to the melanomastromal interface in primary and metastatic lesions and correlates with progression. Of interest, MMP-2 is activated in association with a complex that includes MT1-MMP, TIMP-2, and integrin αvβ3 (Hofmann et al. 2000).

M. R. Zaidi et al.

Melanoma: Cell of Origin Pluripotent stem cells are defined by their abilities to indefinitely self-renew and to give rise to an organ-specific differentiated repertoire of cell types. These stem cells tend to be relatively quiescent and to express drug resistance and antiapoptotic genes. Based mostly on data from hematologic malignancies and a few solid tumors, it has long been suggested that tumors harbor relatively rare cells with similar properties, called cancer stem cells or tumor-initiating cells (Batlle and Clevers 2017). Several critical questions followed upon acceptance of this hypothesis. One involved the cancer cell of origin: Do cancers arise from stem cells that have become malignant, or do cancers arise from more differentiated cell types, which then acquire the ability to selfrenew? A second question concerned the ability of tumors to recur after drug treatment: Would cancer stem cells be intrinsically resistant to conventional therapeutic treatment? Adult melanocyte stem cells have been identified and reside within a specialized niche within the hair follicle called the bulge region (Nishimura 2011; Nishimura et al. 2005; Nishimura et al. 2002). During cyclic hair renewal, dormant bulge melanocyte stem cells are activated by specific signaling molecules and then give rise to transiently amplifying progeny, which proliferate, differentiate into pigment-producing melanocytes, and migrate to the base of the hair follicle. Transiently amplifying melanocytes can also migrate from the bulge region to the epidermis where they function as differentiated, pigmented melanocytes. Interestingly, a number of factors implicated in regulating melanocyte stem cell function – MITF, PAX3, NOTCH, and SLUG (SNAI2) – were also linked to melanomagenesis (Garraway and Sellers 2006; Gupta et al. 2005; Hoek et al. 2004; Nickoloff et al. 2005; Plummer et al. 2008), supporting the idea that a cancer stem cell was the melanoma cell of origin. Fang et al. (2005) cultured nonadherent spheroids from human melanoma cells or from metastases in an ES-like medium and identified a CD20-rich subpopulation that could be induced to differentiate into multiple phenotypes, consistent with stem

Biology of Melanocytes and Primary Melanoma

cell properties. Frank and colleagues (Frank et al. 2005; Schatton et al. 2008) identified a subpopulation of human melanoma cells expressing the chemoresistance mediator ABCB5, which was capable of self-renewal and differentiation as well as serial and limiting-dilution transplantation. It was postulated that the ABCB5-expressing cancer stem cells would represent a population highly resistant to conventional melanoma therapy. Recently, however, it has been reported that differentiated pigment-producing melanocytes located in the interfollicular epidermis can give rise to melanoma (Kohler et al. 2017). Using a BrafV600E-driven mouse melanoma model, Kohler et al. showed that the melanocyte stem cells residing in the hair follicle bulge region are refractory to melanomagenesis, and in fact, melanoma originates from the pigment-producing, but not amelanotic, melanocytes located in the interfollicular regions. These pigment-producing mature melanocytes exhibited loss of differentiation markers before dermal invasion, suggesting that a transcriptomic reprogramming had caused a reversion to a dedifferentiated tumor-initiating state (Kohler et al. 2017). It is unclear what microenvironmental cues play a role in dictating this reprogramming.

Melanoma and the Environment Sun Exposure and Epidemiology Melanoma rates continue to rise in most populations at a time when the incidence of many other cancers is on the decline (Cronin et al. 2018; Nikolaou and Stratigos 2014). Extensive epidemiologic studies have provided strong support for the notion that UVR is a significant environmental carcinogen for melanoma (Garibyan and Fisher 2010; Moan et al. 2008a; Tran et al. 2008). The relentless rise in skin cancer incidence despite broad knowledge of the carcinogenic activity of UVR is likely a manifestation of recent evidence demonstrating that release of betaendorphin and consequent opiate-like responses occur as by-products of the UVR-tanning response (Fell et al. 2014). Supporting

27

experimental evidence for UVR’s carcinogenic activity has come from studies of the effects of UVR on a variety of animal models, including the opossum, the fish Xiphophorus, the Sinclair swine, and several lines of genetically engineered mice (Jhappan et al. 2003; Zaidi et al. 2008). However, significant questions remain as to how UVR causes melanoma and under what circumstances (Maddodi and Setaluri 2008). For example, unlike other skin cancers, melanoma risk does not appear to be associated with cumulative lifetime exposure to UVR; instead, intermittent burning doses of UVR represent a more significant risk factor. Other epidemiologic data draw a critical association between melanoma and childhood UVR exposure (Green et al. 2011; Whiteman et al. 2001). Immunologic differences between children and adults, or differences in their proportion of melanocytic progenitors, have been suggested as possible explanations for this observation (Zaidi et al. 2012). So, while the association between melanoma and sunlight exposure has long been appreciated, the relationship between exposure and risk is complex, and the underlying mechanisms are largely unknown. In contrast, great strides have been made in classifying melanomas based on epidemiologic and molecular associations (Bastian 2014; Cancer Genome Atlas 2015). In perhaps the best example, Curtin et al. (2005) demonstrated that melanomas arising from skin sites not subjected to chronic sun damage (e.g., those receiving intermittent sun exposure) were highly likely to carry characteristic BRAF mutations, whereas in melanomas arising from skin with chronic sun damage, or regions that virtually never see sunlight (e.g., acral and mucosal melanomas), such mutations were very rare. In support of this relationship, congenital nevi arising without UVR exposure also lacked BRAF mutations (Bauer et al. 2007). Interestingly, as if to compensate for the lack of BRAF mutations, melanomas from chronically sundamaged skin frequently harbor activating oncogenic mutations in KIT, a receptor tyrosine kinase upstream of BRAF (Curtin et al. 2006). Despite the general agreement on the causal role of UVR in melanoma, the efficacy of sunscreens in the prevention of melanoma remained

28

controversial; in fact, some earlier epidemiologic data suggested that people who used sunscreens were more likely to develop skin cancer, not less (Bastuji-Garin and Diepgen 2002). Part of this controversy can be traced to the fact that the early sunscreens were designed to block UVB, and not UVA, because erythema (the reddening accompanying burning) is caused by UVB. If in fact UVA played a significant role in melanomagenesis, early epidemiologic data would reflect the failure of early sunscreens to effectively filter out this waveband. However, the role of UVA in melanoma was itself controversial. Another interpretation of these data was that they reflected improper use of sunscreens in terms of frequency and thickness of sunscreen application by those spending significant time out in the sun. Recent studies have started to shed light on these controversies. Klug et al. (2010) performed a matched logistic regression analysis of sunscreen use adjusted for average UVB intensity, sun exposure, skin type, number of sunburns, age group, study site, and gender to show that sunscreen use was not associated with melanomagenesis. Moreover, utilizing the HGF transgenic mouse model of UVR-induced melanomagenesis, they showed that SPF-15 sunscreen-treated mice developed significantly fewer melanomas than the vehicle-treated control mice; and this effect was associated with significant reduction in UVB-induced DNA damage, indicating a distinct protective effect of sunscreen (Klug et al. 2010). Using a Braf-mutant mouse model, Viros et al. showed evidence that sunscreen at least partially protected against UVR-induced melanomagenesis (Viros et al. 2014). In another in vivo study using the UVR-susceptible NRas61R mutant mouse model of melanoma, Hennessey et al. recently showed that aerosol sunscreens effectively blocked UVR-induced DNA damage and delayed melanoma formation (Hennessey et al. 2017). Yet another controversy surrounded the opposing values of restricting sunlight to prevent the development of melanoma and other skin cancers and of receiving enough sunlight to ensure optimal production of vitamin D, which in adequate amounts prevents a variety of diseases (Moan et

M. R. Zaidi et al.

al. 2008b). For example, in a post-hoc analysis of a randomized controlled trial, vitamin D supplementation was not associated with reduced melanoma risk (Tang et al. 2011). On the other hand, vitamin D deficiency has been shown to be associated with a worse prognosis in metastatic melanoma patients (Timerman et al. 2017). Extensive studies in the last two decades have provided strong evidence that vitamin D deficiency is associated with reduced survival of melanoma patients and single-nucleotide polymorphisms (SNPs) in the vitamin D receptor (VDR) gene affect melanomagenesis and/or disease outcome (Slominski et al. 2017). Active forms of vitamin D3 protect against oxidative DNA damage by upregulating genes such as TP53 and GADD45 (Fleet et al. 2012), and its newly discovered derivatives exhibit VDR-mediated antiproliferative effects on melanoma cell lines (Janjetovic et al. 2011; Slominski et al. 2012). Based on these and other extensive studies, multiple randomized clinical trials are currently underway to evaluate whether vitamin D supplementation improves survival outcome in melanoma patients. Importantly, however, these questions do not lead to the conclusion that exposure to UVR would be an optimal means of maintaining healthy vitamin D levels, as oral vitamin supplements are effective, inexpensive, and unlike UVR, not associated with any known carcinogenic activity.

Photobiology and Melanoma UVB versus UVA To begin to understand how UVR affects biological systems (photobiology), one needs to appreciate the complexity of UVR itself. UVR from sunlight is subdivided into shorter wavelength UVC (200–280 nm), middle wavelength UVB (280–320 nm), and longer wavelength UVA (320–400 nm). UVC is biologically irrelevant, because it is absorbed by the atmospheric ozone layer. UVB and UVA are very different in their properties and biologic consequences. UVB penetrates poorly into the skin but is efficiently absorbed directly by DNA and proteins (Bruls et al. 1984). The two major types of DNA damage induced by UVB are cyclobutane

Biology of Melanocytes and Primary Melanoma

pyrimidine dimers (CPDs) and 6–4 photoproducts, which are repaired via the nucleotide excision repair pathway. The UVB-induced CPDs are characterized by signature C to T or CC to TT transitions. However, in melanomas such CPD signature mutations are relatively rare in the oncogenes or tumor suppressors that have been implicated in melanomagenesis. For example, 90% of BRAF mutations contain a T to A transversion (T1799A) (Wan et al. 2004). Only UVB can cause skin erythema. In contrast to shortwave UVB, UVA penetrates much more deeply into the skin but is poorly absorbed by DNA. UVA does not directly damage DNA, instead exerting its effects through photosensitizers. Thus, UVA induces oxidative DNA damage, such as the formation of 8-oxo-7,8-dihydroguanine through the generation of species such as reactive melanin radicals (Hill and Hill 2000). UVR and ROS In eukaryotic cells, mitochondrial-mediated oxidative phosphorylation produces energy and, because electron transfer is not completely efficient, also generates potentially harmful side products called reactive oxygen species (ROS) (Wittgen and van Kempen 2007). Some ROS are required as secondary messengers of signaling cascades and for other normal cell functions, so cells have developed an antioxidant network that allows excess ROS to be scavenged, resulting in a state of homeostasis. However, melanocytes face a daunting ROS burden because of their ability to produce melanin and because they are routinely exposed to UVR, a situation that can have melanomagenic consequences. UVR stimulates the synthesis of melanin that, although designed for protection from further UVR damage, can become a prooxidant under oxidative stress conditions. Specifically, UVR exposure can induce the formation of oxidized melanin, forming a variety of radicals and disrupting ROS homeostasis. UVR, and UVA in particular, can damage DNA through the production of these melaninderived radical species. Using the HGF transgenic mouse model, Noonan et al. (2012) showed that UVA was indeed melanomagenic but required the presence of melanin pigment and was associated with oxidative DNA damage, whereas UVB-induced melanomagenesis was independent of melanin.

29

Interestingly, ROS activity is related to the hypoxic response, mediated by the transcription factor hypoxia-inducible factor-1 (HIF-1α/β) (Wittgen and van Kempen 2007). The epidermis of the normal skin exists in a mildly hypoxic environment, an environment that can promote melanomagenesis. Bedogni et al. (2005) demonstrated that AKT-mediated transformation of melanocytes only occurs under HIF-lα-inducing hypoxic conditions similar to those found in the normal skin. In addition, AKT hyperactivation can bring about ROS generation by inducing NOX4 and by inhibiting apoptosis, thereby damaging mitochondria (Govindarajan et al. 2007). UVR and the microenvironment UVR is known to induce an increase in the number of melanocytes and to enhance melanin synthesis, and ultimately UVR stimulates an increased transfer of melanin to those keratinocytes that it contacts (Miyamura et al. 2007; Zaidi et al. 2011). UVR can alter the melanocyte microenvironment, which subsequently influences the melanocytic response. D’Orazio et al. (2006) reported that UVR induces enhanced synthesis and secretion of αMSH by keratinocytes, which in turn triggers an appropriate melanocyte response through MC1R and the cAMP pathway, activating melaninsynthesizing enzymes and melanin production. In fact, UVR affects most of the components of the skin including melanocytes, keratinocytes, fibroblasts, and elements of the vascular and immune systems (Slominski and Pawelek 1998). UVR induces secretion of multiple factors by multiple cellular components, including neuropeptides (POMCderived αMSH, ACTH, and β-endorphin), growth factors (FGF2, IGF-1, TGFα), and numerous cytokines (TNF-α, CSF2, IL-1, IL-6, IL-8, IL-10) (Slominski and Pawelek 1998). UVR triggers inflammatory reactions and has an overall immunoinhibitory effect on the skin (Hart and Norval 2017; Kripke 1990). Moreover, the effect of UVR on the immune system in mouse models may be most influential at the neonatal stage, perhaps explaining the elevated risk associated with childhood UVR exposure (Zaidi et al. 2011, 2012). More specifically, UVR suppresses the ability of antigen-presenting Langerhans cells

30

in the skin to induce immunity by decreasing their number and their expression of MHC. UVR also stimulates an influx of macrophages and a decrease in the viability of T-lymphocytes. It has also been shown that inflammation can have positive influences on tumorigenesis. Together, it is evident that the abilities of UVR to induce mutagenesis, inflammation, and long-term immunosuppression can all profoundly influence melanomagenesis. Importantly, however, there is evidence to suggest some UVR-independent melanoma risk in certain circumstances. In individuals with the red-hair/light-skin phenotype, melanoma risk is elevated in both sunexposed and non-sun-exposed skin (in contrast to their risk of non-melanoma skin cancers whose incidence is elevated only in sun-exposed skin). Mitra et al. (2012) utilized mouse models of red-hair/Mc1r-deficiency to show that red pigment (pheomelanin) is intrinsically carcinogenic – elevating the spontaneous incidence of BRAFV600Eassociated melanomas in the complete absence of UVR exposure. When red-haired mice were crossed to albinos (carrying Tyr mutations, thus unable to make any melanins), the melanoma risk was fully blocked. Significant levels of pheomelanin are dominantly present in the skin of not only redheads but also other Caucasians, suggesting that it may contribute chemically to melanomagenesis. However, it is also very likely that UVR amplifies the ROSinducing capacity of pheomelanin, suggesting that UVR significantly exacerbates skin carcinogenicity in humans with fair skin.

Conclusion It has been more than 200 years since the Parisian physician René Laennec first reported “melanoma” in Europe, while back in 1857 William Norris (1857) recognized that melanoma had a familial component. It is unlikely that either could have foreseen that by the twenty-first century, cutaneous malignant melanoma would become one of the fastest increasing cancers. In the United States alone, melanoma incidence has more than quadrupled over the past four decades, and it is estimated that more than 90,000 new cases will be diagnosed in 2018 (Cronin et al.

M. R. Zaidi et al.

2018). Melanoma remains one of only a few major malignancies demonstrating significant positive rates of increase in the United States. We have come a long way with respect to our understanding of the origins of melanoma and the mechanisms by which it develops. As documented in this chapter, the first decade of the twenty-first century featured tremendous progress in the molecular technologic weaponry with which to address the most outstanding questions about the melanoma disease initiation, progression, and metastasis. Utilizing these impressive tools, researchers generated a treasure trove of data, gathered from both broadly global and exquisitely detailed analyses of melanocytic lesions from patients and from genetically engineered mouse models. In the second decade of this century, these data have advanced our understanding of the fundamental mechanisms of the disease, which has fueled an explosive expansion of therapeutic strategies that have revolutionized the clinical management of melanoma. There is excited anticipation within the melanoma research community that the holy grail of eliminating melanoma-related death is within reach. Acknowledgments This chapter has been updated from the previous edition, which was written by Drs. Glenn Merlino and Vince Hearing. Because of space limitations, we have relied heavily on the citation of many review articles and therefore apologize to the authors of numerous significant primary papers we were unable to reference. M.R.Z. acknowledges research support from the National Institutes of Health (K22CA163799 and R01CA193711), US Department of Defense (CA150492), Melanoma Research Foundation, and W.W. Smith Charitable Trust. D.E.F. acknowledges research support from the National Institutes of Health (5P01 CA163222, 1R01CA222871, R01AR072304, and 5R01 AR043369-22), the Melanoma Research Alliance, and the Dr. Miriam and Sheldon G. Adelson Medical Research Foundation. D.E.F. has a financial interest in Soltego, Inc., a company developing SIK inhibitors for topical skin-darkening treatments.

References Ackermann J, Frutschi M, Kaloulis K, McKee T, Trumpp A, Beermann F (2005) Metastasizing melanoma formation caused by expression of activated N-RasQ61K on an INK4a-deficient background. Cancer Res

Biology of Melanocytes and Primary Melanoma 65:4005–4011. https://doi.org/10.1158/0008-5472. CAN-04-2970 Aguissa-Toure AH, Li G (2012) Genetic alterations of PTEN in human melanoma. Cell Mol Life Sci 69:1475–1491. https://doi.org/10.1007/s00018011-0878-0 Ainger SA, Jagirdar K, Lee KJ, Soyer HP, Sturm RA (2017) Skin pigmentation genetics for the clinic. Dermatology 233:1–15. https://doi.org/10.1159/0004 68538 Aoude LG, Wadt KA, Pritchard AL, Hayward NK (2015) Genetics of familial melanoma: 20 years after CDKN2A. Pigment Cell Melanoma Res 28:148–160. https://doi.org/10.1111/pcmr.12333 Barral DC, Seabra MC (2004) The melanosome as a model to study organelle motility in mammals. Pigment Cell Res 17:111–118. https://doi.org/10.1111/j.16000749.2004.00138.x Bastian BC (2014) The molecular pathology of melanoma: an integrated taxonomy of melanocytic neoplasia. Annu Rev Pathol 9:239–271. https://doi.org/10.1146/ annurev-pathol-012513-104658 Bastuji-Garin S, Diepgen TL (2002) Cutaneous malignant melanoma, sun exposure, and sunscreen use: epidemiological evidence. Br J Dermatol 146(Suppl 61):24–30 Batlle E, Clevers H (2017) Cancer stem cells revisited. Nat Med 23:1124–1134. https://doi.org/10.1038/nm.4409 Bauer J, Bastian BC (2006) Distinguishing melanocytic nevi from melanoma by DNA copy number changes: comparative genomic hybridization as a research and diagnostic tool. Dermatol Ther 19:40–49. https://doi.org/10.1111/j.1529-8019.2005.00055.x Bauer J, Curtin JA, Pinkel D, Bastian BC (2007) Congenital melanocytic nevi frequently harbor NRAS mutations but no BRAF mutations. J Invest Dermatol 127:179–182. https://doi.org/10.1038/sj.jid.5700490 Baxter LL, Pavan WJ (2013) The etiology and molecular genetics of human pigmentation disorders. Wiley Interdiscip Rev Dev Biol 2:379–392. https://doi.org/ 10.1002/wdev.72 Baxter LL, Watkins-Chow DE, Pavan WJ, Loftus SK (2018) A curated gene list for expanding the horizons of pigmentation biology Pigment Cell & Melanoma Research 32(3):348–358. https://doi.org/10.1111/ pcmr.12743 Beauvais-Jouneau A et al (1999) A novel model to study the dorsolateral migration of melanoblasts. Mech Dev 89:3–14 Bedogni B, Welford SM, Cassarino DS, Nickoloff BJ, Giaccia AJ, Powell MB (2005) The hypoxic microenvironment of the skin contributes to Akt-mediated melanocyte transformation. Cancer Cell 8:443–454. https://doi.org/10.1016/j.ccr.2005.11.005 Bennett DC (2008) How to make a melanoma: what do we know of the primary clonal events? Pigment Cell Melanoma Res 21:27–38. https://doi.org/10.1111/j.1755148X.2007.00433.x Ben-Porath I, Weinberg RA (2005) The signals and pathways activating cellular senescence. Int J Biochem

31 Cell Biol 37:961–976. https://doi.org/10.1016/j.bioce l.2004.10.013 Berson JF, Harper DC, Tenza D, Raposo G, Marks MS (2001) Pmel17 initiates premelanosome morphogenesis within multivesicular bodies. Mol Biol Cell 12:3451–3464. https://doi.org/10.1091/mbc.12. 11.3451 Bharti K, Nguyen MT, Skuntz S, Bertuzzi S, Arnheiter H (2006) The other pigment cell: specification and development of the pigmented epithelium of the vertebrate eye. Pigment Cell Res 19:380–394. https://doi.org/ 10.1111/j.1600-0749.2006.00318.x Blake JA, Ziman MR (2005) Pax3 transcripts in melanoblast development. Develop Growth Differ 47:627–635. https://doi.org/10.1111/j.1440-169X.200 5.00835.x Bonaventure J, Domingues MJ, Larue L (2013) Cellular and molecular mechanisms controlling the migration of melanocytes and melanoma cells. Pigment Cell Melanoma Res 26:316–325. https://doi.org/10.1111/ pcmr.12080 Borgdorff V et al (2014) A chemical biology approach identifies AMPK as a modulator of melanoma oncogene MITF. Oncogene 33:2531–2539. https:// doi.org/10.1038/onc.2013.185 Botchkarev VA et al (2006) Neurotrophins in skin biology and pathology. J Invest Dermatol 126:1719–1727. https://doi.org/10.1038/sj.jid.5700270 Braeuer RR, Zigler M, Villares GJ, Dobroff AS, Bar-Eli M (2011) Transcriptional control of melanoma metastasis: the importance of the tumor microenvironment. Semin Cancer Biol 21:83–88. https://doi.org/10.1016/ j.semcancer.2010.12.007 Brenner M, Hearing VJ (2008) Modifying skin pigmentation-approaches through intrinsic biochemistry and exogenous agents. Drug Discov Today 5:e189–e199 Bronner ME, LeDouarin NM (2012) Development and evolution of the neural crest: an overview. Dev Biol 366:2–9. https://doi.org/10.1016/j.ydbio.2011.12.042 Bruls WA, van Weelden H, van der Leun JC (1984) Transmission of UV-radiation through human epidermal layers as a factor influencing the minimal erythema dose. Photochem Photobiol 39:63–67 Buac K, Pavan WJ (2007) Stem cells of the melanocyte lineage. Cancer Biomark 3:203–209 Buchbinder EI et al (2015) Phase 2 study of sunitinib in patients with metastatic mucosal or acral melanoma. Cancer 121:4007–4015. https://doi.org/10.1002/ cncr.29622 Burd CE et al (2014) Mutation-specific RAS oncogenicity explains NRAS codon 61 selection in melanoma. Cancer Discov 4:1418–1429. https://doi.org/10.1158/ 2159-8290.CD-14-0729 Cancer Genome Atlas N (2015) Genomic classification of cutaneous melanoma. Cell 161:1681–1696. https://doi.org/10.1016/j.cell.2015.05.044 Carbone M, Yang H, Pass HI, Krausz T, Testa JR, Gaudino G (2013) BAP1 and cancer. Nat Rev Cancer 13:153–159. https://doi.org/10.1038/nrc3459

32 Carvajal RD et al (2011) KIT as a therapeutic target in metastatic melanoma. JAMA 305:2327–2334. https://doi.org/10.1001/jama.2011.746 Chen H, Weng QY, Fisher DE (2014) UV signaling pathways within the skin. J Invest Dermatol 134:2080–2085. https://doi.org/10.1038/jid.2014.161 Chi A et al (2006) Proteomic and bioinformatic characterization of the biogenesis and function of melanosomes. J Proteome Res 5:3135–3144. https://doi.org/10.1021/ pr060363j Chin L et al (1997) Cooperative effects of INK4a and ras in melanoma susceptibility in vivo. Genes Dev 11:2822–2834 Chin L, Garraway LA, Fisher DE (2006) Malignant melanoma: genetics and therapeutics in the genomic era. Genes Dev 20:2149–2182. https://doi.org/10.1101/ gad.1437206 Chudnovsky Y, Adams AE, Robbins PB, Lin Q, Khavari PA (2005) Use of human tissue to assess the oncogenic activity of melanoma-associated mutations. Nat Genet 37:745–749. https://doi.org/10.1038/ ng1586 Clark WH Jr, Elder DE, Guerry D, Epstein MN, Greene MH, Van Horn M (1984) A study of tumor progression: the precursor lesions of superficial spreading and nodular melanoma. Hum Pathol 15:1147–1165 Coelho SG et al (2009) Short- and long-term effects of UV radiation on the pigmentation of human skin. J Investig Dermatol Symp Proc 14:32–35. https://doi.org/ 10.1038/jidsymp.2009.10 Costin GE, Hearing VJ (2007) Human skin pigmentation: melanocytes modulate skin color in response to stress. FASEB J 21:976–994. https://doi.org/10.1096/fj.066649rev Costin GE, Valencia JC, Vieira WD, Lamoreux ML, Hearing VJ (2003) Tyrosinase processing and intracellular trafficking is disrupted in mouse primary melanocytes carrying the underwhite (uw) mutation. A model for oculocutaneous albinism (OCA) type 4. J Cell Sci 116:3203–3212. https://doi.org/10.1242/jcs.00598 Coudrier E (2007) Myosins in melanocytes: to move or not to move? Pigment Cell Res 20:153–160. https://doi.org/10.1111/j.1600-0749.2007.00376.x Cronin KA et al (2018) Annual report to the nation on the status of Cancer, part I: national cancer statistics. Cancer 124:2785–2800. https://doi.org/10.1 002/cncr.31551 Crowson AN, Magro C, Miller A, Mihm MC Jr (2007) The molecular basis of melanomagenesis and the metastatic phenotype. Semin Oncol 34:476–490. https://doi.org/10.1053/j.seminoncol.2007.09.007 Cui R et al (2007) Central role of p53 in the suntan response and pathologic hyperpigmentation. Cell 128:853–864. https://doi.org/10.1016/j.cell.200 6.12.045 Cummings SD, Ryu B, Samuels MA, Yu X, Meeker AK, Healey MA, Alani RM (2008) Id1 delays senescence of primary human melanocytes. Mol Carcinog 47:653–659. https://doi.org/10.1002/mc.20422

M. R. Zaidi et al. Curtin JA et al (2005) Distinct sets of genetic alterations in melanoma. N Engl J Med 353:2135–2147. https://doi. org/10.1056/NEJMoa050092 Curtin JA, Busam K, Pinkel D, Bastian BC (2006) Somatic activation of KIT in distinct subtypes of melanoma. J Clin Oncol 24:4340–4346. https://doi.org/10.1200/ JCO.2006.06.2984 D’Orazio JA et al (2006) Topical drug rescue strategy and skin protection based on the role of Mc1r in UVinduced tanning. Nature 443:340–344. https://doi.org/ 10.1038/nature05098 Dahl C, Guldberg P (2007) The genome and epigenome of malignant melanoma. APMIS 115:1161–1176. https:// doi.org/10.1111/j.1600-0463.2007.apm_855.xml.x Damsky WE, Bosenberg M (2017) Melanocytic nevi and melanoma: unraveling a complex relationship. Oncogene 36:5771–5792. https://doi.org/10.1038/onc. 2017.189 Damsky W et al (2015) mTORC1 activation blocks BrafV600E-induced growth arrest but is insufficient for melanoma formation. Cancer Cell 27:41–56. https://doi.org/10.1016/j.ccell.2014.11.014 Davies H et al (2002) Mutations of the BRAF gene in human cancer. Nature 417:949–954. https://doi.org/ 10.1038/nature00766 Dell’Angelica EC (2003) Melanosome biogenesis: shedding light on the origin of an obscure organelle. Trends Cell Biol 13:503–506 Delmas V et al (2007) Beta-catenin induces immortalization of melanocytes by suppressing p16INK4a expression and cooperates with N-Ras in melanoma development. Genes Dev 21:2923–2935. https://doi. org/10.1101/gad.450107 Dhomen N et al (2009) Oncogenic Braf induces melanocyte senescence and melanoma in mice Cancer. Cell 15:294–303. https://doi.org/10.1016/j. ccr.2009.02.022 Di Leva G, Garofalo M, Croce CM (2014) MicroRNAs in cancer. Annu Rev Pathol 9:287–314. https://doi.org/ 10.1146/annurev-pathol-012513-104715 Diwakar G, Zhang D, Jiang S, Hornyak TJ (2008) Neurofibromin as a regulator of melanocyte development and differentiation. J Cell Sci 121:167–177. https://doi.org/10.1242/jcs.013912 Dorsky RI, Raible DW, Moon RT (2000) Direct regulation of nacre, a zebrafish MITF homolog required for pigment cell formation, by the Wnt pathway. Genes Dev 14:158–162 Dunn KJ, Brady M, Ochsenbauer-Jambor C, Snyder S, Incao A, Pavan WJ (2005) WNT1 and WNT3a promote expansion of melanocytes through distinct modes of action. Pigment Cell Res 18:167–180. https://doi. org/10.1111/j.1600-0749.2005.00226.x Edward M, MacKie RM (1993) Cell-cell and cell-extracellular matrix interactions during melanoma cell invasion and metastasis. Melanoma Res 3:227–234 Eller MS, Gilchrest BA (2000) Tanning as part of the eukaryotic SOS response. Pigment Cell Res 13(Suppl 8):94–97

Biology of Melanocytes and Primary Melanoma Fang D et al (2005) A tumorigenic subpopulation with stem cell properties in melanomas. Cancer Res 65:9328–9337. https://doi.org/10.1158/00085472.CAN-05-1343 Fattore L et al (2017) MicroRNAs in melanoma development and resistance to target therapy. Oncotarget 8:22262–22278. https://doi.org/10.18632/ oncotarget.14763 Fell GL, Robinson KC, Mao J, Woolf CJ, Fisher DE (2014) Skin beta-endorphin mediates addiction to UV light. Cell 157:1527–1534. https://doi.org/10.1016/j. cell.2014.04.032 Fleet JC, DeSmet M, Johnson R, Li Y (2012) Vitamin D and cancer: a review of molecular mechanisms. Biochem J 441:61–76. https://doi.org/10.1042/BJ20 110744 Frank NY et al (2005) ABCB5-mediated doxorubicin transport and chemoresistance in human malignant melanoma. Cancer Res 65:4320–4333. https://doi.org/ 10.1158/0008-5472.CAN-04-3327 Fulda S (2015) Targeting apoptosis for anticancer therapy. Semin Cancer Biol 31:84–88. https://doi.org/10.1016/j. semcancer.2014.05.002 Gaffal E, Landsberg J, Bald T, Sporleder A, Kohlmeyer J, Tuting T (2011) Neonatal UVB exposure accelerates melanoma growth and enhances distant metastases in Hgf-Cdk4(R24C) C57BL/6 mice. Int J Cancer 129:285–294. https://doi.org/10.1002/ijc.25913 Garcia-Castro MI, Marcelle C, Bronner-Fraser M (2002) Ectodermal Wnt function as a neural crest inducer. Science 297:848–851 Garibyan L, Fisher DE (2010) How sunlight causes melanoma. Curr Oncol Rep 12:319–326. https://doi.org/ 10.1007/s11912-010-0119-y Garnett MJ, Marais R (2004) Guilty as charged: B-RAF is a human oncogene. Cancer Cell 6:313–319. https://doi. org/10.1016/j.ccr.2004.09.022 Garraway LA, Sellers WR (2006) Lineage dependency and lineage-survival oncogenes in human cancer. Nat Rev Cancer 6:593–602. https://doi.org/10.1038/nrc1947 Garraway LA et al (2005) Integrative genomic analyses identify MITF as a lineage survival oncogene amplified in malignant melanoma. Nature 436:117–122. https:// doi.org/10.1038/nature03664 Ginger RS et al (2008) SLC24A5 encodes a trans-Golgi network protein with potassium-dependent sodium-calcium exchange activity that regulates human epidermal melanogenesis. J Biol Chem 283:5486–5495. https:// doi.org/10.1074/jbc.M707521200 Goel VK et al (2009) Melanocytic nevus-like hyperplasia and melanoma in transgenic BRAFV600E mice. Oncogene 28:2289–2298. https://doi.org/10.1038/onc. 2009.95 Goldstein AM, Tucker MA (2013) Dysplastic nevi and melanoma. Cancer Epidemiol Biomark Prev 22:528–532. https://doi.org/10.1158/1055-9965. EPI-12-1346 Govindarajan B et al (2007) Overexpression of Akt converts radial growth melanoma to vertical

33 growth melanoma. J Clin Invest 117:719–729. https://doi.org/10.1172/JCI30102 Graf J, Hodgson R, van Daal A (2005) Single nucleotide polymorphisms in the MATP gene are associated with normal human pigmentation variation. Hum Mutat 25:278–284. https://doi.org/10.1002/humu.20143 Green AC, Wallingford SC, McBride P (2011) Childhood exposure to ultraviolet radiation and harmful skin effects: epidemiological evidence. Prog Biophys Mol Biol 107:349–355. https://doi.org/10.1016/j.pbio molbio.2011.08.010 Grossman D, McNiff JM, Li F, Altieri DC (1999) Expression and targeting of the apoptosis inhibitor, survivin, in human melanoma. J Invest Dermatol 113:1076–1081. https://doi.org/10.1046/j.1523-1747.1 999.00776.x Guo J et al (2011) Phase II, open-label, single-arm trial of imatinib mesylate in patients with metastatic melanoma harboring c-Kit mutation or amplification. J Clin Oncol 29:2904–2909. https://doi.org/10.1200/ JCO.2010.33.9275 Gupta PB et al (2005) The melanocyte differentiation program predisposes to metastasis after neoplastic transformation. Nat Genet 37:1047–1054. https://doi.org/ 10.1038/ng1634 Guth SI, Wegner M (2008) Having it both ways: sox protein function between conservation and innovation. Cell Mol Life Sci 65:3000–3018. https://doi.org/ 10.1007/s00018-008-8138-7 Guyonneau L, Murisier F, Rossier A, Moulin A, Beermann F (2004) Melanocytes and pigmentation are affected in dopachrome tautomerase knockout mice. Mol Cell Biol 24:3396–3403 Ha L et al (2007) ARF functions as a melanoma tumor suppressor by inducing p53-independent senescence. Proc Natl Acad Sci U S A 104:10968–10973. https://doi.org/10.1073/pnas.0611638104 Haass NK, Herlyn M (2005) Normal human melanocyte homeostasis as a paradigm for understanding melanoma. J Investig Dermatol Symp Proc 10:153–163. https://doi.org/10.1111/j.1087-0024.2005.200407.x Haass NK, Smalley KS, Li L, Herlyn M (2005) Adhesion, migration and communication in melanocytes and melanoma. Pigment Cell Res 18:150–159. https://doi.org/ 10.1111/j.1600-0749.2005.00235.x Hanahan D, Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144:646–674. https://doi.org/ 10.1016/j.cell.2011.02.013 Haqq C et al (2005) The gene expression signatures of melanoma progression. Proc Natl Acad Sci USA 102: 6092–6097. https://doi.org/10.1073/pnas.0501564102 Hari L et al (2012) Temporal control of neural crest lineage generation by Wnt/beta-catenin signaling. Development 139:2107–2117. https://doi.org/10.124 2/dev.073064 Harris ML, Baxter LL, Loftus SK, Pavan WJ (2010) Sox proteins in melanocyte development and melanoma. Pigment Cell Melanoma Res 23:496–513. https://doi. org/10.1111/j.1755-148X.2010.00711.x

34 Hart PH, Norval M (2017) Ultraviolet radiation-induced immunosuppression and its relevance for skin carcinogenesis. Photochem Photobiol Sci. https://doi.org/ 10.1039/c7pp00312a Hearing VJ, Leong SPL (2006) From melanocytes to melanoma: the progression to malignancy. Humana Press, Totowa Hennessey RC et al (2017) Ultraviolet radiation accelerates NRas-mutant melanomagenesis: a cooperative effect blocked by sunscreen. Pigment Cell Melanoma Res 30:477–487. https://doi.org/10.1111/pcmr.12601 Herlyn M, Berking C, Li G, Satyamoorthy K (2000) Lessons from melanocyte development for understanding the biological events in naevus and melanoma formation. Melanoma Res 10:303–312 Herraiz C, Garcia-Borron JC, Jimenez-Cervantes C, Olivares C (2017) MC1R signaling. Intracellular partners and pathophysiological implications. Biochim Biophys Acta 1863:2448–2461. https://doi.org/10.1 016/j.bbadis.2017.02.027 Hill HZ, Hill GJ (2000) UVA, pheomelanin and the carcinogenesis of melanoma. Pigment Cell Res 13(Suppl 8):140–144 Hirobe T (2011) How are proliferation and differentiation of melanocytes regulated? Pigment Cell Melanoma Res 24:462–478. https://doi.org/10.1111/j.1755-148X. 2011.00845.x Hirobe T, Ootaka H (2007) Interleukin-1alpha stimulates the differentiation of melanocytes but inhibits the proliferation of melanoblasts from neonatal mouse epidermis. Zool Sci 24:959–970. https://doi.org/10.2108/ zsj.24.959 Hirobe T, Osawa M, Nishikawa S (2004) Hepatocyte growth factor controls the proliferation of cultured epidermal melanoblasts and melanocytes from newborn mice. Pigment Cell Res 17:51–61 Hoashi T, Watabe H, Muller J, Yamaguchi Y, Vieira WD, Hearing VJ (2005) MART-1 is required for the function of the melanosomal matrix protein PMEL17/ GP100 and the maturation of melanosomes. J Biol Chem 280:14006–14016. https://doi.org/10.1074/jbc. M413692200 Hodi FS et al (2013) Imatinib for melanomas harboring mutationally activated or amplified KIT arising on mucosal, acral, and chronically sun-damaged skin. J Clin Oncol 31:3182–3190. https://doi.org/10.1200/ JCO.2012.47.7836 Hoek KS (2007) DNA microarray analyses of melanoma gene expression: a decade in the mines. Pigment Cell Res 20:466–484. https://doi.org/10.1111/j.16000749.2007.00412.x Hoek K et al (2004) Expression profiling reveals novel pathways in the transformation of melanocytes to melanomas. Cancer Res 64:5270–5282. https://doi.org/ 10.1158/0008-5472.CAN-04-0731 Hofmann UB, Westphal JR, Van Kraats AA, Ruiter DJ, Van Muijen GN (2000) Expression of integrin alpha(v) beta(3) correlates with activation of membrane-type matrix metalloproteinase-1 (MT1-MMP) and matrix

M. R. Zaidi et al. metalloproteinase-2 (MMP-2) in human melanoma cells in vitro and in vivo. Int J Cancer 87:12–19 Hopkins BD et al (2013) A secreted PTEN phosphatase that enters cells to alter signaling and survival. Science 341:399–402. https://doi.org/10.1126/science. 1234907 Hornyak TJ, Hayes DJ, Chiu LY, Ziff EB (2001) Transcription factors in melanocyte development: distinct roles for Pax-3 and Mitf. Mech Dev 101:47–59 Hu ZZ et al (2007) Comparative bioinformatics analyses and profiling of lysosome-related organelle proteomes. Int J Mass Spectrom 259:147–160. https://doi.org/ 10.1016/j.ijms.2006.09.024 Huarte M (2015) The emerging role of lncRNAs in cancer. Nat Med 21:1253–1261. https://doi.org/10.10 38/nm.3981 Hubbard JK, Uy JA, Hauber ME, Hoekstra HE, Safran RJ (2010) Vertebrate pigmentation: from underlying genes to adaptive function. Trends Genet 26:231–239. https://doi.org/10.1016/j.tig.2010.02.002 Hussein MR, Haemel AK, Wood GS (2003a) Apoptosis and melanoma: molecular mechanisms. J Pathol 199:275–288. https://doi.org/10.1002/path.1300 Hussein MR, Haemel AK, Wood GS (2003b) p53-related pathways and the molecular pathogenesis of melanoma. Eur J Cancer Prev 12:93–100. https://doi.org/ 10.1097/01.cej.0000062797.86004.35 Ignatius MS, Moose HE, El-Hodiri HM, Henion PD (2008) colgate/hdac1 repression of foxd3 expression is required to permit mitfa-dependent melanogenesis. Dev Biol 313:568–583. https://doi.org/10.1016/j. ydbio.2007.10.045 Imokawa G (2004) Autocrine and paracrine regulation of melanocytes in human skin and in pigmentary disorders. Pigment Cell Res 17:96–110. https://doi.org/ 10.1111/j.1600-0749.2003.00126.x Ito S, IFPCS (2003) The IFPCS presidential lecture: a chemist’s view of melanogenesis. Pigment Cell Res 16:230–236 Ito M et al (1999) Removal of stem cell factor or addition of monoclonal anti-c-KIT antibody induces apoptosis in murine melanocyte precursors. J Invest Dermatol 112:796–801. https://doi.org/10.1046/j.152 3-1747.1999.00552.x Ivanov VN, Bhoumik A, Ronai Z (2003) Death receptors and melanoma resistance to apoptosis. Oncogene 22:3152–3161. https://doi.org/10.1038/sj.onc.1206456 Jacobs JJ et al (2000) Senescence bypass screen identifies TBX2, which represses Cdkn2a (p19(ARF)) and is amplified in a subset of human breast cancers. Nat Genet 26:291–299. https://doi.org/10.1038/81583 Janjetovic Z et al (2011) High basal NF-kappaB activity in nonpigmented melanoma cells is associated with an enhanced sensitivity to vitamin D3 derivatives. Br J Cancer 105:1874–1884. https://doi.org/10.1038/ bjc.2011.458 Jhappan C, Noonan FP, Merlino G (2003) Ultraviolet radiation and cutaneous malignant melanoma. Oncogene 22:3099–3112. https://doi.org/10.1038/sj.onc.1206450

Biology of Melanocytes and Primary Melanoma Jiang R, Norton CR, Copeland NG, Gilbert DJ, Jenkins NA, Gridley T (1998) Genomic organization, expression and chromosomal localization of the mouse Slug (Slugh) gene. Biochim Biophys Acta 1443:251–254 Jin EJ, Erickson CA, Takada S, Burrus LW (2001) Wnt and BMP signaling govern lineage segregation of melanocytes in the avian embryo. Dev Biol 233:22–37. https://doi.org/10.1006/dbio.2001.0222 Jonsson G et al (2007) Genomic profiling of malignant melanoma using tiling-resolution arrayCGH. Oncogene 26:4738–4748. https://doi.org/10.1038/sj. onc.1210252 Jouneau A, Yu YQ, Pasdar M, Larue L (2000) Plasticity of cadherin-catenin expression in the melanocyte lineage. Pigment Cell Res 13:260–272 Kadekaro AL, Kavanagh RJ, Wakamatsu K, Ito S, Pipitone MA, Abdel-Malek ZA (2003) Cutaneous photobiology. The melanocyte vs. the sun: who will win the final round? Pigment Cell Res 16:434–447 Kamb A et al (1994) Analysis of the p16 gene (CDKN2) as a candidate for the chromosome 9p melanoma susceptibility locus. Nat Genet 8:23–26. https://doi.org/ 10.1038/ng0994-22 Kasof GM, Gomes BC (2001) Livin, a novel inhibitor of apoptosis protein family member. J Biol Chem 276:3238–3246. https://doi.org/10.1074/jbc.M0 03670200 Kaufmann WK et al (2008) Defective cell cycle checkpoint functions in melanoma are associated with altered patterns of gene expression. J Invest Dermatol 128:175–187. https://doi.org/10.1038/sj.jid.5700935 Kaur A, Webster MR, Weeraratna AT (2016) In the Wnt-er of life: Wnt signalling in melanoma and ageing. Br J Cancer 115:1273–1279. https://doi.org/10.1038/ bjc.2016.332 Kawakami A, Fisher DE (2017) The master role of microphthalmia-associated transcription factor in melanocyte and melanoma biology. Lab Investig 97:649–656. https://doi.org/10.1038/labinvest.2017.9 Kawakami Y et al (2000) T cell immune responses against melanoma and melanocytes in cancer and autoimmunity. Pigment Cell Res 13(Suppl 8):163–169 Kim M et al (2006) Comparative oncogenomics identifies NEDD9 as a melanoma metastasis gene. Cell 125:1269–1281. https://doi.org/10.1016/j.cell.200 6.06.008 Klug HL et al (2010) Sunscreen prevention of melanoma in man and mouse. Pigment Cell Melanoma Res 23:835–837. https://doi.org/10.1111/j.1755-148X. 2010.00756.x Kohler C et al (2017) Mouse cutaneous melanoma induced by mutant BRaf arises from expansion and dedifferentiation of mature pigmented melanocytes. Cell Stem Cell 21(679-693):e676. https://doi.org/10.1016/ j.stem.2017.08.003 Kos L, Aronzon A, Takayama H, Maina F, Ponzetto C, Merlino G, Pavan W (1999) Hepatocyte growth factor/scatter factor-MET signaling in neural crest-

35 derived melanocyte development. Pigment Cell Res 12:13–21 Kos R, Reedy MV, Johnson RL, Erickson CA (2001) The winged-helix transcription factor FoxD3 is important for establishing the neural crest lineage and repressing melanogenesis in avian embryos. Development 128:1467–1479 Kripke ML (1990) Effects of UV radiation on tumor immunity. J Natl Cancer Inst 82:1392–1396 Kuphal S, Bauer R, Bosserhoff AK (2005) Integrin signaling in malignant melanoma. Cancer Metastasis Rev 24:195–222. https://doi.org/10.1007/s10555-0051572-1 Kwong LN et al (2012) Oncogenic NRAS signaling differentially regulates survival and proliferation in melanoma. Nat Med 18:1503–1510. https://doi.org/ 10.1038/nm.2941 LaBonne C, Bronner-Fraser M (1998) Neural crest induction in Xenopus: evidence for a two-signal model. Development 125:2403–2414 Lamason RL et al (2005) SLC24A5, a putative cation exchanger, affects pigmentation in zebrafish and humans. Science 310:1782–1786. https://doi.org/ 10.1126/science.1116238 Landi MT et al (2006) MC1R germline variants confer risk for BRAF-mutant melanoma. Science 313:521–522. https://doi.org/10.1126/science.1127515 Lang D et al (2005) Pax3 functions at a nodal point in melanocyte stem cell differentiation. Nature 433:884–887. https://doi.org/10.1038/nature03292 Larue L, Delmas V (2006) The WNT/Beta-catenin pathway in melanoma. Front Biosci 11:733–742 Lavado A, Montoliu L (2006) New animal models to study the role of tyrosinase in normal retinal development. Front Biosci 11:743–752 Lee HO, Levorse JM, Shin MK (2003) The endothelin receptor-B is required for the migration of neural crest-derived melanocyte and enteric neuron precursors. Dev Biol 259:162–175 Lee YR, Chen M, Pandolfi PP (2018) The functions and regulation of the PTEN tumour suppressor: new modes and prospects. Nat Rev Mol Cell Biol. https://doi.org/ 10.1038/s41580-018-0015-0 Leucci E, Coe EA, Marine JC, Vance KW (2016) The emerging role of long non-coding RNAs in cutaneous melanoma. Pigment Cell Melanoma Res 29:619–626. https://doi.org/10.1111/pcmr.12537 Lewis JL, Bonner J, Modrell M, Ragland JW, Moon RT, Dorsky RI, Raible DW (2004) Reiterated Wnt signaling during zebrafish neural crest development. Development 131:1299–1308. https://doi.org/10.1242/ dev.01007 Liebermann DA, Hoffman B, Vesely D (2007) p53 induced growth arrest versus apoptosis and its modulation by survival cytokines. Cell Cycle 6:166–170. https://doi.org/10.4161/cc.6.2.3789 Lin JY, Fisher DE (2007) Melanocyte biology and skin pigmentation. Nature 445:843–850. https://doi.org/10. 1038/nature05660

36 Lin X et al (2017) C-myc overexpression drives melanoma metastasis by promoting vasculogenic mimicry via cmyc/snail/Bax signaling. J Mol Med (Berl) 95:53–67. https://doi.org/10.1007/s00109-016-1452-x Lindsay CR et al (2011) P-Rex1 is required for efficient melanoblast migration and melanoma metastasis. Nat Commun 2:555. https://doi.org/10.1038/ncomms 1560 Liu W, Sharpless NE (2012) Senescence-escape in melanoma. Pigment Cell Melanoma Res 25:408–409 Liu ZJ et al (2006) Notch1 signaling promotes primary melanoma progression by activating mitogen-activated protein kinase/phosphatidylinositol 3-kinase-Akt pathways and up-regulating N-cadherin expression. Cancer Res 66:4182–4190. https://doi.org/10.1158/00085472.CAN-05-3589 Liu J, Fukunaga-Kalabis M, Li L, Herlyn M (2014) Developmental pathways activated in melanocytes and melanoma. Arch Biochem Biophys 563:13–21. https://doi.org/10.1016/j.abb.2014.07.023 Liu XL, Ding J, Meng LH (2018) Oncogene-induced senescence: a double edged sword in cancer. Acta Pharmacol Sin. https://doi.org/10.1038/aps.2017.198 Lowe SW, Lin AW (2000) Apoptosis in cancer. Carcinogenesis 21:485–495 Lowe SW, Cepero E, Evan G (2004) Intrinsic tumour suppression. Nature 432:307–315. https://doi.org/ 10.1038/nature03098 Maddodi N, Setaluri V (2008) Role of UV in cutaneous melanoma. Photochem Photobiol 84:528–536. https://doi.org/10.1111/j.1751-1097.2007.00283.x McGary EC, Lev DC, Bar-Eli M (2002) Cellular adhesion pathways and metastatic potential of human melanoma. Cancer Biol Ther 1:459–465 McGill GG et al (2002) Bcl2 regulation by the melanocyte master regulator Mitf modulates lineage survival and melanoma cell viability. Cell 109:707–718 Merlino G (2005) Cancer biology: the weakest link? Nature 436:33–35. https://doi.org/10.1038/436033a Meulemans D, Bronner-Fraser M (2004) Gene-regulatory interactions in neural crest evolution and development. Dev Cell 7:291–299. https://doi.org/ 10.1016/j.devcel.2004.08.007 Michaloglou C et al (2005) BRAFE600-associated senescence-like cell cycle arrest of human naevi. Nature 436:720–724. https://doi.org/10.1038/nature03890 Micillo R, Panzella L, Koike K, Monfrecola G, Napolitano A, d’Ischia M (2016) “Fifty shades” of black and red or how carboxyl groups fine tune Eumelanin and Pheomelanin properties. Int J Mol Sci 17. https://doi. org/10.3390/ijms17050746 Mitra D et al (2012) An ultraviolet-radiation-independent pathway to melanoma carcinogenesis in the red hair/fair skin background. Nature 491:449–453. https://doi.org/10.1038/nature11624 Miyamura Y et al (2007) Regulation of human skin pigmentation and responses to ultraviolet radiation. Pigment Cell Res 20:2–13. https://doi.org/10.1111/ j.1600-0749.2006.00358.x

M. R. Zaidi et al. Moan J, Porojnicu AC, Dahlback A (2008a) Ultraviolet radiation and malignant melanoma. Adv Exp Med Biol 624:104–116. https://doi.org/ 10.1007/978-0-387-77574-6_9 Moan J, Porojnicu AC, Dahlback A, Setlow RB (2008b) Addressing the health benefits and risks, involving vitamin D or skin cancer, of increased sun exposure. Proc Natl Acad Sci USA 105:668–673. https://doi.org/ 10.1073/pnas.0710615105 Mooi WJ, Peeper DS (2006) Oncogene-induced cell senescence – halting on the road to cancer. N Engl J Med 355:1037–1046. https://doi.org/10.1056/NEJ Mra062285 Moran B, Silva R, Perry AS, Gallagher WM (2018) Epigenetics of malignant melanoma. Semin Cancer Biol 51:80–88. https://doi.org/10.1016/j.semcancer. 2017.10.006 Murali R et al (2013) BAP1 expression in cutaneous melanoma: a pilot study. Pathology 45:606–609. https://doi.org/10.1097/PAT.0b013e3283653818 Natali PG, Nicotra MR, Di Renzo MF, Prat M, Bigotti A, Cavaliere R, Comoglio PM (1993) Expression of the c-Met/HGF receptor in human melanocytic neoplasms: demonstration of the relationship to malignant melanoma tumour progression. Br J Cancer 68: 746–750 Nazarian R et al (2010) Melanomas acquire resistance to B-RAF(V600E) inhibition by RTK or N-RAS upregulation. Nature 468:973–977. https://doi.org/10. 1038/nature09626 Nickoloff BJ, Hendrix MJ, Pollock PM, Trent JM, Miele L, Qin JZ (2005) Notch and NOXA-related pathways in melanoma cells. J Investig Dermatol Symp Proc 10:95–104. https://doi.org/10.1111/j.1087-0024. 2005.200404.x Niihori T et al (2006) Germline KRAS and BRAF mutations in cardio-facio-cutaneous syndrome. Nat Genet 38:294–296. https://doi.org/10.1038/ng1749 Nikolaou V, Stratigos AJ (2014) Emerging trends in the epidemiology of melanoma. Br J Dermatol 170:11–19. https://doi.org/10.1111/bjd.12492 Nishikawa S, Osawa M (2007) Generating quiescent stem cells. Pigment Cell Res 20:263–270. https://doi.org/ 10.1111/j.1600-0749.2007.00388.x Nishimura EK (2011) Melanocyte stem cells: a melanocyte reservoir in hair follicles for hair and skin pigmentation. Pigment Cell Melanoma Res 24:401–410. https://doi. org/10.1111/j.1755-148X.2011.00855.x Nishimura EK et al (2002) Dominant role of the niche in melanocyte stem-cell fate determination. Nature 416:854–860. https://doi.org/10.1038/416854a Nishimura EK, Granter SR, Fisher DE (2005) Mechanisms of hair graying: incomplete melanocyte stem cell maintenance in the niche. Science 307:720–724. https://doi. org/10.1126/science.1099593 Noonan FP et al (2012) Melanoma induction by ultraviolet a but not ultraviolet B radiation requires melanin pigment. Nat Commun 3:884. https://doi.org/10.1038/ ncomms1893

Biology of Melanocytes and Primary Melanoma Nordlund JJ (2006) The pigmentary system: physiology and pathophysiology, 2nd edn. Blackwell Publication, Malden Norris W (1857) Eight cases of melanosis with pathological and therapeutical remarks on that disease. Longman, London Osawa M et al (2005) Molecular characterization of melanocyte stem cells in their niche. Development 132:5589–5599. https://doi.org/10.1242/dev.02161 Palmieri G et al (2015) Multiple molecular pathways in Melanomagenesis: characterization of therapeutic targets. Front Oncol 5:183. https://doi.org/10.3389/ fonc.2015.00183 Park HY, Kosmadaki M, Yaar M, Gilchrest BA (2009) Cellular mechanisms regulating human melanogenesis. Cell Mol Life Sci 66:1493–1506. https://doi.org/ 10.1007/s00018-009-8703-8 Passeron T et al (2007) SOX9 is a key player in ultraviolet B-induced melanocyte differentiation and pigmentation. Proc Natl Acad Sci USA 104:13984–13989. https://doi.org/10.1073/pnas.0705117104 Pavan WJ, Raible DW (2012) Specification of neural crest into sensory neuron and melanocyte lineages. Dev Biol 366:55–63. https://doi.org/ 10.1016/j.ydbio.2012.02.038 Perez-Guijarro E, Day CP, Merlino G, Zaidi MR (2017) Genetically engineered mouse models of melanoma. Cancer 123:2089–2103. https://doi.org/10.1002/ cncr.30684 Petitjean A, Mathe E, Kato S, Ishioka C, Tavtigian SV, Hainaut P, Olivier M (2007) Impact of mutant p53 functional properties on TP53 mutation patterns and tumor phenotype: lessons from recent developments in the IARC TP53 database. Hum Mutat 28:622–629. https://doi.org/10.1002/humu.20495 Pinon P, Wehrle-Haller B (2011) Integrins: versatile receptors controlling melanocyte adhesion, migration and proliferation. Pigment Cell Melanoma Res 24:282–294. https://doi.org/10.1111/j.1755-148X.201 0.00806.x Plummer RS, Shea CR, Nelson M, Powell SK, Freeman DM, Dan CP, Lang D (2008) PAX3 expression in primary melanomas and nevi. Mod Pathol 21:525–530. https://doi.org/10.1038/modpatho l.3801019 Pollock PM et al (2003) High frequency of BRAF mutations in nevi. Nat Genet 33:19–20. https://doi.org/ 10.1038/ng1054 Polsky D, Young AZ, Busam KJ, Alani RM (2001) The transcriptional repressor of p16/Ink4a, Id1, is up-regulated in early melanomas. Cancer Res 61:6008–6011 Poser I, Bosserhoff AK (2004) Transcription factors involved in development and progression of malignant melanoma. Histol Histopathol 19:173–188. https://doi. org/10.14670/HH-19.173 Prota G (1992) Melanins and melanogenesis. Academic, San Diego Puntervoll HE et al (2013) Melanoma prone families with CDK4 germline mutation: phenotypic profile

37 and associations with MC1R variants. J Med Genet 50:264–270. https://doi.org/10.1136/jmedgenet-2012101455 Putz U, Howitt J, Doan A, Goh CP, Low LH, Silke J, Tan SS (2012) The tumor suppressor PTEN is exported in exosomes and has phosphatase activity in recipient cells. Sci Signal 5:ra70. https://doi.org/ 10.1126/scisignal.2003084 Qiu T, Wang H, Wang Y, Zhang Y, Hui Q, Tao K (2015) Identification of genes associated with melanoma metastasis. Kaohsiung J Med Sci 31:553–561. https://doi.org/10.1016/j.kjms.2015.10.002 Raj D, Liu T, Samadashwily G, Li F, Grossman D (2008) Survivin repression by p53, Rb and E2F2 in normal human melanocytes. Carcinogenesis 29:194–201. https://doi.org/10.1093/carcin/bgm219 Rao C et al (2003) A defect in a novel ADAMTS family member is the cause of the belted white-spotting mutation. Development 130:4665–4672. https://doi.org/ 10.1242/dev.00668 Riker AI et al (2008) The gene expression profiles of primary and metastatic melanoma yields a transition point of tumor progression and metastasis. BMC Med Genet 1:13. https://doi.org/10.1186/1755-8794-1-13 Robertson GP (2005) Functional and therapeutic significance of Akt deregulation in malignant melanoma. Cancer Metastasis Rev 24:273–285. https://doi.org/ 10.1007/s10555-005-1577-9 Rodriguez-Viciana P et al (2006) Germline mutations in genes within the MAPK pathway cause cardio-faciocutaneous syndrome. Science 311:1287–1290. https://doi.org/10.1126/science.1124642 Rouzaud F, Kadekaro AL, Abdel-Malek ZA, Hearing VJ (2005) MC1R and the response of melanocytes to ultraviolet radiation. Mutat Res 571:133–152. https://doi.org/10.1016/j.mrfmmm.2004.09.014 Sakai C, Kawakami Y, Law LW, Furumura M, Hearing VJJ (1997) Melanosomal proteins as melanoma-specific immune targets. Melanoma Res 7:83–95 Saldana-Caboverde A, Kos L (2010) Roles of endothelin signaling in melanocyte development and melanoma. Pigment Cell Melanoma Res 23:160–170. https://doi. org/10.1111/j.1755-148X.2010.00678.x Sanchez-Martin M, Rodriguez-Garcia A, Perez-Losada J, Sagrera A, Read AP, Sanchez-Garcia I (2002) SLUG (SNAI2) deletions in patients with Waardenburg disease. Hum Mol Genet 11:3231–3236 Satyamoorthy K, Muyrers J, Meier F, Patel D, Herlyn M (2001) Mel-CAM-specific genetic suppressor elements inhibit melanoma growth and invasion through loss of gap junctional communication. Oncogene 20:4676–4684. https://doi.org/10.1038/sj.onc.1204616 Schatton T et al (2008) Identification of cells initiating human melanomas. Nature 451:345–349. https://doi.org/10.1038/nature06489 Schiaffino MV (2010) Signaling pathways in melanosome biogenesis and pathology. Int J Biochem Cell Biol 42:1094–1104. https://doi.org/10.1016/j. biocel.2010.03.023

38 Scott MC, Suzuki I, Abdel-Malek ZA (2002) Regulation of the human melanocortin 1 receptor expression in epidermal melanocytes by paracrine and endocrine factors and by ultraviolet radiation. Pigment Cell Res 15:433–439 Scott G, Leopardi S, Parker L, Babiarz L, Seiberg M, Han R (2003) The proteinase-activated receptor-2 mediates phagocytosis in a Rho-dependent manner in human keratinocytes. J Invest Dermatol 121:529–541. https://doi.org/10.1046/j.1523-1747.2003.12427.x Shain AH et al (2015) The genetic evolution of melanoma from precursor lesions. N Engl J Med 373:1926–1936. https://doi.org/10.1056/NEJMoa1502583 Sharpless NE, Kannan K, Xu J, Bosenberg MW, Chin L (2003) Both products of the mouse Ink4a/Arf locus suppress melanoma formation in vivo. Oncogene 22:5055–5059. https://doi.org/10.1038/sj.onc.1206809 Shay JW (2016) Role of telomeres and telomerase in aging and Cancer. Cancer Discov 6:584–593. https://doi.org/ 10.1158/2159-8290.CD-16-0062 Shriver MD et al (2003) Skin pigmentation, biogeographical ancestry and admixture mapping. Hum Genet 112:387–399. https://doi.org/10.1007/s00439002-0896-y Silver DL, Hou L, Pavan WJ (2006) The genetic regulation of pigment cell development. Adv Exp Med Biol 589:155–169. https://doi.org/10.1007/9780-387-46954-6_9 Silver DL, Hou L, Somerville R, Young ME, Apte SS, Pavan WJ (2008) The secreted metalloprotease ADAMTS20 is required for melanoblast survival. PLoS Genet 4:e1000003. https://doi.org/10.1371/journal.pgen.1000003 Simanshu DK, Nissley DV, McCormick F (2017) RAS proteins and their regulators in human disease. Cell 170:17–33. https://doi.org/10.1016/j.cell.2017.06.009 Slominski A, Pawelek J (1998) Animals under the sun: effects of ultraviolet radiation on mammalian skin. Clin Dermatol 16:503–515 Slominski A, Wortsman J, Plonka PM, Schallreuter KU, Paus R, Tobin DJ (2005) Hair follicle pigmentation. J Invest Dermatol 124:13–21. https://doi.org/10.1111/ j.0022-202X.2004.23528.x Slominski AT et al (2012) Novel vitamin D hydroxyderivatives inhibit melanoma growth and show differential effects on normal melanocytes. Anticancer Res 32:3733–3742 Slominski AT et al (2017) Vitamin D signaling and melanoma: role of vitamin D and its receptors in melanoma progression and management. Lab Investig 97: 706–724. https://doi.org/10.1038/labinvest.2017.3 Smith MP et al (2013) Effect of SMURF2 targeting on susceptibility to MEK inhibitors in melanoma. J Natl Cancer Inst 105:33–46. https://doi.org/10.1093/ jnci/djs471 Smith MP et al (2016) Inhibiting drivers of non-mutational drug tolerance is a salvage strategy for targeted melanoma therapy. Cancer Cell 29:270–284. https://doi.org/ 10.1016/j.ccell.2016.02.003

M. R. Zaidi et al. Soengas MS et al (2001) Inactivation of the apoptosis effector Apaf-1 in malignant melanoma. Nature 409:207–211. https://doi.org/10.1038/35051606 Soura E, Eliades PJ, Shannon K, Stratigos AJ, Tsao H (2016) Hereditary melanoma: update on syndromes and management: genetics of familial atypical multiple mole melanoma syndrome. J Am Acad Dermatol 74:395–407; quiz 408-310. https://doi.org/10.1016/j. jaad.2015.08.038 Stark M, Hayward N (2007) Genome-wide loss of heterozygosity and copy number analysis in melanoma using high-density single-nucleotide polymorphism arrays. Cancer Res 67:2632–2642. https://doi.org/10.1158/ 0008-5472.CAN-06-4152 Sturm RA, Larsson M (2009) Genetics of human iris colour and patterns. Pigment Cell Melanoma Res 22:544–562. https://doi.org/10.1111/j.1755-148X. 2009.00606.x Tadokoro T et al (2003) UV-induced DNA damage and melanin content in human skin differing in racial/ ethnic origin. FASEB J 17:1177–1179. https://doi.org/ 10.1096/fj.02-0865fje Tadokoro T et al (2005) Mechanisms of skin tanning in different racial/ethnic groups in response to ultraviolet radiation. J Invest Dermatol 124:1326–1332. https://doi.org/10.1111/j.0022-202X.2005.23760.x Takeda K, Yokoyama S, Yasumoto K, Saito H, Udono T, Takahashi K, Shibahara S (2003) OTX2 regulates expression of DOPAchrome tautomerase in human retinal pigment epithelium. Biochem Biophys Res Commun 300:908–914 Talantov D et al (2005) Novel genes associated with malignant melanoma but not benign melanocytic lesions. Clin Cancer Res 11:7234–7242. https://doi.org/ 10.1158/1078-0432.CCR-05-0683 Tang JY et al (2011) Calcium plus vitamin D supplementation and the risk of nonmelanoma and melanoma skin cancer: post hoc analyses of the women’s health initiative randomized controlled trial. J Clin Oncol 29:3078–3084. https://doi.org/10.1200/JCO. 2011.34.5967 Testa JR et al (2011) Germline BAP1 mutations predispose to malignant mesothelioma. Nat Genet 43:1022–1025. https://doi.org/10.1038/ng.912 Thomas AJ, Erickson CA (2008) The making of a melanocyte: the specification of melanoblasts from the neural crest. Pigment Cell Melanoma Res 21:598–610. https://doi.org/10.1111/j.1755-148X.2008.00506.x Thong HY, Jee SH, Sun CC, Boissy RE (2003) The patterns of melanosome distribution in keratinocytes of human skin as one determining factor of skin colour. Br J Dermatol 149:498–505 Timerman D et al (2017) Vitamin D deficiency is associated with a worse prognosis in metastatic melanoma. Oncotarget 8:6873–6882. https://doi.org/10.18632/ oncotarget.14316 Tobin DJ (2008) Human hair pigmentation – biological aspects. Int J Cosmet Sci 30:233–257. https://doi.org/ 10.1111/j.1468-2494.2008.00456.x

Biology of Melanocytes and Primary Melanoma Tormo D et al (2006) Rapid growth of invasive metastatic melanoma in carcinogen-treated hepatocyte growth factor/scatter factor-transgenic mice carrying an oncogenic CDK4 mutation. Am J Pathol 169:665–672. https://doi.org/10.2353/ajpath.2006.060017 Toyofuku K et al (2002) The etiology of oculocutaneous albinism (OCA) type II: the pink protein modulates the processing and transport of tyrosinase. Pigment Cell Res 15:217–224 Tran S, Rizos H (2013) Human nevi lack distinguishing senescence traits. Aging (Albany NY) 5:98–99. https://doi.org/10.18632/aging.100537 Tran TT, Schulman J, Fisher DE (2008) UV and pigmentation: molecular mechanisms and social controversies. Pigment Cell Melanoma Res 21:509–516. https://doi. org/10.1111/j.1755-148X.2008.00498.x Tsao H, Chin L, Garraway LA, Fisher DE (2012) Melanoma: from mutations to medicine. Genes Dev 26:1131–1155. https://doi.org/10.1101/gad.191 999.112 Tsutsumida A et al (2004) Epigenetic silencing of E- and Pcadherin gene expression in human melanoma cell lines. Int J Oncol 25:1415–1421 Vance KW, Carreira S, Brosch G, Goding CR (2005) Tbx2 is overexpressed and plays an important role in maintaining proliferation and suppression of senescence in melanomas. Cancer Res 65:2260–2268. https://doi.org/10.1158/0008-5472.CAN-04-3045 Villanueva J, Vultur A, Herlyn M (2011) Resistance to BRAF inhibitors: unraveling mechanisms and future treatment options. Cancer Res 71:7137–7140. https://doi.org/10.1158/0008-5472.CAN-11-1243 Viros A et al (2014) Ultraviolet radiation accelerates BRAF-driven melanomagenesis by targeting TP53. Nature 511:478–482. https://doi.org/10.1038/ nature13298 Wadt K et al (2012) A cryptic BAP1 splice mutation in a family with uveal and cutaneous melanoma, and paraganglioma. Pigment Cell Melanoma Res 25:815–818. https://doi.org/10.1111/pcmr.12006 Wan PT et al (2004) Mechanism of activation of the RAFERK signaling pathway by oncogenic mutations of BRAF. Cell 116:855–867 Wehrle-Haller B (2003) The role of Kit-ligand in melanocyte development and epidermal homeostasis. Pigment Cell Res 16:287–296 Weinstein IB, Joe A (2008) Oncogene addiction. Cancer Res 68:3077–3080; discussion 3080. https://doi.org/ 10.1158/0008-5472.CAN-07-3293 Whiteman DC, Parsons PG, Green AC (1999) Determinants of melanocyte density in adult human skin. Arch Dermatol Res 291:511–516 Whiteman DC, Whiteman CA, Green AC (2001) Childhood sun exposure as a risk factor for melanoma: a systematic review of epidemiologic studies. Cancer Causes Control 12:69–82 Wiesner T et al (2011) Germline mutations in BAP1 predispose to melanocytic tumors. Nat Genet 43: 1018–1021. https://doi.org/10.1038/ng.910

39 Wiesner T et al (2012) Toward an improved definition of the tumor spectrum associated with BAP1 germline mutations. J Clin Oncol 30:e337–e340. https://doi. org/10.1200/JCO.2011.41.2965 Wilkie AL, Jordan SA, Jackson IJ (2002) Neural crest progenitors of the melanocyte lineage: coat colour patterns revisited. Development 129:3349–3357 Willmore-Payne C, Holden JA, Tripp S, Layfield LJ (2005) Human malignant melanoma: detection of BRAF- and c-kit-activating mutations by high-resolution amplicon melting analysis. Hum Pathol 36:486–493. https://doi.org/10.1016/j.humpath.2005. 03.015 Wittgen HG, van Kempen LC (2007) Reactive oxygen species in melanoma and its therapeutic implications. Melanoma Res 17:400–409. https://doi.org/10.1097/ CMR.0b013e3282f1d312 Wolf Horrell EM, Boulanger MC, D’Orazio JA (2016) Melanocortin 1 receptor: structure, function, and regulation. Front Genet 7:95. https://doi.org/10.3389/ fgene.2016.00095 Wu X, Hammer JA (2014) Melanosome transfer: it is best to give and receive. Curr Opin Cell Biol 29:1–7. https:// doi.org/10.1016/j.ceb.2014.02.003 Wu X, Hammer JA 3rd (2000) Making sense of melanosome dynamics in mouse melanocytes. Pigment Cell Res 13:241–247 Xu L et al (2008) Gene expression changes in an animal melanoma model correlate with aggressiveness of human melanoma metastases. Mol Cancer Res 6:760–769. https://doi.org/10.1158/1541-7786.MCR07-0344 Yamaguchi Y, Hearing VJ (2006) Melanocyte distribution and function in human skin. In: Hearing VJ, Leong SPL (eds) From melanocytes to melanoma. Humana Press, Totowa Yamaguchi Y et al (2004) Mesenchymal-epithelial interactions in the skin: increased expression of dickkopf1 by palmoplantar fibroblasts inhibits melanocyte growth and differentiation. J Cell Biol 165:275–285. https://doi.org/10.1083/jcb.200311122 Yamaguchi Y, Brenner M, Hearing VJ (2007) The regulation of skin pigmentation. J Biol Chem 282:27557–27561. https://doi.org/10.1074/jbc.R7000 26200 Yamaguchi Y et al (2008) Dickkopf 1 (DKK1) regulates skin pigmentation and thickness by affecting Wnt/betacatenin signaling in keratinocytes. FASEB J 22:1009–1020. https://doi.org/10.1096/fj.07-9475com Yang G, Zhang G, Pittelkow MR, Ramoni M, Tsao H (2006) Expression profiling of UVB response in melanocytes identifies a set of p53-target genes. J Invest Dermatol 126:2490–2506. https://doi.org/10.1038/sj. jid.5700470 Yasumoto K, Takeda K, Saito H, Watanabe K, Takahashi K, Shibahara S (2002) Microphthalmia-associated transcription factor interacts with LEF-1, a mediator of Wnt signaling. EMBO J 21:2703–2714. https://doi. org/10.1093/emboj/21.11.2703

40 Yokoyama S et al (2008) Pharmacologic suppression of MITF expression via HDAC inhibitors in the melanocyte lineage. Pigment Cell Melanoma Res 21:457–463. https://doi.org/10.1111/j.1755-148X.2008.00480.x Yokoyama S et al (2011) A novel recurrent mutation in MITF predisposes to familial and sporadic melanoma. Nature 480:99–103. https://doi.org/10.1038/nature10630 Yoshida Y et al (2007) Functional analysis of keratinocytes in skin color using a human skin substitute model composed of cells derived from different skin pigmentation types. FASEB J 21:2829–2839. https://doi.org/ 10.1096/fj.06-6845com Young AR (2006) Acute effects of UVR on human eyes and skin. Prog Biophys Mol Biol 92:80–85. https://doi.org/10.1016/j.pbiomolbio.2006.02.005 Zaidi MR, Day CP, Merlino G (2008) From UVs to metastases: modeling melanoma initiation and progression

M. R. Zaidi et al. in the mouse. J Invest Dermatol 128:2381–2391. https://doi.org/10.1038/jid.2008.177 Zaidi MR et al (2011) Interferon-gamma links ultraviolet radiation to melanomagenesis in mice. Nature 469: 548–553. https://doi.org/10.1038/nature09666 Zaidi MR, De Fabo EC, Noonan FP, Merlino G (2012) Shedding light on melanocyte pathobiology in vivo. Cancer Res 72:1591–1595. https://doi.org/10.1158/ 0008-5472.CAN-11-2586 Zhan T, Rindtorff N, Boutros M (2017) Wnt signaling in cancer. Oncogene 36:1461–1473. https://doi.org/ 10.1038/onc.2016.304 Zhang Y, Xiong Y, Yarbrough WG (1998) ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell 92: 725–734

Immunology of Melanoma Paul F. Robbins and Yong-Chen Lu

Contents Innate Immunity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Adaptive Immunity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 Immune Regulation and Tolerance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Co-stimulatory/Co-inhibitory Receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Regulatory T Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cytokines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

44 44 47 48

Animal Tumor Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Antigens Recognized by Tumor-Reactive T Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 Identification of Tumor Antigens Recognized by T Cells: General Principles . . . . . . . . . . 50 Melanoma Neoantigens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 Conclusions: Implications for Immunotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Abstract

Historically, the basis of cancer immunology comes from observations of therapy for cancer dating back to antiquity indicating that the immune system can recognize and reject cancers. Ancient Eastern and Western medical writings described the treatment of “tumors” by injecting purulent materials from infected wounds, leading to infection of the tumor mass and surrounding tissues and sometimes leading

P. F. Robbins (*) · Y.-C. Lu Surgery Branch, National Cancer Institute, National Institutes of Health, Bethesda, MD, USA e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_44

to tumor necrosis and regression. This phenomenon, linking cancer therapy with inflammation, has evolved into the modern-day disciplines of cancer immunology and immunotherapy. Melanoma is among the most highly immunogenic human tumors and has become the prototype for identifying tumor antigens and developing antigen-specific immunotherapies, including checkpoint inhibitors, cancer vaccines, and adoptive T cell transfer (see chapters ▶ “Cytokines (IL-2, IFN, GM-CSF, etc.) Melanoma,” ▶ “Checkpoint Inhibitors in the Treatment of Metastatic Melanoma,” ▶ “Melanoma Vaccines,” and ▶ “Cellular Therapy for Melanoma”). Many of the

41

42

P. F. Robbins and Y.-C. Lu

principles of cancer immunology learned from studies of melanoma are now being extended to other types of cancer. Future research aims to develop new strategies to improve the efficacy of current melanoma immunotherapies, such as combination therapies. Keywords

Innate immunity · Adaptive immunity · Immune regulation and tolerance · Animal tumor models · Antigens recognized by tumorreactive T cells · Implications for immunotherapy

Innate Immunity The immune system has both innate and adaptive components. Innate immunity arises during development of the immune system, and imprinted components of the innate immune system are poised to recognize “foreign invaders” without any further adaptation. Components of the innate immune system include soluble factors (e.g., complement, cytokines), inflammatory cells (granulocytes, macrophages), and natural killer (NK) cells. These components become biologically active without prior antigen exposure and underlie the initial responses to microorganisms as well as tumor cells. In 1893, William B. Coley, a surgeon in New York, developed a treatment for cancer patients using live cultures of bacteria or bacterial filtrates (so-called Coley’s toxins) injected around the tumor. This treatment was successful in select patients with advanced cancers, which was quite remarkable in an age that had no formal notion of clinical research or clinical trials. Despite some successes in which large tumors regressed or disappeared in individual patients, this cancer therapy was controversial and was eventually replaced by radiation therapy and chemotherapy. However, the legacy of William B. Coley and tumor hemorrhagic necrosis led to the discovery of the tumor necrosis factor (TNF), a soluble component of innate immunity which can induce profound inflammation and necrosis of tumors (Carswell et al. 1975). As a cytokine induced by bacterial infection, TNF is produced by activated

macrophages and other leukocytes. In mice, systemic administration of TNF causes tumor hemorrhagic necrosis similar to the phenomenon associated with bacterial infection. Unfortunately, the fact that humans are much more sensitive than mice to the adverse effects of TNF has limited its use in patients to clinical trials involving surgical administration of TNF in conjunction with alkylating chemotherapy interventional administration by closed-circuit regional perfusion for treatment of melanoma and other tumors localized to the liver or extremities. One class of cells that may be involved with innate tumor immunity is the NK cell subset (Lanier 2005). NK cells are particularly interesting because they recognize the “absence of self,” or the lack of expression of specific major histocompatibility complex (MHC) molecules. MHC molecules are expressed by almost all somatic cells and are required for presenting peptide antigens to T cells, eliciting cellular immunity. However, a subset of tumor cells, which are deficient in the expression of MHC molecules, can escape T cell-mediated antitumor responses. NK cells provide a redundant system to kill target cells which have lost expression of specific MHC alleles. NK cells are normally inhibited by the binding of their specific receptors to MHC molecules, but when MHC molecules are missing on cancer cells or virus-infected cells, the lack of inhibitory signals allows NK cell activation. NK cells also have activation receptors that may play important roles in tumor immunity (Diefenbach et al. 2001). The relative importance of NK cells in cancer immunity in humans is uncertain, although activated NK cells may partially mediate the antitumor effects of therapy with IL-2. Another class of cells that mediate innate antitumor immunity is the category of antigen-presenting cells that include macrophages and dendritic cells (DCs) (Trombetta and Mellman 2005). These cells present tumor antigens to T cells and trigger subsequent adaptive immune responses. Another important feature of these cells is sensing the danger signals from pathogens or tumor cells through the recognition of unique molecular patterns (Kroemer et al. 2013). For example, following chemotherapy or radiotherapy, dying tumor cells secret HMGB1 (high

Immunology of Melanoma

mobility group box 1), which can be recognized by TLR4 (toll-like receptor 4) expressed by DCs. This signaling induces DCs to present tumor antigens to T cells and trigger antitumor responses (Apetoh et al. 2007). Further understanding of the innate immune activation within the tumor microenvironment may provide new strategies and therapeutic targets for cancer immunotherapies (Woo et al. 2015).

Adaptive Immunity The hallmarks of the adaptive immune system, which includes T cells and B cells, are diversity, specificity, and memory. The adaptive arm of the immune system forms the basis for melanoma immunotherapies involving checkpoint inhibitors, vaccines, and adoptive T cell transfer. As the immune system responds to antigens carried by infectious agents or cancer, B cells and T cells that recognize individual antigens proliferate, become activated, and produce high affinity receptors (IgGs or T cell receptors, respectively). There is enormous molecular diversity in these receptors, such that B cells and T cells can theoretically recognize 1018–1022 discrete targets. In addition, the adaptive immune system is exquisitely specific, with the capacity to recognize minute chemical modifications on antigens, such as phosphate or methyl groups (Zarling et al. 2006; Deng et al. 2007). A small proportion of antigen-specific B cells or T cells enter the memory pool following antigen stimulation, where they can reside for long periods of time (months to years) awaiting possible antigenic rechallenge. Over the past two decades, compelling evidence has shown that the adaptive immune system can specifically react against antigens expressed by cancer cells. Many studies have revealed the molecular characteristics of melanoma antigens, through advances in the fields of biochemistry, allowing protein and carbohydrate purification, and recombinant DNA technology, leading to the cloning of genes encoding melanoma antigens. Proteins, sugars, and lipids can form surface structures on melanoma cells that are recognized serologically. In addition, intracellular antigens

43

derived from transmembrane, cytosolic, or nuclear proteins can be processed into short peptides and loaded onto MHC molecules for transit and display on the cell surface, where they can be recognized by T lymphocytes. Cytosolic and nuclear proteins are typically processed in the cytoplasm and endoplasmic reticulum, and derivative peptides generally 8–11 amino acids in length are complexed to MHC class I molecules for recognition by CD8+ cytolytic T cells. In contrast, transmembrane or extracellular proteins are typically processed in a specialized endosomal compartment, and derivative peptides with an average length of 15–18 amino acids are complexed to MHC class II molecules for recognition by CD4+ T helper cells or inhibitory cells termed regulatory T cells (Tregs). Thus, tumor antigens from diverse subcellular locations can potentially be presented for recognition by different subclasses of interacting T cells, generating potent antitumor immunity. CD4+ T cells are particularly important for providing help to cytotoxic CD8+ T cells and to antibody-producing B lymphocytes and may play a central role in anti-melanoma immunity (Pardoll and Topalian 1998). CD4+ T cells can also help to orchestrate inflammatory responses involving macrophages, eosinophils, and other immune cells that can destroy tumors. Although many types of CD4+ helper T cells have been identified, there are three predominant cell types that have been investigated for their role in tumor immunity: Th1, Th2, and Th17. Th1 cells provide help for generating CD8+ T cell responses, secrete IFNγ and other cytokines, and in some animal models are crucial components of antitumor immunity, particularly models in which cytotoxic CD8+ T cells are required for tumor rejection. Th2 cells secrete IL-4 and IL-13 and help to elicit certain types of antibody responses and are particularly important for fighting parasitic infections. Although many similarities exist between parasites and tumor cells, the role of Th2 responses in antitumor immunity remains to be completely defined. It is likely that under some circumstances Th2 cells can exert antitumor effects by eliciting inflammatory responses. For instance, tumor cells expressing IL-4 elicit inflammatory responses that are sufficient to mediate tumor rejection,

44

suggesting that Th2 cells can play a role in tumor immunity (Golumbek et al. 1991). Th17 cells secrete IL-17 and IL-6, and their differentiation is promoted by IL-6 and TGF-β, two cytokines often found in abundance in the tumor microenvironment (Weaver et al. 2006; Korn et al. 2009). In animal models, Th17 cells have been shown to play a role in antimicrobial gut immunity as well as autoimmunity. Results of a murine tumor model system indicate that Th17 cells may be more effective at mediating tumor rejection than Th1 cells in some circumstances (Muranski et al. 2008), although the role of Th17 cells in mediating human antitumor responses remains unclear (Asadzadeh et al. 2017).

Immune Regulation and Tolerance Conventional CD8+ and CD4+ T cells play dominant roles in mediating immune responses to foreign antigens, regulating responses to self-antigens, and mediating antitumor responses. Evaluation of approaches for modulating T cell-mediated immune responses has become a major research focus for cancer immunotherapies.

Co-stimulatory/Co-inhibitory Receptors Although the antigen-specific TCR controls the specificity of a T cell reaction, an additional receptor-ligand interaction, i.e., an antigen-independent co-signal, is essential for the optimal activation/ inhibition of T cells. These co-signaling molecules can be classified as either co-stimulators or co-inhibitors depending on whether they enhance/ sustain or quench/terminate TCR-mediated responses, respectively. Co-signaling is dependent on the TCR signal, and in the absence of TCR signaling, T cells fail to generate an effective response (Chen 1998; Schwartz 2003; Alcover et al. 2018). A classic example of co-stimulation is the interaction between the CD28 co-receptor expressed on naïve T cells and its ligands CD80/ CD86 (B7.1/B7.2), expressed on antigen-presenting cells such as dendritic cells (DCs). When

P. F. Robbins and Y.-C. Lu

T cells receive an appropriate signal as a result of TCR cross-linking, this co-stimulatory interaction triggers proliferation and IL-2 secretion and prevents the death of activated T cells (Lenschow et al. 1996; Sharpe and Freeman 2002). Following this initial signaling event, T cells differentiate into effector cells, and a subset will ultimately become long-lived memory T cells, with each step positively or negatively controlled by other co-signaling molecules. Co-signaling molecules were originally described as either receptors or ligands and as either stimulatory or inhibitory; however, studies have shown that such designations are overly restrictive. For example, CD80 and CD86 can act as co-stimulatory ligands if they bind to the CD28 receptor on naïve T cells, but they function as co-inhibitory ligands when they bind to the CTLA-4 receptor on activated T cells. Furthermore, a ligand can also act as a receptor, termed “reverse signaling,” if its engagement leads to the intracellular transmission of a signal. For example, binding of CD80/CD86 on DC by CTLA-4 on T cells induces indoleamine 2,3-dioxygenase (IDO) expression in DC, which subsequently inhibits their activation (Grohmann et al. 2002). In this context, CD80 and CD86 have also been termed “counter-receptors.” There is an expanding array of co-signaling molecules, including the B7-CD28 superfamily and the tumor necrosis factor receptor (TNFR) superfamily. Although discussed separately below, it is also likely that cross talk occurs among these families (Croft 2005). While the exact dynamics of these co-signaling events are still being elucidated, it is already clear that each can potentially serve as a therapeutic target to modulate anticancer immune response.

B7-CD28 Family The ligation of the B7 family members CD80 and CD86 (B7-1/B7-2) with CD28 or CTLA-4 is the best-described co-signaling event for activating or inhibiting T cells, respectively. Knowledge of the B7-CD28 family has been expanding as new pathways have been discovered in rapid succession, including the PD-1 and PD-L1/PD-L2 axis and other B7 family members.

Immunology of Melanoma

CD28/CTLA-4-CD80/CD86 axis: CD28 is constitutively expressed on naïve T cells, and CD80/CD86 engagement by CD28 transmits activation and survival signals to T cells. Cell surface expression of CTLA-4, which binds to CD80 and CD86 with a greater than tenfold higher affinity than CD28, is upregulated following T cell activation, curtailing T cell responses. This inhibitory pathway likely exists to prevent over-activation or aberrant host immune responses and is critical in the maintenance of self-tolerance. CTLA-4 appears to be a master checkpoint for T cell suppression because CTLA-4-deficient mice show massive infiltration of self-reactive T cells in peripheral organs (Tivol et al. 1995; Waterhouse et al. 1995). In addition to inducing T cell suppression, CTLA-4 may aid in the activation of regulatory T cells (Read et al. 2000; Salomon et al. 2000; Takahashi et al. 2000) and may negatively reverse signal through CD80/CD86 expressed on DC, further inducing tolerance (Fallarino et al. 2003). Conversely, the blockade of CTLA-4 can reactivate T cells. James Allison’s group was the first to demonstrate that in vivo administration of CTLA-4 blockade antibody can induce the rejection of pre-established tumors in mice, suggesting that the blockade of CTLA-4 can enable effective T cell-mediated immune responses against tumors (Leach et al. 1996). Administration of an antiCTLA-4 monoclonal antibody to cancer patients has been shown in several studies to induce autoimmune toxicities in multiple organs while leading to clinical responses up to 17% of treated melanoma patients (Phan et al. 2003; Robinson et al. 2004; Blansfield et al. 2005). PD-1-PD-L1/PD-L2 axis: Discovered by Tasuku Honjo’s group in 1992, PD-1 (programmed death 1) is a novel member of the immunoglobulin gene superfamily, which is induced upon programmed cell death (Ishida et al. 1992). Later studies demonstrate that PD-1 is a co-inhibitor found on activated T cells, B cells, and thymocytes, and it is thought to play a role in maintaining peripheral self-tolerance (Agata et al. 1996; Nishimura et al. 1999; Blank et al. 2003; Keir et al. 2008; Nguyen and Ohashi 2015). PD-1 antagonists accelerate autoimmune diseases

45

and graft-versus-host disease in animal models, supporting this mechanism (Blazar et al. 2003), and PD-1 knockout mice develop autoimmune diseases, although much later and with lesser severity than CTLA-4 knockout mice (Salama et al. 2003; Tsushima et al. 2003). The ligands for PD-1, PD-L1, and PD-L2 (also known as B7-H1 and B7-DC, respectively) possess differential expression patterns: PD-L1 is detected on a proportion of tissue macrophages as well as DC, and its expression is inducible in most peripheral tissues by inflammatory cytokines such as interferons, while PD-L2 expression is limited to macrophages and DC (Dong et al. 1999, 2002; Ishida et al. 2002; Yamazaki et al. 2002). This suggests that PD-L1 may function to downregulate effector T cell functions in inflamed peripheral tissues, while PD-L2 may modulate T cell responses in lymphoid organs. Importantly, PD-L1 is constitutively expressed on many human tumors, including melanomas, and by ligating PD-1 it can suppress antitumor immunity. Conversely, the blockade of PD-1/PD-L1 pathway can reactivate T cells and induce antitumor immune responses (Baumeister et al. 2016). Treatment of patients with metastatic melanoma with immune checkpoint blockade using monoclonal antibodies directed against PD-1 (nivolumab, pembrolizumab, etc.) is associated with significant clinical responses and objective response rates of between 25% and 38% (Brahmer et al. 2010; Topalian et al. 2012, 2014; Hamid et al. 2013; Weber et al. 2013; Robert et al. 2014), while lower but still significant clinical response rates of between 17% and 30% were seen in response to anti-PD-L1 treatment (Brahmer et al. 2012; Herbst et al. 2014). Early murine studies, as well as recent human studies, provide evidence for the importance of neoantigen-reactive T cells in effective cancer immunotherapy (Foley 1953; Tran et al. 2017). These observations are consistent with the hypothesis that immune checkpoint blockade can lead to reactivation of potent tumor-specific, neoantigen-reactive T cells. The PD-1 immunotherapies and the potential mechanisms of action will be discussed in detail in chapter

46

▶ “Checkpoint Inhibitors in the Treatment of Metastatic Melanoma.” Other B7 family members: Shortly after B7-H1 (PD-L1) was reported, other members of this family were reported, including B7-H2, B7-H3, B7-H4, B7-H5, and B7-H6 (Ceeraz et al. 2013). B7-H2 expression on DC may have a role in co-stimulating Th2-type responses and B-cell function by binding its receptor, inducible co-stimulator (ICOS) (Ling et al. 2000; Wang et al. 2000; Yoshinaga et al. 2000). B7-H3 is broadly expressed in many types of tissues, including lymphoid tissues, suggesting that it may play both a stimulatory and inhibitory role (Flies and Chen 2007). B7-H4 mRNA is widely expressed by many tissues, but the B7-H4 protein expression is relatively limited (Choi et al. 2003; MacGregor and Ohashi 2017). More importantly, B7-H4 is highly expressed in a variety of tumors, including melanoma (Quandt et al. 2011). Similar to other family members, studies suggest that B7-H4 participates in the suppression of T cell immunity (Prasad et al. 2003; Sica et al. 2003). Conversely, a recent study suggests that B7-H4 expressed in nonhematopoietic cells can promote antitumor immunity (Rahbar et al. 2015). B7-H5, also known as VSIR or VISTA, is highly expressed on monocytes, neutrophils, and dendritic cells. Ligation of B7-H5 inhibits T cell proliferation and activation (Flies et al. 2014; Lines et al. 2014). B7-H6 is expressed on tumor cells, including melanoma, and can trigger the activation of NK cells through the interaction with its ligand NKp30 (Brandt et al. 2009). Similar to PD-1/PDL1, the blockade of these B7 family members may induce tumor regressions. As a result, investigators are actively studying the clinical activities of the antibodies specific to these B7 family members in recent clinical trials.

TNFR Family Following the initial co-stimulatory event between CD28 and CD80/CD86, TNFR family members including CD27, CD30, herpesvirus entry mediator (HVEM), OX40 (CD134), 4-1BB (CD137), and glucocorticoid-induced TNFR family-related gene (GITR), as well as their ligands CD70, CD30L, LIGHT, OX40L, 4-1BBL, and

P. F. Robbins and Y.-C. Lu

GITRL, respectively, are key mediators of sustaining T cell responses (Watts 2005). HVEM is constitutively expressed by naïve T cells, and its ligand, LIGHT, is constitutively expressed by immature DC (Morel et al. 2000; Tamada et al. 2000a). Blockade of this interaction results in delayed T cell activation and clonal expansion and notably lessens acute graft-versus-host disease (Tamada et al. 2000b). CD27 and GITR are also constitutively expressed, but at lower levels, and their expression peaks after T cell activation (Borst et al. 1989; Shimizu et al. 2002). Notably, high levels of GITR are constitutively expressed on Treg (see below) and are further upregulated following activation (McHugh et al. 2002). CD70, the ligand for CD27, is not constitutively expressed but is induced early after antigen encounter (Lens et al. 1996). GITRL is transiently upregulated and then downmodulated on DC after T cell activation (Tone et al. 2003; Stephens et al. 2004). Consistent with this pattern of expression, CD27-CD70 interactions influence early T cell expansion after antigen exposure (Hintzen et al. 1995). This interaction also strongly impacts the formation of T memory cells. CD27-deficient mice show a significant decrease in the number of antigen-specific T memory cells (Hendriks et al. 2000), while mice immunized with CD70transfected tumor cells showed an increase in the number of tumor-specific effector T cells and partial immunologic protection when rechallenged with wild-type tumors (Lorenz et al. 1999). The pattern of GITR and GITRL expression also suggests that this interaction plays a role following initial T cell activation. One proposed model is that during T cell activation, the GITR/GITRL interaction helps to stimulate IL-2 production and T cell differentiation into effector cells. During this time, local Treg also expressing GITR interact with GITRL on DC in an environment with abundant IL-2, resulting in a parallel expansion in the suppressor population. When the transient expression of GITRL by DC passes, the effector T cells are vulnerable to suppression by the Treg, thereby turning off the response (Shevach and Stephens 2006). The other TNFR family members, OX40, 4-1BB, CD30, and their ligands are induced later

Immunology of Melanoma

in the immune response, peaking between 1 and 4 days following activation (Ellis et al. 1993; Pollok et al. 1995; Gramaglia et al. 1998). Cells lacking OX40 are capable of initial activation, including IL-2 production and clonal expansion, but subsequently lose Bcl-xL and Bcl-2 expression and undergo apoptosis (Rogers et al. 2001). Similarly, mice deficient in OX40 show a decreased number of antigen-specific T cells in the primary response and a decreased number of memory T cells (Gramaglia et al. 2000). 4-1BB impacts on both CD8+ and CD4+ T cell survival and cytokine production, and also enhances CD8 cytotoxic capacity (Myers and Vella 2005). CD30 appears to have a similar function in that it influences clonal T cell expansion following activation as well as the memory response. CD30L-deficient mice show a limited expansion of the CD8 population following activation, as well as limited secondary expansion following boosting (Podack et al. 2002). CD30 also appears to play a role in humoral immunity by promoting B-cell proliferation and antibody production (Shanebeck et al. 1995). Taken together, these results suggest that OX40, 4-1BB, and CD30 promote effector T cell activation and expansion in both primary and secondary immune responses. Several investigators have attempted to capitalize on this understanding to expand tumor antigenspecific T cells, by transfecting tumors with TNFR ligands or utilizing TNFR agonists (Weinberg et al. 2000). The key to therapeutic success will no doubt rest in identifying and targeting the TNFR co-stimulatory molecules on the surface of the tumor antigen-specific T cells such that this population is expanded and activated, boosting the antitumor immune response.

TIM Family The T cell immunoglobulin mucin (TIM) proteins are a recently described family of co-signaling molecules. The gene family consists of eight members in mice, TIM-1 through TIM-8, and three in humans, TIM-1, TIM-3, and TIM-4 (Kuchroo et al. 2003). The three human members are differentially expressed on immune cells and are emerging as key regulators of Th1- and Th2mediated immune responses. TIM-1 is not expressed on naïve T cells but on activated

47

T cells and is expressed to a higher degree on Th2 cells compared to Th1 cells (Umetsu et al. 2005). TIM-4 is the ligand for TIM-1 and is expressed on macrophages and DC, and its engagement provides a co-stimulatory signal resulting in T cell proliferation and Th2 cytokine production (Meyers et al. 2005). Similarly, TIM-3 is not expressed on naïve T cells, but is expressed on the surface of fully differentiated Th1 cells. Its ligand, galectin-9, is found on a number of hematopoietic cells including activated T lymphocytes and DC (Monney et al. 2002; Dai et al. 2005). Engagement of TIM-3 with its ligand galectin-9 engagement can lead to the death of TIM3+ Th1 cell (Zhu et al. 2005). This co-signaling pathway therefore negatively regulate effector T cells, curtailing clonal expansion, inducing peripheral tolerance, and potentially preventing protracted inflammation by turning off the Th1 response. Lastly, tumor-associated DCs express a high level of TIM-3, which can suppress innate DC immune responses via an HMGB1-dependent, galectin-9-independent mechanism (Chiba et al. 2012). The TIM-3-galectin-9 pathway is of particular interest in studying the immune response to tumors, since tumor cells have been shown to express members of the galectin family, including galectin-9 (Lahm et al. 2001). Tumors may suppress immune surveillance by tumor-infiltrating lymphocytes through this mechanism, resulting in the inhibition of tumor-specific effector T cells. In addition, TIM-3 blocking antibody can promote T cell-mediated, IFN-γ-dependent antitumor responses in mouse tumor models (Ngiow et al. 2011). As a result, using agents to block TIM-3-galectin-9 interaction holds promise as a way to boost anticancer immune responses (Anderson 2014), and the results of ongoing clinical trials of TIM-3 blockade therapies should help to elucidate the importance of this pathway in modulating antitumor immunity.

Regulatory T Cells The clonal deletion of autoreactive thymocytes during development is not complete, and therefore

48

peripheral mechanisms to control self-reactive T cells and induce immunologic self-tolerance are necessary. In multiple studies, Tregs have been shown to mediate this process as well as to curtail ongoing exuberant immune responses. This population of T cells arises in the thymus and constitutes approximately 5–10% of circulating CD4+ T cells (Bopp et al. 2007; Josefowicz et al. 2012). Studies have identified a transcription factor forkhead box P3 (FOXP3, also known as scurfin) as the most crucial regulator for Treg (Fontenot et al. 2003; Hori et al. 2003; Khattri et al. 2003). Mice with defective Foxp3 gene show severe autoimmune syndromes, while overexpressing Foxp3 can convert naïve T cells toward a Treg phenotype. Multiple subpopulations of Treg termed iTreg and nTreg have also been identified, representing cells that have been “induced” to differentiate into Treg cells in the periphery and those that occur “naturally” in the thymus, respectively (Sakaguchi 2004; Bilate and Lafaille 2012). The nTreg population appears to influence the formation of iTreg through a process termed “infectious tolerance,” whereby nTreg suppress self-reactive T cells by cell contact-dependent mechanisms, conferring the suppressive phenotype of an iTreg to these previously reactive cells. These iTreg then secrete soluble mediators such as IL-10 and TGF-β, amplifying peripheral tolerance including the downregulation of bystander T cells with different antigen specificities (Jonuleit et al. 2002; Stassen et al. 2004). Depletion of CD25+CD4+ T cells in experimental models has been found to enhance antitumor immunity, but also can result in severe autoimmune sequela. Not surprisingly, increased numbers of Treg have been identified in patients with a variety of cancers, and the frequency of tumor-inflitrating Treg has been positively associated with poor prognosis (Gray et al. 2003; Li et al. 2005; Petersen et al. 2006). Results reported for a phase I trial for patients with type 1 diabetes, however, did not demonstrate significant improvement in patients’ metabolic functions following adoptive transfer of polyclonal Treg cells (Bluestone et al. 2015). The importance of Treg in immunotherapies thus remains unclear, and the depletion of Treg alone might not be sufficient to induce antitumor responses in cancer patients.

P. F. Robbins and Y.-C. Lu

Cytokines Immune cells and other cell types such as endothelial cells and fibroblasts produce membranebound or membrane-soluble mediators termed cytokines that bind to surface receptors on other immune cells, facilitating communication and allowing for a robust, coordinated, and self-limited immune response. Like many components of the immune system, the cytokine nomenclature has evolved and changed as our understanding of these molecules has expanded. Many cytokines have retained their historical descriptive names, such as platelet derived growth factor (PDGF) and colony-stimulating factors (CSFs), even though these cytokines are now known to have a greater variety of actions than their original names would suggest. Cytokines secreted by leukocytes are typically given an interleukin (or IL) designation. A complete discussion of all known cytokines and their receptors is beyond the scope of this chapter; however, select cytokines will be discussed here with the aim of providing a context for the discussion of immunotherapeutic modulation. Functionally distinct CD4+ T helper subsets have been characterized based upon the local cytokine milieu that influences their development as well as the cytokines they produce. For example, IFN-γ and IL-12 stimulate Th1 cell differentiation, leading to IL-2 and IFN-γ secretion, macrophage activation, and the expansion and activation of specific T cells involved with in delayed-type hypersensitivity (DTH) reactions. In contrast, IL-4, IL-5, and IL-13 promote Th2-type responses that lead to the further production of IL-4, IL-10, and IL-13, which act on B cells and stimulate antibody-mediated responses as well as suppress Th1 phenotype development. CD8+ T cells also produce cytokines, predominantly of Th1 cytokines, while Treg are characterized by their secretion of IL-10 and TGF-β which then downregulate immune responses. Other cytokines such as IL-18 and IL-15 stimulate T cell memory, while IL-15 stimulates NK cell activity. In humans, the distinction between Th1 and Th2 responses is not absolute (Kelso 1995); however, the type of T cell response generated by

Immunology of Melanoma

cytokines influences immune equilibrium, i.e., whether patients are susceptible to autoimmune disease or cancer (Liblau et al. 1995; Lan et al. 2006). Recombinant cytokines have been used as treatment modalities, either singly or in combination with traditional treatments, for the therapy of cancers including melanoma (see chapter ▶ “Cytokines (IL-2, IFN, GM-CSF, etc.) Melanoma”) (Atkins 2006). Systemic administration of IL-2 appears to activate tumor-specific T cells and NK cells, and IL-12 administration may induce a Th1-type response, resulting in the subsequent production of the inflammatory cytokine IFN-γ that acts to promote antitumor T cell responses (Trinchieri 2003). In clinical trials conducted to date, the major limitations of cytokine therapy have included systemic side effects and varied individual responses to treatment. Future areas of investigation include further evaluation of local cytokine delivery, which may lead to a better understanding of the factors that influence patient response to cytokine therapy, and evaluation of the effectiveness of different cytokine combinations with other treatment modalities.

Animal Tumor Models Many animal model systems have been developed for the study of melanoma, including syngeneic/ xenograft transplantation models, physically/ chemically induced models, as well as genetically engineered mouse models (Kuzu et al. 2015; van der Weyden et al. 2016; Perez-Guijarro et al. 2017). Immunity to melanoma has been intensively studied in murine melanoma models. Evidence for specific tumor recognition by the immune system was obtained from experiments first conducted in the 1950s using murine tumors generated using the mutagen methylcholanthrene (MCA). In these studies, mice that had received a surgical resection of previously inoculated tumors could be protected against a subsequent challenge with the same tumor (Foley 1953), whereas only limited protection was observed against challenge with additional MCA tumors (Prehn and Main 1957; Basombrio 1970). Subsequent studies

49

revealed that CD8+ cytotoxic T cells were primarily responsible for mediating tumor rejection (Old and Boyse 1964; Ward et al. 1989, 1990). In recent years, one of the most common used mouse models for the study of antitumor immunity is the B16 melanoma model, a syngeneic transplantation model. Several studies have shown that non-mutated differentiation antigens are commonly recognized by T cells and can serve as targets for tumor rejection. In one study, B16 melanoma cells were injected subcutaneously into C57BL/6 mice and grown in vivo for ~2 weeks to establish large tumors (>50 mm2). These mice were then treated with (1) gp100-specific CD8+ T cells isolated from pmel-1 TCR transgenic mice, (2) recombinant vaccinia virus encoding the human gp100 peptide, and (3) recombinant human IL-2 twice daily for six doses. These three elements, antigen-specific T cells, vaccination, and a T cell growth/activation cytokine, were essential for inducing the complete regression of large, established tumors (Overwijk et al. 2003). Similarly, antigen-specific CD4+ T cells can also induce tumor regression in a comparable setting. For instance, tumor rejection was observed in B16 tumor-bearing mice treated with Th17-polarized, TRP-1-specific T cells, followed by the injections of recombinant TRP-1 vaccinia virus and IL-2 (Muranski et al. 2008). Downregulation of immune responses is a natural consequence of immunity to protect tissues from overactive immune cells; however, these same processes can also hamper tumor immunity. For example, in studies that served as the basis for clinical trials described above, treatment of mice immunized against melanoma with an antiCTLA-4 antibody was effective at inducing antitumor immunity (van Elsas et al. 1999). The antitumor responses induced by CTLA-4 blockade were mediated by CD8+ T cells, which recognize TRP-2 and probably other melanoma differentiation antigens as well (van Elsas et al. 2001). Antitumor immunity can be downregulated by suppressor cells, termed T regulatory (Treg) cells, which produce molecules with suppressive effects including IL-10 and TGF-β. Depletion of Treg cells with antibodies can enhance melanoma immunity in animal models and, when used in

50

combination with CTLA-4 inhibition, is effective at preventing the establishment of the aggressive and poorly immunogenic B16 melanoma (Sutmuller et al. 2001), and combined blockade of CTLA-4 and PD-1 is more effective than either alone at mediating the rejection of B16 melanoma (Curran et al. 2010). One potential consequence of immunity against non-mutated differentiation antigens is autoimmunity. For melanoma differentiation antigens, this cross-reactive autoimmunity is manifested as vitiligo. Immunity against the major differentiation antigens in mice, including tyrosinase, TRP-1, TRP-2, and gp100, leads to hypopigmentation and can, under some conditions, lead to blindness (Hara et al. 1995; Weber et al. 1998; Bowne et al. 1999; Overwijk et al. 1999; Colella et al. 2000; Hawkins et al. 2000). The observation that patients treated with TCRs directed against MART-1 and gp100 developed similar toxicities demonstrates the issues with strategies aimed at enhancing antitumor immunity, as these are likely to simultaneously exacerbate autoimmunity sequelae. Neoantigens may represent better targets for immunotherapy than self-antigens due to the inability of neoantigenreactive T cells to recognize normal tissues. In addition, neoantigen-reactive T cells may in general possess higher affinities than self-antigenreactive T cells, whose affinities may be limited by self-tolerance mechanisms. A series of murine tumor studies have also been carried out focused on targeting neoepitopes in attempts to alleviate many of the toxicities associated with targeting self-antigens. The results of early murine model studies have indicated that neoepitopes can represent potent tumor rejection antigens (De Plaen et al. 1988; Sibille et al. 1990; Van Pel et al. 1991; Monach et al. 1995; Dubey et al. 1997), but recent investigations of the role of neoepitope-reactive T cells in antitumor immunity have taken advantage of advances in high-throughput sequencing methods. In one of the earlies examples of this approach, candidate neoepitopes were identified by combining an algorithm used to predict peptide binding to individual MHC alleles with highthroughput sequencing of tumor cell DNA and

P. F. Robbins and Y.-C. Lu

RNA obtained from murine sarcomas generated in immune-deficient Rag2 knockout mice (Matsushita et al. 2012). Using this approach, a mutated spectrin-β2 neoepitope was identified as a dominant tumor rejection antigen in this mouse model. The results of additional murine studies indicated that vaccination against candidate tumor neoepitopes identified using whole exome sequencing (WES) could treat mice with relatively small tumor burdens (Castle et al. 2012; Gubin et al. 2014; Kreiter et al. 2015). In another example of identifying candidate neoepitopes from tumor samples, WES and RNA-seq were coupled with mass spectrometric analysis of peptides eluted from cell surface MHC molecules. Partial tumor treatment was observed in a murine tumor vaccine model system following immunization with three neoepitopes identified from the murine colorectal tumor cell line MC28 using this approach (Yadav et al. 2014).

Antigens Recognized by TumorReactive T Cells Identification of Tumor Antigens Recognized by T Cells: General Principles The identification of tumor antigens that are associated with objective clinical responses, with the eventual aim of developing more effective therapies, has been carried out using a variety of approaches. In studies first carried out in the early 1990s, tumor-specific T cells generated by repeated in vitro sensitization of lymphocytes from peripheral blood (PBL) or lymph nodes with tumor cells or by culturing tumor-infiltrating lymphocytes (TIL) were used to screen antigenpresenting cells (APCs) expressing the appropriate MHC restriction element that were also transfected with tumor cDNA libraries. Another approach involves screening of sensitized T cells or TIL for their ability to recognize APCs pulsed with peptides eluted from tumor cells or fractionated tumor cell lysates. Screening of sera from cancer patients using a method termed SEREX (serological analysis of recombinant cDNA

Immunology of Melanoma

expression libraries) (Preuss et al. 2002) has also led to the identification of thousands of target molecules (online list available at http://www2. licr.org/CancerImmunomeDB/) that are generally expressed at relatively high levels in tumor cells relative to normal tissues. These responses are presumably elicited due to the release and subsequent processing and presentation of these nonmutated cellular proteins, many of which represent intracellular proteins normally sequestered from the humoral immune system, by APC present in the tumor and draining lymph nodes. Multiple humoral antigen targets are also recognized by HLA class I- and class II-restricted T cells, indicating that this process may be involved with eliciting responses from both arms of the adaptive immune system. Many of the approaches to identify tumor-reactive T cells have involved the use of “candidate antigen” approaches. In this approach, gene products overexpressed in a cell lineage-specific manner by a particular tumor type have also been examined for the presence of peptides capable of binding to MHC I or II alleles, using published peptide-MHC-binding motifs (http://www.cbs.dtu. dk/services/NetMHCpan/, www.syfpeithi.de). Repetitive in vitro sensitization of PBL expressing the appropriate MHC allele with predicted epitopes has led to the generation of T cells that recognize previously unidentified antigens. Caution is needed when interpreting the results of these assays, however, as this method can result in the generation of T cells with relatively low avidities that may not recognize physiological antigen levels endogenously expressed by tumor cell. Investigations of the role of T cells that recognize tumor-specific neoantigens, peptides encoded by non-synonymous somatic mutations in mediating clinical responses to therapies such as immune checkpoint blockade, represent an area of increasing interest that has been revolutionized by the development of rapid and inexpensive next-generation sequencing tools. These include wholeexome sequencing, a method that involves the isolation, amplification, and sequencing of coding exons generated from genomic DNA, and RNAseq analysis, which involves the sequencing of cDNA libraries generated from tumor cell mRNA.

51

Much of the early work on tumor antigen identification was carried out in patients with melanoma, in part due to the relative ease with which melanomas can be adapted to in vitro culture. The results of more recent studies have indicated that neoantigen-reactive T cells can be identified from a variety of tumor types with frequencies that are related to the number of non-synonymous somatic mutations that are present in those tumors. Nevertheless, many ongoing studies have continued to focus on melanoma due to the relatively high response rates to adoptive immunotherapy and immune checkpoint blockade. Hopefully, paradigms established by investigations of immune responses to melanoma will be relevant to additional cancer types. As described below, tumor antigens can be grouped into four general categories: cancer germline antigens, tissue-specific differentiation antigens, overexpressed gene products, and nonself-antigens, a category that includes neoantigens and foreign antigens, predominantly antigens expressed by virally infected tumors.

Cancer Germline Antigens The first antigen identified as a target for human tumor-reactive T cells, termed MAGE (melanoma antigen)-1, was isolated by screening a genomic DNA library generated from a melanoma issue culture cell line using an autologous HLA-A1 restricted tumor-specific CD8+ clone (van der Bruggen et al. 1991). The isolated gene was non-mutated and was a member of a large, previously unidentified gene family. The recognized T cell epitope, EADPTGHSY, corresponds to amino acids 161–169 of the MAGE-1 protein (Traversari et al. 1992) (see Table 1). Another HLA-A1restricted CD8+ clone recognized MAGE-A3, a MAGE family member with 73% similarity to MAGE-1 (Gaugler et al. 1994). Tumor-reactive T cells were subsequently shown to recognize products of several additional members of the MAGE gene family (Van der Bruggen and Van Den Eynde 1994; Zorn and Hercend 1999). These antigens, which are expressed exclusively in testis and placenta but not in other normal adult tissues, have been termed cancer germline antigens. Cancer germline antigens are generally expressed in

52

P. F. Robbins and Y.-C. Lu

Table 1 Neoantigens recognized by human tumor-reactive T cells Tumor 1876 1913 1913 2098 2098 2098 2098 2197 2197 2197 2197 2197 2224 2359 2369 2369 2369 2556 2556 2591 3309 3309 3466 3466 3466 3466 3466 3466 3678 3678 3678 3678 3678 3703 3713 3713 3713 3713 3713 3713 3713 3713 3713 3713 3713 3716 3716 3784

Gene OSBPL8 HLA-A11 CDKN2A GAS7 GAPDH CSNK1A1 HAUS3 MED15 NCAPH2 CUL4B SNRPA OLFML3 KPNA5 KIF2C PPP1R3B PLEKHM2 DOPEY2 RAC1 MYH14 POLA2 MATN2 CDK12 COL18A1 TEAD1 ERBB2 PDZD8 PXMP4 KHSRP UGGT2 FBXO21 XPNPEP1 CORO7 RECQL5 NSDHL SRPX WDR46 CENPL HELZ2 PRDX3 GCN1L1 PLSCR4 AFMID SEC22C TPX2 AHNAK TFDP2 ZMYM4 GNB5

HLA restriction element Unknown class I – HLA-A*11:01 HLA-A*02:01 HLA-A*02:01 HLA-A*02:01 HLA-A*02:01 HLA-B*51:01 HLA-B*51:01 Unknown class I Unknown class I Unknown class I HLA-A*02:01 HLA-A*02:05 HLA-A*01:01 HLA-A*01:01 HLA-A*26:01 HLA-A*01:01 HLA-A*01:01 HLA-C*07:01 HLA-A*11:01 HLA-A*11:01 A*02:01 A*02:01 A*02:01 B*44:02 B*39:01 B*39:01 A*02:01 C*14:02 A*03:01 B*51:01 B*44:02 A*02:01 A*02:01 A*02:01 A*29:02 A*29:02 A*29:02 A*29:02 A*29:02 A*29:02 B*44:03 B*44:03 A*02:01 B*15:01 B*15:01 B*07:02

Amino acid change p.D735N p.S35F p.G74fs, p.G74C p.H229Y p.M133I p.S27L p.T160A p.P681S p.S208Y p.L189F p.G38C p.L355F p.P384S p.A16T p.P176H p.H902Y p.P2168L p.P29S p.A600V p.L420F p.E226K p.E928K p.S306F p.L388F p.H458Y p.I311N p.S176C p.P592L p.P882L p.S250Y p.S663T p.S33L p.P816L p.A290V p.P55L p.T300I p.P79L p.D614N p.P101L p.P769L p.R247C p.A52V p.H218Y p.H458Y p.S4460F p.A406T p.H432Y p.P377L

Reference Kalora et al. (2018) Huang et al. (2004 Huang et al. (2004) Zhou et al. (2005) Zhou et al. (2005) Robbins et al. (2013) Robbins et al. (2013) Kalora et al. (2016) Kalora et al. (2016) Kalora et al. (2018) Kalora et al. (2018) Kalora et al. (2018) Unpublished data Lu et al. (2014) Lu et al. (2013) Robbins et al. (2013) Unpublished data Unpublished data Unpublished data Lu et al. (2014) (Robbins et al. 2013) Robbins et al. (2013) Parkhurst et al. (2017) Cohen et al. (2015) Cohen et al. (2015) Unpublished data Unpublished data Unpublished data Parkhurst et al. (2017) Parkhurst et al. (2017) Parkhurst et al. (2017) Parkhurst et al. (2017) Parkhurst et al. (2017) Cohen et al. (2015) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) Prickett et al. (2016) (Prickett et al. 2016) Cohen et al. (2015) Parkhurst et al. (2017) Unpublished data Parkhurst et al. (2017) (continued)

Immunology of Melanoma

53

Table 1 (continued) Tumor 3784 3784 3795 3795 3868 3868 3903 3903 3903 3919 3919 3919 3998 3998 4000 4000 4000 4000 4084 4155 4155 4189 4189 4202

Gene KIF16B SON NRAS RBBP6 GANAB SLC5A3 PHKA1 KIF1BP CCAR2 TRIP12 CFDP1 TRIP12 MAGEA6 MED13 HIVEP2 AMPH GPD2 DBT DHX34 RSL1D1 SFMBT1 DDX3X HEYL RANBP2

HLA restriction element B*07:02 B*07:02 A*01:01 A*01:01 A*02:01 Unknown class I B*38:01 B*38:01 B*38:01 A*01:01 A*01:01 Unknown class II A*30:02 A*30:02/B*15:01 A*02:01 A*02:01 B*57:01 B*57:01 Unknown class II C*01:01 Unknown class II C*08:02 Unknown class II Unknown class I

between 10% and 50% of multiple tumor types that include melanoma, breast, prostate, and esophageal cancer. The MAGE family of genes has now been grouped into subfamilies, and the MAGE-1 gene, now termed MAGE-A1, has been grouped with a closely related set of 15 genes termed MAGE-A. The entire human MAGE gene family consists of 55 genes that have been grouped into the MAGE-A-K, MAGE-L2, and NECDIN genes and genes families (Chomez et al. 2001). Cancer germline antigens that possess similar expression patterns to the MAGE gene family, but that are unrelated to this family, have also been identified. These include the BAGE (Boel et al. 1995) and GAGE (Van den Eynde et al. 1995) families of genes, as well as NY-ESO-1, a gene that was initially identified using SEREX techniques (Chen et al. 1997). Additional studies have indicated that NY-ESO-1 might represent a

Amino acid change p.L1020P p.R1927C p.Q61K p.R793C p.S320F p.P587L p.P34L p.P246S p.H227Y p.F1577S p.P128S p.F1577S p.E168K p.P1691S p.P1682L p.E83K p.A332V p.A2V p.R300H p.R183C p.P323S p.A349fs, pR351fs p.S39W p.K3185R

Reference (Parkhurst et al. 2017) Parkhurst et al. (2017) Unpublished data Unpublished data Cohen et al. (2015) Unpublished data Parkhurst et al. (2017) Parkhurst et al. (2017) Parkhurst et al. (2017) Cohen et al. (2015) Unpublished data Unpublished data Gros et al. (2016) Gros et al. (2016) Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data Unpublished data

particularly immunogenic tumor antigen, as IgG antibodies directed against NY-ESO-1 were detected in the sera 10 of 12 patients with NYESO-1 positive tumors (Jager et al. 1999), whereas antibody responses against other human tumor antigens have been observed only infrequently. Tumor burden has been correlated with the titer of anti-NY-ESO-1 antibodies (Valmori et al. 2000), and the presence of anti-NY-ESO-1 antibodies was associated with CD8+ T cells reactive against the HLA-A*02:01-restricted NYESO-1:157–165 epitope (Jager et al. 1998, 2000). A screening study was also carried out to evaluate the effects of altering the TCR complementarity-determining region (CDR) amino acid sequences of 1G4, a TCR directed against the NYESO-1:157–165 epitope, in an attempt to enhance the affinity of the TCR for cognate antigen (Robbins et al. 2008). A modified TCR with a modestly enhanced activity in retrovirally transduced CD8+

54

T cells but a significantly enhanced activity in transduced CD4+ T cells was then used to carry out a small-scale phase II adoptive immunotherapy trial. Complete responses lasting 2 years or more were observed in 2 of 20 patients with metastatic melanoma and 2 of 18 patients with metastatic synovial cell sarcoma in response to autologous lymphocytes transduced with the affinity-enhanced 1G4 TCR NY-ESO-1-reactive TCR described above (Robbins et al. 2015). Nevertheless, the relatively low response rate, coupled with the observation that less than 5% of common cancers appear to express homogeneous levels of NY-ESO-1, limits the applicability of this therapy to a relatively small patient population (Kerkar et al. 2016). Durable clinical responses have also been observed in patients enrolled in clinical trials utilizing TCRs that target more broadly expressed CG antigens such as MAGE-A3; however, unexpected severe toxicities and patient deaths associated with the cell transfers were also observed in these trials, findings likely resulting from the expression of cross-reactive epitopes in vital normal tissues (Morgan et al. 2013; Linette et al. 2013).

Tissue-Specific Differentiation Antigens Tissue-specific gene products that are expressed in melanomas as well as precursor normal melanocytes present in the skin and retina, termed melanocyte differentiation antigens (MDAs), are frequently targeted by melanoma-reactive T cells. Many of the MDAs, including gp100, tyrosinase (Wolfel et al. 1994), TRP-1, TRP-2, and OA-1, are previously known enzymes or structural proteins located within melanosomes and involved in the synthesis of melanin pigment. Immune responses directed against gp100 and MART-1/ Melan-A, discussed below, are representative of those seen against this group of melanoma antigens. The gp100 MDA was identified as an immunodominant target of melanoma-reactive T cells using multiple approaches. In one study, pools of peptides were eluted from MHC molecules on the surface of melanoma cells, fractionated on HPLC columns, tested for their ability to sensitize APC for melanoma-specific Tcell recognition, and sequenced

P. F. Robbins and Y.-C. Lu

using mass spectrometry (Cox et al. 1994). A peptide derived from the gp100 protein, YLEPGPVTA (gp100:280–288), was found to be recognized by five out of five cultured CD8+ Tcell lines. Separately, a gene-based approach also identified gp100 as a melanoma antigen recognized by CD8+ TIL (Kawakami et al. 1994a). In additional studies, four gp100-reactive HLA-A2-restricted TIL lines were screened for their ability to recognize APC that were pulsed with 169 individual synthetic candidate epitopes from the gp100 protein sequence, selected using published HLA-A2 peptide-binding motifs. Using this approach, three additional gp100 epitopes, ITDQVPFSV (gp100:209–217), KTWGQYWQV (gp100:154–162), and VLYRYGSFSV (gp100:476–485), were found to be recognized by melanoma-reactive T cells (Kawakami et al. 1995). Interestingly, the gp100:209–217 and gp100:154–162 peptides did not conform to the ideal HLA-A2-binding motif, an observation that led to further studies described later in this chapter examining the immunogenicity of altered peptide ligands. The gene encoding a previously unknown MDA, designated MART-1/Melan-A (hereafter referred to as MART-1), was discovered by screening melanoma cDNA libraries with HLAA2-restricted melanoma-specific CD8+ T cells (Coulie et al. 1994; Kawakami et al. 1994b). The product of the MART-1 gene, a relatively small 118 amino acid protein, was found to be expressed in 80–90% of fresh melanomas and cultured melanoma cell lines (Marincola et al. 1996). The majority of melanoma-reactive, HLA-A2restricted TIL were shown to recognize MART1, indicating that this is an immunodominant antigen (Kawakami et al. 1994c). In an attempt to identify the recognized epitope(s), a set of 23 candidate peptides from MART-1 were synthesized and tested for their ability to stimulate MART-reactive T cells. An HLA-A2-restricted, MART-1-reactive TIL culture was initially found to recognize a single nonamer peptide AAGIGILTV(MART-1:27–35) as well as two overlapping decamers EAAGIGILTV and AAGIGILTVI. Subsequent observations indicated that nine of ten HLA-A2-restricted TIL cultures recognizing the MART-1 protein all reacted

Immunology of Melanoma

against the MART-1:27–35 epitope. Furthermore, the MART-1:27–35 nonamer, but not the overlapping decamers, was identified among HLAA2-associated peptides eluted from the surface of melanoma cells, indicating that it represented the predominant naturally processed MART-1 epitope (Skipper et al. 1999). The MART1:27–35 peptide binds HLA-A2 with a relatively low affinity and contains an alanine residue rather than the preferred leucine, isoleucine, or methionine residues typically found at the dominant position 2 MHC anchor position in high-affinity HLA-A2-associated peptides. This is not entirely unexpected, since high-affinity T cell interactions with self-peptide-MHC complexes are generally selected against during thymic development. Interestingly, however, relatively high frequencies of T cells recognizing MART-1:27–35 have been observed in the peripheral blood of many HLAA2+ melanoma patients. In one study, the frequency of MART-1-reactive T cells was >1:2000 in the peripheral blood of four out of nine melanoma patients and on the order of 1:40,000 PBMC in the remaining patients (Anichini et al. 1999). In a second study, detectable levels of MART-1-reactive T cells were found in the peripheral blood of 10 out of 13 melanoma patients and 6 out of 10 healthy HLAA2+ individuals (Pittet et al. 1999). Cross-reactivity of MART-1-reactive T cells with a variety of peptides derived from viral or other sources has been observed, providing a potential explanation for the relatively high frequency of T cells recognizing this epitope (Loftus et al. 1996; Dutoit et al. 2002). Multiple MHC class II-restricted epitopes derived from MDAs have also been described, including two tyrosinase epitopes recognized in the context of HLA-DRβ1*0401 (Topalian et al. 1996) and epitopes with nearly identical sequences derived from TRP-1 and TRP-2 and recognized in the context of HLA-DRβ1*1502 (Robbins et al. 2002). Several class II-restricted epitopes derived from the gp100 glycoprotein have also been described (Li et al. 1998; Touloukian et al. 2000; Parkhurst et al. 2004). These studies employed candidate antigen, peptide elution, or cDNA library screening

55

approaches for antigen identification. The recognized epitopes appeared to be naturally processed and presented, as demonstrated by the ability of CD4+ T cells reactive with these epitopes to also recognize intact tumor cells or antigen-presenting cells that had been pulsed with tumor cell lysates. The MDAs have been targeted in multiple clinical trials involving both passive adoptive T cell transfer and active immunization approaches. Objective tumor regressions were seen in 30% of patients with metastatic melanoma who received autologous peripheral blood mononuclear cells transduced with a high-avidity MART-1-reactive TCR and 19% of patients who received an avid gp100-reactive TCR (Johnson et al. 2009). All of the patients had subsequent tumor recurrences with the exception of a single patient in each trial arm, however, (unpublished observations), and many of these patients developed severe dose-limiting toxicities that appeared to result from the ability of the administered T cells to recognize normal melanocytes in the eyes and ears expressing these gene products (Johnson et al. 2009) (Fig. 1). The observation that many of the epitopes targeted using these approaches bound to cognate MHC molecules with relatively low affinities led

Fig. 1 Autoimmunity induced by passive immunization against the tyrosinase family protein TRP-1. (a) A C57BL/ 6 mouse that received a control antibody. (b) A C57BL/6 mouse that received an injection of a mouse antibody recognizing TRP-1

56

to studies aimed at identifying altered peptide ligands (APL) with enhanced immunogenicity. Substitution of an optimal methionine for a suboptimal threonine at position 2 of the native HLA-A*02:01-restricted gp100:209–217 peptide (gp100:209–217, 210M) led to a ninefold enhancement in peptide-MHC-binding affinity (Parkhurst et al. 1996). The results of a clinical trial in which melanoma patients were vaccinated with the modified gp100:209–217 (210M) revealed that in 44% of immunized patients, 1–10% of peripheral CD8+ T cells recognized the modified and native peptide and in 17% of patients, greater than 10% of peripheral CD8+ T cells recognized both peptides. These results contrasted with immunization studies using the native sequence gp100 peptide, in which generally 1/mm2), and age revealed an independent impact of GEP and AJCC stage on DFS. In the most recently reported analysis to date (Zager et al. 2018), the GEP was assessed in an independent cohort of 523 primary melanoma patients with at least 5 years of follow-up or documented relapse. Class 1 patients had a 98% 5-year MSS rate, compared with 78% for class 2 patients. Multivariate Cox regression analysis of RFS and DMFS examining GEP, tumor thickness, mitotic rate (>1/mm2), ulceration, and SLN status revealed an independent impact of tumor thickness, SLN status, followed by GEP, for DFS and DMFS. One major limitation of this assay is its strong dependency on tumor cell content making it rather unlikely of help particularly for the large number of thin melanomas which eventually progress. In addition, a separate nine-gene qRT-PCRbased test termed MelaGenix has been offered by NeraCare to predict melanoma prognosis. The assay was developed on fresh-frozen melanomas and tested on a training subset of 38 cases, a training cohort of 91 cases, and a validation cohort of 44 cases (Brunner et al. 2013). A dichotomized risk score was developed using this gene signature and was significantly predictive of OS in the training cohort. Multivariate Cox regression analysis of molecular risk score, AJCC stage, Clark level, age, and sex, revealed AJCC stage, followed by the risk score, as independently predictive of OS.

D. Schadendorf et al.

The signature-based GEP score was recently confirmed on formalin-fixed paraffin-embedded (FFPE) melanoma and was shown to be independent of hospital-specific tissue fixation procedures and highly stable even in aged FFPE samples. Interestingly, the MelaGenix GEP score is determined in whole FFPE tissue sections and does not require microdissection of tumor tissue. Out of the GEP score, seven genes originate from tumor stroma making this assay particularly suitable for analyzing the prognosis of primary cutaneous melanomas. Taken together, these GEP assays await evaluation in prospectively collected cohorts with defined eligibility. In addition, to date, a predictive signal for either of the assays in identifying benefit to any adjuvant therapy regimen for melanoma (e.g., interferon alpha, ipilimumab, anti-PD-1 antibody, or targeted therapy) has not been demonstrated. As a result, at this time, use of GEP assays has not been recommended by the National Comprehensive Cancer Network (NCCN) guidelines for melanoma patients outside of the setting of a clinical trial.

IHC Analysis of Melanoma Prognostic Markers Beyond gene expression profiles, several putative prognostic factors for melanoma have been assessed for their predictive impact using IHC analysis. Specifically, two multi-marker signatures have been developed and undergone more extensive analysis and will be discussed in detail here. In 2009, Rimm and colleagues reported the development and performance of a melanoma prognostic model following an analysis of 38 candidate markers using the automated quantitative analysis (AQUA) method (Gould Rothberg et al. 2009a). Assessment of these markers in a tissue microarray (TMA) cohort, including a training sample of 192 cases, identified a consistent prognostic signal for five markers (ATF2, p21/WAF1, p16/CDKN2A, β-catenin, and fibronectin). An algorithm was developed to combine marker expression scores, and a dichotomized analysis of low- versus high-risk subgroups based on

Biomarkers for Melanoma

marker expression showed a significant difference in MSS between the two subgroups in the training set. This differentiation was also observed in the validation set, with a 10-year survival of greater than 90% in the low-risk group versus 60% in the high-risk group, albeit with a trend toward statistical significance (P = 0.09). Multivariate analysis of MSS that included the multi-marker score, tumor thickness, age, anatomical site, sentinel lymph node status, and receipt of nonsurgical therapy revealed (in an order of descending statistical significance) receipt of nonsurgical therapy, age, sentinel lymph node status, the multi-marker score, and tumor thickness to be significantly predictive of MSS. Separately, also in 2009, Kashani-Sabet and colleagues reported the performance of a threemarker IHC assay incorporating the following markers: NCOA3, SPP1, and RGS1 (KashaniSabet, Venna, et al. 2009). This assay was assessed initially in a TMA cohort of 395 primary melanoma patients from the USA and separately in tissue sections from a cohort of 141 patients from two German centers. Marker expression was assessed both by pathologist scoring and using a digital imaging platform. An index was developed in each cohort to combine marker expression levels. In the US cohort, by multivariate logistic regression analysis, the multi-marker expression score was independently predictive of SLN status, following age, but with an impact greater than tumor thickness. Multivariate Cox regression analysis of tumor thickness, SLN status, ulceration, Clark level, age, gender, anatomical location, and multi-marker score revealed the multi-marker score as the top factor predicting disease-specific survival (DSS). A dichotomization of multimarker scores revealed a 5-year survival of 96% in the low-risk group versus 60% in the high-risk group. Separately, the multi-marker score was also independently predictive of DSS in the German cohort, surpassing tumor thickness and other available factors. More recently, this three-marker IHC assay was assessed on tissues collected as part of the Eastern Cooperative Oncology Group trial E1690 examining the utility of two doses of interferon alpha versus observation in patients with resected,

85

high-risk melanoma, including eligible patients with a tumor thickness of greater than 4 mm or node-positive disease (Kashani-Sabet et al. 2017). The tissue cohort from the E1690 trial included both primary melanoma specimens and lymph node metastases. IHC analysis was performed to determine expression of NCOA3, SPP1, and RGS1, and marker analysis was assessed using a digital imaging platform. Once again, an index was developed to combine marker analysis and was dichotomized to split the cohort into low-risk and high-risk subgroups. By Kaplan-Meier analysis, the multi-marker score was significantly predictive of relapse-free survival (RFS) and OS in the entire cohort. By stepwise multivariate Cox regression analysis, multi-marker score was the only factor significantly predictive of RFS and was followed by tumor thickness as the only factors significantly predictive of OS in the entire cohort. When a potential interaction between marker expression and treatment assignment was analyzed, the interferon-treated arms (combining both low-dose and high-dose cohorts) had a significantly improved RFS versus the observation arm in the molecularly defined low-risk subgroup. These results demonstrated the independent prognostic significance of this assay in a prospectively collected cohort amassed in a cooperative groupled clinical trial and identified a potential subset of patients that could benefit from systemic therapy with IFN. Additional validation studies of this IHC assay are currently planned in other clinical trial cohorts. In conclusion, significant progress has also been made in the development of molecular prognostic markers for primary melanoma. This effort has been facilitated by genome-wide profiling efforts of distinct stages of melanoma progression that have identified a plethora of putative prognostic markers. However, the development of molecular prognostic markers for melanoma has been hampered by the lack of large, well-annotated tissue cohorts of primary melanoma patients with sufficient follow-up. For the promising multi-marker assays developed, further development has been hampered by the lack of sufficient prospective validation of these assays. Lastly, while accurately predicting melanoma prognosis

86

would be useful, current treatment guidelines do not recommend aggressive radiographic surveillance in high-risk patient cohorts. As a result, identifying a high-risk cohort may not be clinically actionable. However, given the recent development of landmark, effective adjuvant therapies for melanoma (both immunotherapies and targeted therapies), these assays would have their greatest clinical utility in identifying subsets of patients that derive either the maximum benefit or, alternatively, no benefit from routinely performed clinical interventions.

Tumor Environment-Based Non-soluble Biomarkers Melanomagenesis and tumor progression are a complex and dynamic process that is manifested by tumor heterogeneity and a myriad of yet to be fully understood interactions within the tumor microenvironment. Historically, the most robust diagnostic and prognostic parameters in cutaneous melanoma have been primary tumor characteristics detected histopathologically on routinely processed hematoxylin and eosin-stained tissue sections. These factors include tumor thickness, ulceration, proliferation activity, lymphovascular invasion, and the presence of micrometastases. In the last two decades, tumor-associated molecular biomarkers have been identified, individually and in combination, that correlate with diagnosis and prognosis. Most recently, attention has been given to non-tumor cell markers in the primary tumor environment that facilitate local tumor progression and metastasis. In addition to tumor-infiltrating lymphocytes and other immune factors, biomarkers in the tumor stromal environment may aid in diagnosis, predict prognosis, and even serve as therapeutic targets (Jacobs et al. 2012). It is clear that the tumor microenvironment is shaped by the cross talk between mesenchymal stromal cells and immune cells (English 2013). This realm will likely serve as the next frontier of effective melanoma therapy. Melanoma tumor cell and immune cell interactions have been known to correlate with prognosis since the late 1980s when Clark et al. identified an

D. Schadendorf et al.

improved prognosis in patients with primary cutaneous melanomas that had robust “brisk” lymphocytic infiltrates (Clark et al. 1989). Further refinements of this histopathological factor revealed that there is a complex interplay between the distribution and density of the infiltrates and potential for tumor immune escape or tumor suppression (Clemente et al. 1996). Most recently, chemokines and their receptors have been described to orchestrate melanoma cell and immune cell dynamics (Neagu et al. 2015). Chemoattractant cytokines were initially identified as factors that recruited leukocytes in inflammatory and immune responses (Cyster 1999). It is now known that chemokines play a role in a broad range of melanoma pathways to tumor growth and metastasis including maintenance of tumor-initiating cells, cell proliferation, epithelial-mesenchymal transition-like processes, angiogenesis, senescence, epigenetic responses to oxidative stress, and immune evasion (Sarvaiya et al. 2013).

Tumor-Initiating Cells A relationship between melanoma-initiating cells and antitumor immunity has been identified (Schatton et al. 2010). Stem cells, also known as tumor-initiating cells, are capable of self-renewal and differentiation and are responsible for tumor development and therapeutic resistance (Schatton et al. 2008). The ATP-binding cassette (ABC) efflux transporter ABCB5 is a marker of tumorinitiating cells that has been shown to maintain slow cycling chemoresistant cells through a complex cytokine signaling pathway that includes IL1-beta, IL-8, and CXCR1, thus playing a role in stem cell maintenance and tumor growth (Wilson et al. 2014). Another marker of melanoma-initiating cells, CD133 (human prominin 1), is a transmembrane pentaspan glycoprotein that plays a role in vasculogenic mimicry and formation of a vascular niche (Lai et al. 2012). It is likely that a restricted number of tumor cells may possess the capacity to modulate tumordirected immune responses; clearer understanding of these processes will aid in the development of future therapeutic strategies.

Biomarkers for Melanoma

Epithelial-Mesenchymal Transition Melanoma has been shown to progress through a distinct epithelial-mesenchymal transition (EMT)like process (Caramel et al. 2013). This transition to cells with enhanced migration, invasiveness, resistance to apoptosis, and production of extracellular matrix components is a plastic phenomenon. Tumor cells may cycle between a differentiated state (associated with increased ZEB2 and Slug) and an oncogenic state (with high levels of ZEB1 and twist) (Li et al. 2015). Additionally, tumor cells may alter the environmental niche through release of miRNA containing melanosomes into fibroblasts (Dror et al. 2016). It is likely that the plasticity of this transition between differentiated and oncogenic states contributes to the tumor heterogeneity characteristic of melanoma. Several factors involved in this transition have been explored as potential biomarkers, including the transcription factor SNAI1 (snail1) which may lead to reduced E-cadherin expression and induction of N-cadherin (Miller and Mihm 2006). Indeed, expressions of the EMT-associated proteins N-cadherin, osteopontin, and SPARPC/osteonectin are significantly associated with the risk of metastasis (Alonso et al. 2004). Additionally, twist1 and twist2 are regulatory proteins that induce EMT and may also have a role in limiting oncogene-induced senescence (Ansieau et al. 2008). On the other hand, mechanisms of senescence induction include activation of DNA damage signaling by oncogenes and short telomeres (Bennett 2008). The irreversible arrest of proliferation associated with senescence occurs through the p53 and p16-pRB tumor suppressor pathways (Campisi and d’Adda di Fagagna 2007). In addition to adhesion marker interactions and interference with senescence, stromal-derived proteases may modulate antitumor immune responses. Melanoma-associated fibroblast production of metalloproteinases including MMP-7 decreases tumor cell susceptibility to natural killer cell-mediated tumor necrosis (Ziani et al. 2017). Immunohistochemical analysis of MMP-7 reveals increased expression in melanoma that correlates with tumor thickness and adverse prognosis (Kawasaki et al. 2007). These EMT-like processes

87

likely advance melanoma progression through promoting invasion and a proliferative advantage (Li et al. 2015).

Epigenetic Changes Epigenetics is an important mechanism by which gene expression may be modified in cancer. DNA methylation is one of the epigenetic hallmarks that is most studied. Hypermethylation of CpG islands in the promoter region leads to gene silencing and has been described for genes throughout melanoma progression and metastasis (Rothhammer and Bosserhoff 2007; Schinke et al. 2010; Tanemura et al. 2009). Loss of 5-hydroxymethylcytosine (5-hmC) in melanoma has been described as a fundamental epigenetic event that correlates with tumor progression and is associated with decreased expression of the enzyme teneleven translocase (TET) (Lian et al. 2012). Epigenetic regulation of CD73 has also been described (Wang et al. 2012). CD73 is an ectonucleotidase expressed on Tregs that along with adenosine (ADO) has been implicated in tumor-associated immunosuppression. ADO levels in the extracellular microenvironment are usually low; however, high levels have been identified at the tumor stromal interface. CD73+ is associated with increased ADO production and has been correlated with poor prognosis (Wang et al. 2012). These examples of epigenetic modifications in melanoma are but a few of those that have been recently described. As techniques to detect epigenetic biomarkers evolve, this field may provide important novel information regarding prognosis and potential response to therapy (Greenberg et al. 2014).

Immune Escape Mechanisms Recruitment of suppressor immune cells, including Tregs and tumor-associated macrophages, may facilitate tumor cell evasion of the immune system (Buchbinder and Hodi 2015). Insufficient co-stimulation of the immune system by tumor cells limits antitumor immunity. For example,

88

CTLA-4 is a key inhibitory receptor that blocks co-stimulation. CTLA-4 blockade with ipilimumab is an effective therapy that increases antitumor immune responses through enhanced effector T-cell function and inhibition of Treg activity. The impressive clinical responses seen with immunotherapy, in particular anti-CTLA-4 and anti-PD-1, are based on the critical role the immune system plays in melanoma tumor progression. However, many patients’ tumors are refractory to these therapies. Dendritic cell-mediated tumor evasion mechanisms may contribute to this; tolerized dendritic cells drive Treg differentiation and may establish a milieu of immune privilege. Tumor and stromal cell-derived cytokines, including TGF-beta and prostaglandin-E2, may exert immunosuppressive effects by contributing to the establishment of immune tolerance (Balsamo et al. 2009; Pietra et al. 2012). Similarly, indoleamine (2,3)-deoxygenate (IDO) generates an immunosuppressive tumor microenvironment by suppressing effector T cells and actively tolerizing the tumor microenvironment by promoting Treg development (Holtzhausen et al. 2015). The Wnt-beta-catenin signaling pathway promotes dendritic cell tolerization through induction of IDO. In sentinel lymph nodes, a decrease in interdigitating dendritic cells, antigen-presenting cells involved in T-cell activation, has been associated with a poor prognosis (Cochran et al. 2004). Melanoma metastasis in sentinel lymph nodes is associated with a higher frequency of Foxp3+ CD4+ CD25 high Tregs and IDO-expressing dendritic cells (Lee et al. 2011). Increased IDO expression by dendritic cells in sentinel lymph nodes has been shown to correlate with adverse prognosis (Speeckaert et al. 2012). This cytokine microenvironment is likely to determine the functional immune status of the sentinel lymph nodes. As such, these complex stromal immune interactions remain a topic of intense investigation. There are many other pathways that are being elucidated in the immune evasive mechanisms employed by melanoma. For example, the surface glycoprotein CD47 is a regulator of melanoma immune evasion. This transmembrane integrinassociated protein is present on all normal cells

D. Schadendorf et al.

and upregulated in some melanoma cells; ligands include thrombospondin-1 (TSP1) and signal-regulatory protein alpha (SIRP-alpha) (Brown and Frazier 2001). Increased expression of CD47 in melanoma has been associated with increased risk of metastasis and poor survival (Fu et al. 2017). SIRP-alpha is expressed on myeloid cells including dendritic cells; binding of CD47 to SIRPalpha leads to reduced macrophage phagocytosis allowing melanoma to evade elimination by innate immunity. Indeed, CD47 expression on tumor cells has been coined a “don’t eat me” signal enhancing tumor cell survival by inhibiting phagocytosis by macrophages (Jaiswal et al. 2009; Willingham et al. 2012). TSP1 is highly expressed in tumor stroma and has many functional interactions other than with CD47. CD47TSP1 interactions have been associated with modulation of nitric oxide (NO) signaling and vascular responses (Isenberg et al. 2006). Functional CD47-TSP1 interactions have also been associated with an increased capacity for self-renewal when CD47 is highly expressed on tumor-initiating cells (cancer stem cells) (Kaur and Roberts 2016). Initially identified as an integrin-associated protein, CD47 also is necessary for ligand recognition by a variety of integrins including alphavbeta3, alpha2beta1, and alpha4beta1 (Brown and Frazier 2001). These studies and many others have identified the CD47-SIRPalpha interaction as a promising innate immune checkpoint; however, there are important mechanistic issues still to be resolved. This is a rapidly expanding field which will likely have a major impact on melanoma therapy (Matlung et al. 2017). In summary, the process of melanoma tumorigenesis, tumor progression, and metastasis is a dynamic interplay between the tumor and the surrounding microenvironment including stromal and immune elements. As we learn more about these interactions, it is likely that melanoma microenvironment biomarkers will be clinically deployed. Future investigations of melanoma microenvironment biomarkers will face the same challenges posed in the past: it is critical to the understanding of these pathways that tissuebased studies are performed that identify cell subsets. This will require an ever-diligent attention

Biomarkers for Melanoma

to tumor sample management and annotation. Additionally, the tissue-based technologies that provide for simultaneous detection of multiple biomarkers localized to cell subsets and individual cells will be required for a complete understanding of the complex tumor, immune, stromal interactions in melanoma tumor progression.

Soluble Biomarkers

89

therapeutic interventions in advanced inoperable disease. Due to its low specificity, false-positive results of elevated LDH levels are common and originate from conditions like hemolysis, muscle or liver disease, injuries, or other pathologies. In patients with low or clinically undetectable tumor burden, LDH serum levels are generally normal; thus, serum LDH cannot be recommended as a marker of minimal residual disease or early relapse (Eisenstein et al. 2018).

Lactate Dehydrogenase (LDH)

S100B At present, the most widely used prognostic serum biomarker in the clinical care of melanoma patients is lactate dehydrogenase (LDH). LDH is an unspecific biomarker indicating high metabolism and/or high tumor load in a variety of tumor entities including melanoma (Manola et al. 2000; Egberts et al. 2012). Studies comparing different serological markers including LDH, S100B, and MIA in multivariate analysis showed LDH as the strongest independent prognostic factor in stage IV melanoma patients (Deichmann et al. 1999). Due to its high prognostic significance together with easy, cost-efficient, and widely distributed detection techniques, serum LDH is the only molecular marker so far that has been incorporated into the melanoma staging and classification system of the AJCC, beginning with the edition of 2001 (Balch et al. 2001a). The implementation of serum LDH into this classification system took place after LDH was demonstrated as an independent predictor of overall survival in a very large cohort of nearly 8000 patients with advanced metastatic melanoma (Balch et al. 2001b). This cohort showed a 1-year survival rate of 65% for patients with normal serum LDH, whereas patients with elevated LDH levels had a significantly reduced 1-year survival rate of 32% only. Serum LDH is therefore commonly used in the clinical routine of melanoma patient care, in particular in patients with advanced metastatic disease. It serves as a reliable prognostic marker before the start of a new systemic therapy and in regular intervals thereafter in order to monitor treatment response. It is moreover used as a common stratification parameter in randomized clinical trials testing

The S100 protein is a 21-kd thermolabile acidic dimeric protein which was originally isolated from tissue of the central nervous system. It consists of two subunits, alpha and beta, in different pairings. S100 is of functional importance for the assembly of microtubules and interacts in a calcium-dependent manner with the tumor suppressor gene p53. The beta subunit (S100B) is expressed in cells of the central nervous system as well as in cells of the melanocytic lineage. Therefore, the serum concentration of S100B has been described as a biomarker of central nervous system damage (Persson et al. 1987) as well as of the presence of melanoma metastasis (Guo et al. 1995). The serum level of S100B is an indicator of tumor burden and therefore correlates with the clinical stage of melanoma patients. With regard to prognosis, S100B is a useful marker in melanoma patients with presence of metastases (Schultz et al. 1998; Hauschild et al. 1999a), but fails to provide prognostic significance in patients with microscopic disease, as well as in patients who are clinically tumor-free after surgery (Guo et al. 1995; Acland et al. 2002; Egberts et al. 2010). However, a meta-analysis of 22 studies including a total of 3393 melanoma patients revealed S100B as a significant prognostic factor in all clinical stages of melanoma, even in stages I–III (Mocellin 2008). Despite S100B is a melanoma serum marker of higher specificity than LDH, it still has limitations not only by elevated levels due to central nervous system damage but also by liver or cardiovascular diseases (Vaquero et al. 2003; Li et al. 2011). The stringent correlation of serum

90

S100B concentrations with tumor burden, however, renders it a useful marker for the monitoring of treatment response in patients with advanced metastatic melanoma (Hauschild et al. 1999b; Egberts et al. 2012); see Fig. 2. Its use in the routine clinical care of melanoma patients is still mainly restricted to European countries.

Other Serum Biomarkers (CRP, FGF, IL-8, MIA, SAA, VEGF, YKL-40) Worldwide, the serum biomarker which is most widely implemented into the clinical routine of melanoma patient care is LDH. S100B is the second often used serum biomarker and is of similar prognostic significance as LDH with higher specificity for melanoma versus other cancer entities. Various other serum factors have been investigated for their prognostic significance in melanoma. Up to now, none of them revealed a higher sensitivity-specificity profile than LDH or S100B. An extensively studied serum protein named melanoma inhibitory activity (MIA) was Fig. 2 Monitoring of the course of melanoma disease by the serum marker S100B. (From Ugurel 2005)

D. Schadendorf et al.

originally detected in melanoma cell culture supernatants (Bogdahn et al. 1989) and was shown to exert an important role in cell-matrix interaction, invasion, and metastasis (Blesch et al. 1994). Studies comparing MIA and S100B demonstrated that S100B is superior to MIA in its value as an early indicator of tumor progression, relapse, or metastasis (Deichmann et al. 1999; Krähn et al. 2001). Proangiogenic factors like vascular endothelial growth factor (VEGF), fibroblast growth factor (FGF), and interleukin8 (IL-8) have been demonstrated to reveal prognostic significance in case of elevated serum levels (Ugurel et al. 2001a; Sanmamed et al. 2014; Yuan et al. 2014). Besides, proteins associated with antigen presentation and recognition like HLA molecules, as well as receptor or ligand molecules associated with anticancer immune response like NKG2D, CEACAM, and others, have been described as soluble variants detectable in sera from melanoma patients and correlating to the patients’ prognosis (Ugurel et al. 2001b; Rebmann et al. 2002; Paschen et al. 2009; Sivan et al. 2012). Members of the acute phase proteins, like C-reactive protein (CRP) and serum amyloid

Biomarkers for Melanoma

A (SAA), are associated with inflammatory processes and have also been described as prognostically significant serum factors in melanoma (Deichmann et al. 2004; Findeisen et al. 2009). YKL-40, a chitinase-like glycoprotein produced by cancer cells as well as by inflammatory cells, has recently been described as a prognostic marker in melanoma, giving promising results even in tumor-free early-stage patients (Schmidt et al. 2006). However, in comparative studies, YKL-40 showed an inferior prognostic value as compared to S100B (Egberts et al. 2012). Moreover, YKL-40 was reported to lose its prognostic significance in patients treated with interferons (Krogh et al. 2016). None of the abovementioned serum markers succeeded in reproducible superiority compared to LDH and/ or S100B, resulting in no further development of these biomarkers into clinically applicable test systems.

Circulating Tumor Cells (CTCs) Like cells of other cancer entities, melanoma cells are known for their ability to leave their tissue of origin and enter the blood stream as free floating cells. These circulating tumor cells (CTCs) become detectable in blood draws and may serve as biomarkers of tumor burden, prognosis, and treatment response. The detection and capture of CTCs in the blood stream mainly rely on surface protein or antigen expression. However, in melanoma, this approach is limited due to the inter- as well as intraindividual heterogeneity of the antigen profiles presented by the patients (Khoja et al. 2014; De Souza et al. 2017). Thus, various antigens like tyrosinase, Melan-A/MART-1, gp100, MAGE-3, and EpCAM have been used alone or in different panel combinations to identify CTCs in melanoma patients. Moreover, various techniques have been used to capture and quantify the identified CTCs including magnetic or electrophoretic separation systems, microfluidics-based techniques, filtration approaches, and cell exclusion systems (Khoja et al. 2015; De Souza et al. 2017; Lim et al. 2018). This inconsistency in methodologies resulted in a high variation of results

91

reported of CTC numbers in melanoma patients of different disease stages and led to a questionable clinical applicability of CTC detection and quantification as a biomarker in melanoma (Nezos et al. 2011). However, CTCs have been demonstrated in multiple studies using non-comparable methodologies to be of prognostic value in melanoma patients; see Fig. 3. A meta-analysis of 53 studies describes a correlation of the presence of CTCs in a patient’s blood stream with advanced disease stage and impaired progression-free as well as overall survival (Mocellin et al. 2006). One study described that melanoma patients with a positive detection of CTCs after completion lymph node dissection revealed a higher risk of disease recurrence (Mocellin et al. 2004). Moreover, CTC numbers were shown to be of use in the monitoring of systemic therapies (Khoja et al. 2013). Despite these promising results, the clinical use of CTC detection and quantification as a reliable biomarker will remain limited until techniques will be found providing valid and reproducible results applicable to a majority of patients.

Circulating Tumor DNA (ctDNA) The abovementioned problems in the capture and quantification of CTCs could be overcome by the indirect detection of tumor cells via DNA sequences which are present in tumor cells only and not in benign cells and tissues. In melanoma, the common driver gene mutations like in BRAF or NRAS are perfectly suited for this purpose. The circulating tumor DNAs (ctDNA) mainly originate from apoptotic or necrotic circulating tumor cells and have been shown to be detected and quantified in peripheral blood samples (Schwarzenbach et al. 2011). To this end, plasma samples are superior to serum samples due to a greater extent of cell lysis during the clotting process of serum (Sorber et al. 2017). Due to the low abundance of ctDNA, its detection requires highly sensitive and specific techniques. Thus, modern detection technologies like digital droplet PCR or allele-specific ligation PCR have significantly improved the detection rate of ctDNA in

92

D. Schadendorf et al. 1.0

Probability of relapse-free survival

0 marker 0.8

0.6

1 marker

0.4

2 markers

0.2 p

0.0001

0 0

12

24

36

48

Months

Fig. 3 Prognostic utility of CTCs in blood of patients with melanoma. Kaplan-Meier curves of relapse-free survival of CTC monitoring in patients with AJCC stage III melanoma in a multicenter trial of biochemotherapy treatment. After treatment, relapse-free survival decreased significantly when

blood specimens were qRT (real-time RT-PCR)–positive for MART-1, GalNAc-T, and/or MAGE-A3 ( p = 0.0003, p < 0.0001, and p < 0.0001, respectively). The level of decrease was directly correlated with the number of positive markers ( p < 0.0001). (From Koyanagi et al. 2005)

different tumor entities. Additionally, targeted sequencing techniques like amplicon sequencing or hybrid-capture sequencing have been used, with the disadvantage of high data volumes to be processed after each analysis (Newman et al. 2014). Both the presence and the quantity (copy number) of ctDNA may serve as prognostic markers in cancer patients. In melanoma, peripheral blood samples from patients with known tumor tissue BRAF and NRAS mutational status have been analyzed for these mutations in ctDNA in various studies, revealing an association of positive results with impaired prognosis and reduced survival (Huang and Hoon 2016). ctDNA frequencies were associated with tumor burden, location of metastasis, and tumor cell metabolism (Lim et al. 2018). In melanoma, patients in early disease stages are often negative for ctDNA detection (Daniotti et al. 2004). Nevertheless, in patients in advanced disease stages and clinically detectable tumor burden, ctDNA has been shown to be a useful marker in the monitoring of treatment response and outcome. Patients starting with a

blood draw testing mutation-positive into a treatment with BRAF/MEK inhibition or anti-PD1 immunotherapy revealed impaired treatment outcomes compared to patients who were mutationnegative at baseline (Gray et al. 2015). Moreover, patients starting with mutation-positive ctDNA results but turning mutation-negative during treatment were superior in their treatment outcomes compared to patients who stayed mutation-positive during the continuous treatment course (Lee et al. 2017). However, the concordance between peripheral blood and tumor tissue is not yet satisfactory in all cases. Thus, matched plasma and tumor tissue samples from melanoma patients showed a 68% concordance only for the detection of TERT promoter mutations in ctDNA versus DNA extracted from tissue (McEvoy et al. 2017). Besides of mutations, tumor-specific gene methylation patterns of ctDNA are detectable in the peripheral blood and have been shown to serve as biomarkers in cancer patients (Warton and Samimi 2015). Tumor suppressor genes are known to be frequently inactivated by methylation as an early event in the etiopathogenesis

Biomarkers for Melanoma

93

samples from melanoma patients has been shown to exert diagnostic, prognostic, and predictive relevance (Aftab et al. 2014; Fattore et al. 2017). In many studies, this correlation was not found for only one single miRNA (Kanemaru et al. 2011) but for distinct miRNA profiles (Friedman et al. 2012). A recent study described a seven-miRNA panel (MELmiR-7) out of 17 miRNAs to correctly discriminate between melanoma patients of all disease stages and healthy controls with a sensitivity of 93% and a specificity of 82% (Stark et al. 2015). Moreover, the authors reported that this miRNA panel characterized the patients’ overall survival with higher accuracy than the serum markers LDH and S100B. Notably, the majority of miRNAs are not tumor-specific, but may also be expressed in inflammation, immune activation, and other conditions (Cortez et al. 2011). As another disadvantage, miRNAs in serum or plasma have to be quantified in relation to housekeeping miRNAs like U6, miR-451, or miR-16, which might be deregulated in cancer patients (Aftab et al. 2014). For translating the promising results of miRNA detection and quantification into clinical use in the routine care of melanoma patients, validation studies in large patient

of multiple cancer entities including melanoma (Calapre et al. 2017). Thus, it has been shown that the methylation of tumor suppressor genes like ras association domain family 1 isoform A (RASSF1A), retinoic acid receptor beta 2 (RARb2), or O-6-methylguanine-DNA methyltransferase (MGMT) is associated with an inferior survival and therapy outcome in melanoma patients (Hoon et al. 2004; Mori et al. 2005, 2006); see Fig. 4.

Circulating MicroRNA (miRNA) MicroRNAs (miRNAs) are short, noncoding RNA molecules which are functional in the regulation and modulation of gene transcription, posttranscription, and epigenetic expression. miRNAs are mainly actively secreted by their cells of origin and not passively released after cell apoptosis or necrosis like ctDNAs (Chen et al. 2012). Additionally, miRNAs in comparison to ctDNAs are relatively stable in the blood stream since they are commonly packed in vesicles or bound to proteins or lipoproteins (Vickers et al. 2011) and therefore are interesting biomarker candidates (Mitchell et al. 2008). Circulating miRNA expression in blood 1.0

Nonmethylated (n 33)

Probability of survival

0.8

(n

1 methylated 16)

0.6 0.4 0.2 0

p

0

0.01

10

20

30

40

50

Months

Fig. 4 Prognostic utility of circulating methylated DNA in blood of patients with melanoma. Kaplan-Meier survival curves of patients with stage IV melanoma undergoing biochemotherapy. Correlation of prebiochemotherapy serum methylation of at least one marker with overall survival (logrank test p = 0.01). Patients with serum methylation of only RAR-β2, methylation of only RASSF1A, or

methylation of at least one marker had significantly worse overall survival compared with patients who had no methylated markers (logrank test p = 0.010).  1 methylated marker, patients with serum methylation of at least one marker; nonmethylated, patients with no serum methylation of genes. (From Mori et al. 2005)

94

cohorts, preferably accompanying a clinical trial, are needed.

Circulating Immune Cells The frequency of immunologically active cells in the blood stream has been analyzed for an association with the course of disease and outcome of cancer patients for a long time. Recent advances in suitable technologies like multiparametric flow cytometry allow the quantification of multiple subsets of circulating immune cells in one single blood sample. In melanoma, immune cell frequencies are particularly appealing as biomarkers for immunotherapeutic strategies. Thus, it has been shown for patients treated with the antiCTLA-4 mAb ipilimumab that low peripheral blood baseline counts of monocytes and myeloid-derived suppressor cells (MDSCs), as well as high counts of eosinophils, lymphocytes, and regulatory T cells (Tregs), were associated with a favorable survival of the corresponding patients (Martens et al. 2016). For anti-PD-1 treatment with pembrolizumab, it has been shown that high numbers of eosinophils and lymphocytes are strongly correlated with a favorable treatment outcome and prolonged survival (Weide et al. 2016). Another recent study of T-cell subsets in the peripheral blood of melanoma patients before and after treatment with pembrolizumab revealed a change in the subset of exhausted CD8+ T cells, which in relation to the patients’ tumor burden was correlated with treatment response (Huang et al. 2017). Since the analysis and quantification of immune cell subsets are a straightforward methodology and results are quickly achieved, it might gain more attention in the near future as a useful biomarker in melanoma patients, particularly in association with immunotherapy.

Treatment-Associated Biomarkers As large bodies of evidence have been generated supporting the use of BRAF inhibitor-based therapy and immune checkpoint antibody therapy in metastatic melanoma, extensive retrospective

D. Schadendorf et al.

analyses have been undertaken to understand which patient subpopulations are more or less likely to derive benefit. One theme that has emerged from this research is that previously known prognostic factors have even more significance as predictors of clinical benefit from therapy. As this has borne out to be true for both targeted therapy and immunotherapy in melanoma, we will consider this evidence as it pertains to both types of therapy in this chapter. Molecular features that predict response or resistance to immune checkpoint antibody therapy are discussed in detail in the chapters dealing with this treatment modality. Therefore, molecular predictors of treatment outcome to BRAF inhibitorbased therapy will be further developed here.

Serum Lactate Dehydrogenase (LDH) Serum LDH has been long recognized as a prognostic factor in metastatic melanoma and a component of the AJCC staging system. In the 2009 AJCC analysis, patients with elevated serum LDH had a median overall survival of less than 1 year compared to approximately 2 years in the patients with normal serum LDH (Balch et al. 2009). And, whereas only 10% of patients with elevated serum LDH survive to 5 years, 25% of patients with normal LDH survived to that landmark. Notably, this analysis did not just for sites of metastatic disease, also known to have prognostic significance, with comparable differences in survival time when comparing patients with skin, subcutaneous, distant lymph node metastases versus those with visceral organ involvement beyond the lung. Considering baseline serum LDH values in relation to outcome on BRAF inhibitor-based targeted therapy and immune checkpoint antibody therapy, there is a similar magnitude of difference in intermediate and long-term outcomes. With long-term follow-up of the vemurafenib singleagent phase III trial, the median overall survival in the normal LDH subpopulation was 18.1 months compared to 9.6 months in those with elevated LDH (Chapman et al. 2017). And with mature follow-up data out to 4 years, 22.8% of patients with normal LDH were still alive

Biomarkers for Melanoma

versus 9.0% of patients with elevated LDH at baseline. For BRAF/MEK combination therapy, the largest available database exists for dabrafenib/ trametinib with regard to outcomes in patient subpopulations. Among 563 patients treated with dabrafenib/trametinib across to phase III trials, the 2-year progression-free survival rate was 39% for those with normal baseline serum LDH and 14% for those with elevated LDH (Long et al. 2016). The rates of overall survival at 2 years for these two groups were 66% and 27%, respectively. This difference was sustained with 3-year outcomes: 55% overall survival likelihood for normal LDH patients versus 22% in elevated LDH patients. In the vemurafenib/cobimetinib phase III trial, similarly disparate outcomes were seen based on serum LDH. Patients with normal serum LDH at baseline had a median progressionfree survival of 13.4 months versus 8.2 months in those with elevated LDH. Median overall survival in the normal LDH subgroup had not been defined as of the most recent, updated analysis but was clearly superior to the 14.8-month median overall survival in the elevated LDH subgroup. In the three-arm randomized trial of ipilimumab/nivolumab versus nivolumab versus ipilimumab, 3-year outcomes varied markedly by baseline serum LDH value (Wolchok et al. 2017). For combination ipilimumab/nivolumab therapy, the 3-year progression-free survival rate for patients with normal serum LDH was 45%, 28% for those with elevated serum LDH, and 17% for those with a baseline LDH that was two times the upper limit of normal. For nivolumab monotherapy, the rates of 3-year progression-free survival were 37%, 21%, and 11% for these same subgroups. And, for ipilimumab monotherapy, they were 14%, 3%, and 0%, respectively. For overall survival, ipilimumab/nivolumab combination therapy produced a 66% 3-year overall survival rate in normal LDH patients compared to 44% and 31% in those with elevated LDH or two times the upper limit of normal. Nivolumab monotherapy was associated with a 61% likelihood of survival at 3 years versus 34% in those with elevated LDH and 14% in the two times upper limit of normal subgroup. Lastly,

95

ipilimumab monotherapy yielded survival rates of 42%, 20%, and 7% across these three subgroups.

Burden of Metastatic Disease As is the case for prognostication in the metastatic melanoma population, various measures of extent of metastatic disease have prognostic value independent of serum LDH in the setting of both BRAF inhibitor-based therapy and immune checkpoint therapy. The first such published analysis came from the pooled phase 3 populations treated with dabrafenib/trametinib (Long et al. 2016). Patient demographic and disease characteristics were all entered into a classification and regression trees analysis which selects outcome discriminating features in an unbiased and hierarchical fashion. While serum LDH was the most powerful discriminator of both progression-free and overall survival outcomes, disease burden as described by aggregate size of measured lesions at baseline and number of involved organs/sites were the second and third and only additional discriminating feature for progression-free survival. And, these subgroups only had predictive value within the population with normal serum LDH. The metastatic site classification was not determined based on number of lesions within or across metastatic sites, but rather on the number of discrete tissue or organ sites involved. For example, the lung and liver were each considered as discrete sites of involvement, as were the skin and lymph nodes. Significant stratification in outcomes was observed for those patients with one or two sites of metastatic disease versus those with three or more. This analysis yielded a particularly striking difference in progression-free survival at 3 years, with 42% of those patients with only one or two sites of metastatic disease measuring less than 6.6 cm in aggregate remaining progressionfree versus 0% of the two times upper limit of normal serum LDH population. In a subgroup analysis of the three-arm randomized trial of ipilimumab/nivolumab versus nivolumab versus ipilimumab, two unique measures of disease burden were applied (Wolchok

96

et al. 2017). The entire patient population was divided into quartiles in one subgroup analysis, and, for the other, the number of metastatic sites was so grouped into one, two to three, and greater than three. There was a threefold difference in the aggregate measure of lesions at baseline in the highest quartile versus lowest quartile. The likelihood of 3-year progression-free survival for a nivolumab/nivolumab, nivolumab, and ipilimumab was 45%, 38%, and 15%, respectively, in the lowest tumor burden quartile. This compares to 33%, 29%, and 5% rates of 3-year PFS in the highest tumor burden quartile. More subtle differences were seen for 3-year PFS rate across subgroups defined by number of metastatic sites. Comparing the greater than three-site subgroup to the single-site subgroup, ipilimumab/ nivolumab was associated with a 47% 3-year PFS rate versus 27%, and nivolumab produced very similar rates of 37% and 36%, while ipilimumab yielded similar rates of 12% and 11%. Interestingly, more striking differences were noted in 3-year overall survival outcome across these disease site subgroups: 70% versus 42%, 65% versus 44%, and 48% versus 28% for ipilimumab/nivolumab, nivolumab, and implement, respectively.

Body Mass Index (BMI) A recent, striking analysis has found that elevated body mass index has positive predictive value in the setting of both BRAF inhibitor-based therapy and immune checkpoint antibody therapy (McQuade et al. 2018). Unlike serum LDH and various measures of metastatic disease burden, BMI had a statistically significant predictive value, but no prognostic significance. Very similar improvements in progression-free survival and overall survival were seen between targeted therapy and immunotherapy-treated populations. A 28% improvement in PFS outcomes was observed for obese patients compared to those with normal BMI and a 40% improvement in overall survival among those receiving BRAF/MEK combination therapy. Among those treated with immune checkpoint antibody therapy, 25% improvement

D. Schadendorf et al.

in PFS and 36% improvement in overall survival were seen. Nearly all of this apparently beneficial effect of obesity was observed in men, in whom there was a highly statistically significant 47% advantage in overall survival for men versus a nonsignificant 15% better outcome for obese women. Notably, there was no difference in outcome across the subpopulations when treated with chemotherapy. These findings have led to hypotheses regarding hormonal and metabolic effects on melanoma biology, therapeutic vulnerability, and resistance that are being pursued in ongoing translational studies.

Molecular Features Associated with Outcome on BRAF Inhibitor-Based Therapy Only preliminary data are available regarding cooccurring somatic alterations accompanying BRAF V600 mutations and their association with outcome on BRAF inhibitor-based therapy. Some studies have provided both preclinical and clinical evidence that MAP kinase pathway intrinsic components can confer relative resistance to BRAF inhibitor-based therapy. Among 124 BRAF V600-mutant melanoma patients treated with BRAF inhibitor monotherapy, 10% harbored coexisting P124L/Q/S substitutions in MEK1 (Carlino et al. 2015). The likelihood of response was substantially lower in these patients compared to the rest of the cohort (33% vs. 72% in MEK1P124Q/S vs. MEK1P124 wild-type, p = 0.018) as well as shorter PFS. An analysis of BRAF allele copy number among the cohort of 46 BRAF-mutant melanoma patients treated with MAP kinase pathway inhibitors demonstrated a beneficial association between elevated BRAF copy number and outcome (Stagni et al. 2018). This finding raises the hypothesis that melanoma with both activities mutations and copy number increases of BRAF is most “addicted” to the oncogenic function of this gene. While there are extensive preclinical data supporting the relevance of genetic alterations in PTEN, p53, Rb, and CDKN2A in relation to BRAF inhibitor sensitivity, systemic interrogation of clinical cohorts is still awaited.

Biomarkers for Melanoma

Multiple groups have identified markers of BRAF inhibitor resistance that align with a dedifferentiated, neural crest cell-like phenotype (Konieczkowski et al. 2014; Muller et al. 2014; Zuo et al. 2018). These studies leveraged the availability of large numbers of immortalized melanoma cell lines with variable sensitivity/ resistance to BRAF inhibitors to nominate transcription factors, receptor tyrosine kinases, and activated components of the PI3 kinase pathway as designators of the cell state that is intrinsically resistant to therapy. Fewer studies have included analysis of patient tumor samples but those that have found statistically significant differences in clinical outcome even among small number of samples. In one analysis of just 12 V600 mutant BRAF melanoma patients treated with BRAF/ MEK combination therapy, there was a threefold difference in median progression-free survival when comparing those tumors that had markers of dedifferentiation (low MITF, high AXL) compared to more differentiated tumors (high MITF, low AXL) (Konieczkowski et al. 2014). In another analysis of BRAF inhibitor-treated patients, supportive evidence of this low MITF expressing phenotype conferring resistance was demonstrated by showing the emergence of the low MITF melanoma cells at the time of clinical progression (Muller et al. 2014). But noting that numerous melanoma cell lines and untreated melanoma patient tumor samples feature the low MITF/high AXL state, this appears to be a cell state that can emerge during the evolution of melanoma before the application of therapy. Notably, some of these same molecular features (notably high AXL expression) have been associated with intrinsic resistance to PD-1 antibody therapy (Hugo et al. 2017). This raises the concerning possibility that a significant proportion of melanomas have adopted a cell state that is resistant to either therapeutic modality. A last line of evidence suggests that markers of immune recognition and activated effector T cells in sites of metastatic melanoma positively associate with outcomes on BRAF inhibitor therapy (Massi et al. 2017). Among 39 patients treated with BRAF inhibitor monotherapy and 25 patients treated with BRAF/MEK inhibitor

97

combination therapy, presence of CD8+ T cells strongly associated with likelihood of response and superior overall survival. Interestingly, co-expression of markers of beta-catenin pathway activation additionally informed BRAF inhibitor treatment outcome. Previously published data link beta-catenin pathway signaling and exclusion of T cell from the melanoma tumor microenvironment (Spranger et al. 2015). So, it would be anticipated that activity in this pathway would overlap with the CD8-negative subpopulation. Yet, those patients with the highest level of CD8 T-cell infiltration and lack of beta-catenin pathway activation had a 75% superior PFS and overall survival outcome that was significant even after adjusting for other disease characteristics known to impact likelihood of benefit from BRAF inhibitor therapy. Long known to be associated with favorable prognosis in both early and advanced melanoma, these preliminary findings suggest that greater degrees of immune recognition have even bigger impact on the outcome of patients treated with BRAF inhibitor-based therapy.

References Acland K, Evans AV, Abraha H, Healy CM, Roblin P, Calonje E, Orchard G, Higgins E, Sherwood R, Russell-Jones R (2002) Serum S100 concentrations are not useful in predicting micrometastatic disease in cutaneous malignant melanoma. Br J Dermatol 146:832–835 Aftab MN, Dinger ME, Perera RJ (2014) The role of microRNAs and long non-coding RNAs in the pathology, diagnosis, and management of melanoma. Arch Biochem Biophys 563:60–70 Alonso SR, Ortiz P, Pollan M, Perez-Gomez B, Sanchez L, Acuna MJ, Pajares R, Martinez-Tello FJ, Hortelano CM, Piris MA, Rodriguez-Peralto JL (2004) Progression in cutaneous malignant melanoma is associated with distinct expression profiles a tissue microarraybased study. Am J Pathol 164:193–203 Ansieau S, Bastid J, Doreau A, Morel AP, Bouchet BP, Thomas C, Fauvet F, Puisieux I, Doglioni C, Piccinin S, Maestro R, Voeltzel T, Selmi A, Valsesia-Wittmann S, Caron de Fromentel C, Puisieux A (2008) Induction of EMT by twist proteins as a collateral effect of tumorpromoting inactivation of premature senescence. Cancer Cell 14:79–89 Bailey MH, Tokheim C, Porta-Pardo E, Sengupta S, Bertrand D, Weerasinghe A, Colaprico A, Wendl MC, Kim J, Reardon B, Ng PK, Jeong KJ, Cao S, Wang Z, Gao J, Gao Q, Wang F, Liu EM, Mularoni L, Rubio-

98 Perez C, Nagarajan N, Cortes-Ciriano I, Zhou DC, Liang WW, Hess JM, Yellapantula VD, Tamborero D, Gonzalez-Perez A, Suphavilai C, Ko JY, Khurana E, Park PJ, Van Allen EM, Liang H, Lawrence MS, Godzik A, Lopez-Bigas N, Stuart J, Wheeler D, Getz G, Chen K, Lazar AJ, Mills GB, Karchin R, Ding L (2018) Comprehensive characterization of Cancer driver genes and mutations. Cell 173:371–85.e18 Balch CM, Buzaid AC, Soong SJ, Atkins MB, Cascinelli N, Coit DG, Fleming ID, Gershenwald JE, Houghton A Jr, Kirkwood JM, McMasters KM, Mihm MF, Morton DL, Reintgen DS, Ross MI, Sober A, Thompson JA, Thompson JF (2001a) Final version of the American joint committee on Cancer staging system for cutaneous melanoma. J Clin Oncol 19:3635–3648 Balch CM, Soong SJ, Gershenwald JE, Thompson JF, Reintgen DS, Cascinelli N, Urist M, McMasters KM, Ross MI, Kirkwood JM, Atkins MB, Thompson JA, Coit DG, Byrd D, Desmond R, Zhang Y, Liu PY, Lyman GH, Morabito A (2001b) Prognostic factors analysis of 17,600 melanoma patients: validation of the American joint committee on Cancer melanoma staging system. J Clin Oncol 19:3622–3634 Balch CM, Gershenwald JE, Soong SJ, Thompson JF, Atkins MB, Byrd DR, Buzaid AC, Cochran AJ, Coit DG, Ding S, Eggermont AM, Flaherty KT, Gimotty PA, Kirkwood JM, McMasters KM, Mihm MC Jr, Morton DL, Ross MI, Sober AJ, Sondak VK (2009) Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol 27:6199–6206 Balsamo M, Scordamaglia F, Pietra G, Manzini C, Cantoni C, Boitano M, Queirolo P, Vermi W, Facchetti F, Moretta A, Moretta L, Mingari MC, Vitale M (2009) Melanoma-associated fibroblasts modulate NK cell phenotype and antitumor cytotoxicity. Proc Natl Acad Sci USA 106:20847–20852 Bennett DC (2008) How to make a melanoma: what do we know of the primary clonal events? Pigment Cell Melanoma Res 21:27–38 Berghuis AM, Koffijberg H, Prakash J, Terstappen LW, IJzerman MJ (2017) Detecting blood-based biomarkers in metastatic breast Cancer: a systematic review of their current status and clinical Utility. Int J Mol Sci 18:363 Blesch A, Bosserhoff AK, Apfel R, Behl C, Hessdoerfer B, Schmitt A, Jachimczak P, Lottspeich F, Buettner R, Bogdahn U (1994) Cloning of a novel malignant melanoma-derived growth-regulatory protein, MIA. Cancer Res 54:5695 Bogdahn U, Apfel R, Hahn M, Gerlach M, Behl C, Hoppe J, Martin R (1989) Autocrine tumor cell growthinhibiting activities from human malignant melanoma. Cancer Res 49:5358–5363 Brown EJ, Frazier WA (2001) Integrin-associated protein (CD47) and its ligands. Trends Cell Biol 11:130–135 Brunner G, Reitz M, Heinecke A, Lippold A, Berking C, Suter L, Atzpodien J (2013) A nine-gene signature predicting clinical outcome in cutaneous melanoma. J Cancer Res Clin Oncol 139:249–258

D. Schadendorf et al. Buchbinder E, Hodi FS (2015) Cytotoxic T lymphocyte antigen-4 and immune checkpoint blockade. J Clin Invest 125:3377–3383 Calapre L, Warburton L, Millward M, Ziman M, Gray ES (2017) Circulating tumour DNA (ctDNA) as a liquid biopsy for melanoma. Cancer Lett 404:62–69 Campisi J, d’Adda di Fagagna F (2007) Cellular senescence: when bad things happen to good cells. Nat Rev Mol Cell Biol 8:729–740 Caramel J, Papadogeorgakis E, Hill L, Browne GJ, Richard G, Wierinckx A, Saldanha G, Osborne J, Hutchinson P, Tse G, Lachuer J, Puisieux A, Pringle JH, Ansieau S, Tulchinsky E (2013) A switch in the expression of embryonic EMT-inducers drives the development of malignant melanoma. Cancer Cell 24:466–480 Carlino MS, Fung C, Shahheydari H, Todd JR, Boyd SC, Irvine M, Nagrial AM, Scolyer RA, Kefford RF, Long GV, Rizos H (2015) Preexisting MEK1P124 mutations diminish response to BRAF inhibitors in metastatic melanoma patients. Clin Cancer Res 21:98–105 Chapman PB, Robert C, Larkin J, Haanen JB, Ribas A, Hogg D, Hamid O, Ascierto PA, Testori A, Lorigan PC, Dummer R, Sosman JA, Flaherty KT, Chang I, Coleman S, Caro I, Hauschild A, McArthur GA (2017) Vemurafenib in patients with BRAFV600 mutation-positive metastatic melanoma: final overall survival results of the randomized BRIM-3 study. Ann Oncol 28:2581–2587 Chen X, Liang H, Zhang J, Zen K, Zhang CY (2012) Secreted microRNAs: a new form of intercellular communication. Trends Cell Biol 22:125–132 Clark WH Jr, Elder DE, th Guerry D, Braitman LE, Trock BJ, Schultz D, Synnestvedt M, Halpern AC (1989) Model predicting survival in stage I melanoma based on tumor progression. J Natl Cancer Inst 81:1893–1904 Clarke LE, Warf MB, Flake DD 2nd, Hartman AR, Tahan S, Shea CR, Gerami P, Messina J, Florell SR, Wenstrup RJ, Rushton K, Roundy KM, Rock C, Roa B, Kolquist KA, Gutin A, Billings S, Leachman S (2015) Clinical validation of a gene expression signature that differentiates benign nevi from malignant melanoma. J Cutan Pathol 42:244–252 Clarke LE, Flake DD 2nd, Busam K, Cockerell C, Helm K, McNiff J, Reed J, Tschen J, Kim J, Barnhill R, Elenitsas R, Prieto VG, Nelson J, Kimbrell H, Kolquist KA, Brown KL, Warf MB, Roa BB, Wenstrup RJ (2017) An independent validation of a gene expression signature to differentiate malignant melanoma from benign melanocytic nevi. Cancer 123:617–628 Clemente CG, Mihm MC Jr, Bufalino R, Zurrida S, Collini P, Cascinelli N (1996) Prognostic value of tumor infiltrating lymphocytes in the vertical growth phase of primary cutaneous melanoma. Cancer 77:1303–1310 Cochran AJ, Wen DR, Huang RR, Wang HJ, Elashoff R, Morton DL (2004) Prediction of metastatic melanoma in nonsentinel nodes and clinical outcome based on the primary melanoma and the sentinel node. Mod Pathol 17:747–755

Biomarkers for Melanoma Corona R, Mele A, Amini M, De Rosa G, Coppola G, Piccardi P, Fucci M, Pasquini P, Faraggiana T (1996) Interobserver variability on the histopathologic diagnosis of cutaneous melanoma and other pigmented skin lesions. J Clin Oncol 14:1218–1223 Cortez MA, Bueso-Ramos C, Ferdin J, Lopez-Berestein G, Sood AK, Calin GA (2011) MicroRNAs in body fluids–the mix of hormones and biomarkers. Nat Rev Clin Oncol 8:467–477 Cyster JG (1999) Chemokines and cell migration in secondary lymphoid organs. Science 286:2098–2102 Daniotti M, Oggionni M, Ranzani T, Vallacchi V, Campi V, Di Stasi D, Torre GD, Perrone F, Luoni C, Suardi S, Frattini M, Pilotti S, Anichini A, Tragni G, Parmiani G, Pierotti MA, Rodolfo M (2004) BRAF alterations are associated with complex mutational profiles in malignant melanoma. Oncogene 23:5968–5977 De Souza LM, Robertson BM, Robertson GP (2017) Future of circulating tumor cells in the melanoma clinical and research laboratory settings. Cancer Lett 392:60–70 Deichmann M, Brenner A, Bock M, Jäckel A, Uhl K, Waldmann V, Näher H (1999) S100-beta, melanomainhibiting activity and lactate dehydrogenase discriminate progressive from non-progressive American joint committee on Cancer stage IV melanoma. J Clin Oncol 17(6):1891–6 Deichmann M, Kahle B, Moser K, Wacker J, Wust K (2004) Diagnosing melanoma patients entering American joint committee on Cancer stage IV, C-reactive protein in serum is superior to lactate dehydrogenase. Br J Cancer 91:699–702 Dror S, Sander L, Schwartz H, Sheinboim D, Barzilai A, Dishon Y, Apcher S, Golan T, Greenberger S, Barshack I, Malcov H, Zilberberg A, Levin L, Nessling M, Friedmann Y, Igras V, Barzilay O, Vaknine H, Brenner R, Zinger A, Schroeder A, Gonen P, Khaled M, Erez N, Hoheisel JD, Levy C (2016) Melanoma miRNA trafficking controls tumour primary niche formation. Nat Cell Biol 18:1006–1017 Duffy MJ (2013) Tumor markers in clinical practice: a review focusing on common solid cancers. Med Princ Pract 22:4–11 Duffy MJ, Crown J (2008) A personalized approach to cancer treatment: how biomarkers can help. Clin Chem 54:1770–1779 Duffy MJ, O’Donovan N, Crown J (2011) Use of molecular markers for predicting therapy response in cancer patients. Cancer Treat Rev 37:151–159 Egberts F, Momkvist A, Egberts JH, Kaehler KC, Hauschild A (2010) Serum S100B and LDH are not useful in predicting the sentinel node status in melanoma patients. Anticancer Res 30:1799–1805 Egberts F, Kotthoff EM, Gerdes S, Egberts JH, Weichenthal M, Hauschild A (2012) Comparative study of YKL-40, S-100B and LDH as monitoring tools for stage IV melanoma. Eur J Cancer 48:695–702 Eisenstein A, Gonzalez EC, Raghunathan R, Xu X, Wu M, McLean EO, McGee J, Ryu B, Alani RM (2018)

99 Emerging biomarkers in cutaneous melanoma. Mol Diagn Ther 22:203–218 English K (2013) Mechanisms of mesenchymal stromal cell immunomodulation. Immunol Cell Biol 91:19–26 Farmer ER, Gonin R, Hanna MP (1996) Discordance in the histopathologic diagnosis of melanoma and melanocytic nevi between expert pathologists. Hum Pathol 27:528–531 Fattore L, Costantini S, Malpicci D, Ruggiero CF, Ascierto PA, Croce CM, Mancini R, Ciliberto G (2017) MicroRNAs in melanoma development and resistance to target therapy. Oncotarget 8:22262–22278 Findeisen P, Zapatka M, Peccerella T, Matzk H, Neumaier M, Schadendorf D, Ugurel S (2009) Serum amyloid a as a prognostic marker in melanoma identified by proteomic profiling. J Clin Oncol 23:2199–2208 Friedman EB, Shang S, de Miera EV, Fog JU, Teilum MW, Ma MW, Berman RS, Shapiro RL, Pavlick AC, Hernando E, Baker A, Shao Y, Osman I (2012) Serum microRNAs as biomarkers for recurrence in melanoma. J Transl Med 10:155 Fu W, Li J, Zhang W, Li P (2017) High expression of CD47 predicts adverse prognosis in Chinese patients and suppresses immune response in melanoma. Biomed Pharmacother 93:1190–1196 Fuentes-Arderiu X (2013) What is a biomarker? It’s time for a renewed definition. Clin Chem Lab Med 51:1689–1690 Fuzery AK, Levin J, Chan MM, Chan DW (2013) Translation of proteomic biomarkers into FDA approved cancer diagnostics: issues and challenges. Clin Proteomics 10:13 Gaiser T, Kutzner H, Palmedo G, Siegelin MD, Wiesner T, Bruckner T, Hartschuh W, Enk AH, Becker MR (2010) Classifying ambiguous melanocytic lesions with FISH and correlation with clinical long-term follow up. Mod Pathol 23:413–419 Garrido-Ruiz MC, Requena L, Ortiz P, Perez-Gomez B, Alonso SR, Peralto JL (2010) The immunohistochemical profile of Spitz nevi and conventional (nonSpitzoid) melanomas: a baseline study. Mod Pathol 23:1215–1224 George E, Polissar NL, Wick M (2010) Immunohistochemical evaluation of p16INK4A, E-cadherin, and cyclin D1 expression in melanoma and Spitz tumors. Am J Clin Pathol 133:370–379 Gerami P, Jewell SS, Morrison LE, Blondin B, Schulz J, Ruffalo T, th Matushek P, Legator M, Jacobson K, Dalton SR, Charzan S, Kolaitis NA, Guitart J, Lertsbarapa T, Boone S, LeBoit PE, Bastian BC (2009) Fluorescence in situ hybridization (FISH) as an ancillary diagnostic tool in the diagnosis of melanoma. Am J Surg Pathol 33:1146–1156 Gerami P, Li G, Pouryazdanparast P, Blondin B, Beilfuss B, Slenk C, Du J, Guitart J, Jewell S, Pestova K (2012) A highly specific and discriminatory FISH assay for distinguishing between benign and malignant melanocytic neoplasms. Am J Surg Pathol 36:808–817

100 Gerami P, Cook RW, Russell MC, Wilkinson J, Amaria RN, Gonzalez R, Lyle S, Jackson GL, Greisinger AJ, Johnson CE, Oelschlager KM, Stone JF, Maetzold DJ, Ferris LK, Wayne JD, Cooper C, Obregon R, Delman KA, Lawson D (2015a) Gene expression profiling for molecular staging of cutaneous melanoma in patients undergoing sentinel lymph node biopsy. J Am Acad Dermatol 72:780–5.e3 Gerami P, Cook RW, Wilkinson J, Russell MC, Dhillon N, Amaria RN, Gonzalez R, Lyle S, Johnson CE, Oelschlager KM, Jackson GL, Greisinger AJ, Maetzold D, Delman KA, Lawson DH, Stone JF (2015b) Development of a prognostic genetic signature to predict the metastatic risk associated with cutaneous melanoma. Clin Cancer Res 21:175–183 Gerami P, Yao Z, Polsky D, Jansen B, Busam K, Ho J, Martini M, Ferris LK (2017) Development and validation of a noninvasive 2-gene molecular assay for cutaneous melanoma. J Am Acad Dermatol 76:114–20.e2 Gershenwald JE, Scolyer RA, Hess KR, Sondak VK, Long GV, Ross MI, Lazar AJ, Faries MB, Kirkwood JM, McArthur GA, Haydu LE, Eggermont AMM, Flaherty KT, Balch CM, Thompson JF (2017) Melanoma staging: evidence-based changes in the American joint committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67:472–492 Goossens N, Nakagawa S, Sun X, Hoshida Y (2015) Cancer biomarker discovery and validation. Transl Cancer Res 4:256–269 Gould Rothberg BE, Rimm DL (2010) Biomarkers: the useful and the not so useful–an assessment of molecular prognostic markers for cutaneous melanoma. J Invest Dermatol 130:1971–1987 Gould Rothberg BE, Berger AJ, Molinaro AM, Subtil A, Krauthammer MO, Camp RL, Bradley WR, Ariyan S, Kluger HM, Rimm DL (2009a) Melanoma prognostic model using tissue microarrays and genetic algorithms. J Clin Oncol 27:5772–5780 Gould Rothberg BE, Bracken MB, Rimm DL (2009b) Tissue biomarkers for prognosis in cutaneous melanoma: a systematic review and meta-analysis. J Natl Cancer Inst 101:452–474 Gray ES, Rizos H, Reid AL, Boyd SC, Pereira MR, Lo J, Tembe V, Freeman J, Lee JH, Scolyer RA, Siew K, Lomma C, Cooper A, Khattak MA, Meniawy TM, Long GV, Carlino MS, Millward M, Ziman M (2015) Circulating tumor DNA to monitor treatment response and detect acquired resistance in patients with metastatic melanoma. Oncotarget 6: 42008–42018 Greenberg ES, Chong KK, Huynh KT, Tanaka R, Hoon DS (2014) Epigenetic biomarkers in skin cancer. Cancer Lett 342:170–177 Group, Biomarkers Definitions Working (2001) Biomarkers and surrogate endpoints: preferred definitions and conceptual framework. Clin Pharmacol Ther 69:89–95 Guo HB, Stoffel-Wagner B, Bierwirth T, Mezger J, Klingmuller D (1995) Clinical significance of serum

D. Schadendorf et al. S100 in metastatic malignant melanoma. Eur J Cancer 31A:924–928 Hauschild A, Engel G, Brenner W, Glaeser R, Moenig H, Henze E, Christophers E (1999a) Predictive value of serum S100B for monitoring patients with metastatic melanoma during chemotherapy and/or immunotherapy. Br J Dermatol 140:1065–1071 Hauschild A, Engel G, Brenner W, Glaeser R, Monig H, Henze E, Christophers E (1999b) S100B protein detection in serum is a significant prognostic factor in metastatic melanoma. Oncology 56:338–344 Holtzhausen A, Zhao F, Evans KS, Tsutsui M, Orabona C, Tyler DS, Hanks BA (2015) Melanoma-derived Wnt5a promotes local dendritic-cell expression of IDO and Immunotolerance: opportunities for pharmacologic enhancement of immunotherapy. Cancer Immunol Res 3:1082–1095 Hoon DS, Spugnardi M, Kuo C, Huang SK, Morton DL, Taback B (2004) Profiling epigenetic inactivation of tumor suppressor genes in tumors and plasma from cutaneous melanoma patients. Oncogene 23: 4014–4022 Huang SK, Hoon DS (2016) Liquid biopsy utility for the surveillance of cutaneous malignant melanoma patients. Mol Oncol 10:450–463 Huang AC, Postow MA, Orlowski RJ, Mick R, Bengsch B, Manne S, Xu W, Harmon S, Giles JR, Wenz B, Adamow M, Kuk D, Panageas KS, Carrera C, Wong P, Quagliarello F, Wubbenhorst B, D’Andrea K, Pauken KE, Herati RS, Staupe RP, Schenkel JM, McGettigan S, Kothari S, George SM, Vonderheide RH, Amaravadi RK, Karakousis GC, Schuchter LM, Xu X, Nathanson KL, Wolchok JD, Gangadhar TC, Wherry EJ (2017) T-cell invigoration to tumour burden ratio associated with antiPD-1 response. Nature 545:60–65 Hugo W, Zaretsky JM, Sun L, Song C, Moreno BH, HuLieskovan S, Berent-Maoz B, Pang J, Chmielowski B, Cherry G, Seja E, Lomeli S, Kong X, Kelley MC, Sosman JA, Johnson DB, Ribas A, Lo RS (2017) Genomic and transcriptomic features of response to anti-PD-1 therapy in metastatic melanoma. Cell 168:542 Isenberg JS, Ridnour LA, Dimitry J, Frazier WA, Wink DA, Roberts DD (2006) CD47 is necessary for inhibition of nitric oxide-stimulated vascular cell responses by thrombospondin-10 . J Biol Chem 281:26069–26080 Jacobs JF, Nierkens S, Figdor CG, de Vries IJ, Adema GJ (2012) Regulatory T cells in melanoma: the final hurdle towards effective immunotherapy? Lancet Oncol 13: e32–e42 Jaiswal S, Jamieson CH, Pang WW, Park CY, Chao MP, Majeti R, Traver D, van Rooijen N, Weissman IL (2009) CD47 is upregulated on circulating hematopoietic stem cells and leukemia cells to avoid phagocytosis. Cell 138:271–285 Kanemaru H, Fukushima S, Yamashita J, Honda N, Oyama R, Kakimoto A, Masuguchi S, Ishihara T, Inoue Y, Jinnin M, Ihn H (2011) The circulating microRNA-

Biomarkers for Melanoma 221 level in patients with malignant melanoma as a new tumor marker. J Dermatol Sci 61:187–193 Kashani-Sabet M, Venna S, Rangel J et al (2009) A multimarker prognostic assay for melanoma. Clin Cancer Res 15:6987–6992 Kashani-Sabet M, Nosrati M, Miller JR 3rd, Sagebiel RW, Leong SPL, Lesniak A, Tong S, Lee SJ, Kirkwood JM (2017) Prospective validation of molecular prognostic markers in cutaneous melanoma: a correlative analysis of E1690. Clin Cancer Res 23:6888–6892 Kaur S, Roberts DD (2016) Divergent modulation of normal and neoplastic stem cells by thrombospondin-1 and CD47 signaling. Int J Biochem Cell Biol 81:184–194 Kawasaki K, Kawakami T, Watabe H, Itoh F, Mizoguchi M, Soma Y (2007) Expression of matrilysin (matrix metalloproteinase-7) in primary cutaneous and metastatic melanoma. Br J Dermatol 156:613–619 Khoja L, Lorigan P, Zhou C, Lancashire M, Booth J, Cummings J, Califano R, Clack G, Hughes A, Dive C (2013) Biomarker utility of circulating tumor cells in metastatic cutaneous melanoma. J Invest Dermatol 133:1582–1590 Khoja L, Shenjere P, Hodgson C, Hodgetts J, Clack G, Hughes A, Lorigan P, Dive C (2014) Prevalence and heterogeneity of circulating tumour cells in metastatic cutaneous melanoma. Melanoma Res 24:40–46 Khoja L, Lorigan P, Dive C, Keilholz U, Fusi A (2015) Circulating tumour cells as tumour biomarkers in melanoma: detection methods and clinical relevance. Ann Oncol 26:33–39 Konieczkowski DJ, Johannessen CM, Abudayyeh O, Kim JW, Cooper ZA, Piris A, Frederick DT, Barzily-Rokni M, Straussman R, Haq R, Fisher DE, Mesirov JP, Hahn WC, Flaherty KT, Wargo JA, Tamayo P, Garraway LA (2014) A melanoma cell state distinction influences sensitivity to MAPK pathway inhibitors. Cancer Discov 4:816–827 Koyanagi K, O’Day SJ, Gonzalez R et al (2005) Serial monitoring of circulating melanoma cells during neoadjuvant bio-chemotherapy for stage III melanoma: outcome prediction in a multicenter trial. J Clin Oncol 23:8057 Krähn G, Kaskel P, Sander S, Waizenhofer PJ, Wortmann S, Leiter U, Peter RU (2001) S100 beta is a more reliable tumor marker in peripheral blood for patients with newly occurred melanoma metastases compared with MIA, albumin and lactate-dehydrogenase. Anticancer Res 21:1311–1316 Krogh M, Christensen I, Bouwhuis M, Johansen JS, Norgaard P, Schmidt H, Hansson J, Suciu S, Eggermont AM, Bastholt L (2016) Prognostic and predictive value of YKL-40 in stage IIB-III melanoma. Melanoma Res 26:367–376 Kulasingam V, Diamandis EP (2008) Strategies for discovering novel cancer biomarkers through utilization of emerging technologies. Nat Clin Pract Oncol 5: 588–599 Lai CY, Schwartz BE, Hsu MY (2012) CD133+ melanoma subpopulations contribute to perivascular niche

101 morphogenesis and tumorigenicity through vasculogenic mimicry. Cancer Res 72:5111–5118 Lee JH, Chen Y, Chan JL, Qian YW, Goydos JS (2011) Molecular analysis of melanoma-induced sentinel lymph node immune dysfunction. Cancer Immunol Immunother 60:685–692 Lee JH, Long GV, Boyd S, Lo S, Menzies AM, Tembe V, Guminski A, Jakrot V, Scolyer RA, Mann GJ, Kefford RF, Carlino MS, Rizos H (2017) Circulating tumour DNA predicts response to anti-PD1 antibodies in metastatic melanoma. Ann Oncol 28:1130–1136 Li JP, Lu L, Wang LJ, Zhang FR, Shen WF (2011) Increased serum levels of S100B are related to the severity of cardiac dysfunction, renal insufficiency and major cardiac events in patients with chronic heart failure. Clin Biochem 44:984–988 Li FZ, Dhillon AS, Anderson RL, McArthur G, Ferrao PT (2015) Phenotype switching in melanoma: implications for progression and therapy. Front Oncol 5:31 Lian CG, Xu Y, Ceol C, Wu F, Larson A, Dresser K, Xu W, Tan L, Hu Y, Zhan Q, Lee CW, Hu D, Lian BQ, Kleffel S, Yang Y, Neiswender J, Khorasani AJ, Fang R, Lezcano C, Duncan LM, Scolyer RA, Thompson JF, Kakavand H, Houvras Y, Zon LI, Mihm MC Jr, Kaiser UB, Schatton T, Woda BA, Murphy GF, Shi YG (2012) Loss of 5-hydroxymethylcytosine is an epigenetic hallmark of melanoma. Cell 150:1135–1146 Lim SY, Lee JH, Diefenbach RJ, Kefford RF, Rizos H (2018) Liquid biomarkers in melanoma: detection and discovery. Mol Cancer 17:8 Liu J, Lichtenberg T, Hoadley KA, Poisson LM, Lazar AJ, Cherniack AD, Kovatich AJ, Benz CC, Levine DA, Lee AV, Omberg L, Wolf DM, Shriver CD, Thorsson V, Hu H (2018) An integrated TCGA pan-Cancer clinical data resource to drive high-quality survival outcome analytics. Cell 173:400–16.e11 Long GV, Grob JJ, Nathan P, Ribas A, Robert C, Schadendorf D, Lane SR, Mak C, Legenne P, Flaherty KT, Davies MA (2016) Factors predictive of response, disease progression, and overall survival after dabrafenib and trametinib combination treatment: a pooled analysis of individual patient data from randomised trials. Lancet Oncol 17:1743–1754 Mandala M, Massi D (2014) Tissue prognostic biomarkers in primary cutaneous melanoma. Virchows Arch 464:265–281 Manola J, Atkins M, Ibrahim J, Kirkwood J (2000) Prognostic factors in metastatic melanoma: a pooled analysis of eastern cooperative oncology group trials. J Clin Oncol 18:3782–3793 Martens A, Wistuba-Hamprecht K, Geukes Foppen M, Yuan J, Postow MA, Wong P, Romano E, Khammari A, Dreno B, Capone M, Ascierto PA, Di Giacomo AM, Maio M, Schilling B, Sucker A, Schadendorf D, Hassel JC, Eigentler TK, Martus P, Wolchok JD, Blank C, Pawelec G, Garbe C, Weide B (2016) Baseline peripheral blood biomarkers associated with clinical outcome of advanced melanoma patients treated with Ipilimumab. Clin Cancer Res 22:2908–2918

102 Massi D, Romano E, Rulli E, Merelli B, Nassini R, De Logu F, Bieche I, Baroni G, Cattaneo L, Xue G, Mandala M (2017) Baseline beta-catenin, programmed death-ligand 1 expression and tumour-infiltrating lymphocytes predict response and poor prognosis in BRAF inhibitor-treated melanoma patients. Eur J Cancer 78:70–81 Matlung HL, Szilagyi K, Barclay NA, van den Berg TK (2017) The CD47-SIRPalpha signaling axis as an innate immune checkpoint in cancer. Immunol Rev 276:145–164 McEvoy AC, Calapre L, Pereira MR, Giardina T, Robinson C, Khattak MA, Meniawy TM, Pritchard AL, Hayward NK, Amanuel B, Millward M, Ziman M, Gray ES (2017) Sensitive droplet digital PCR method for detection of TERT promoter mutations in cell free DNA from patients with metastatic melanoma. Oncotarget 8:78890–78900 McQuade JL, Daniel CR, Hess KR, Mak C, Wang DY, Rai RR, Park JJ, Haydu LE, Spencer C, Wongchenko M, Lane S, Lee DY, Kaper M, McKean M, Beckermann KE, Rubinstein SM, Rooney I, Musib L, Budha N, Hsu J, Nowicki TS, Avila A, Haas T, Puligandla M, Lee S, Fang S, Wargo JA, Gershenwald JE, Lee JE, Hwu P, Chapman PB, Sosman JA, Schadendorf D, Grob JJ, Flaherty KT, Walker D, Yan Y, McKenna E, Legos JJ, Carlino MS, Ribas A, Kirkwood JM, Long GV, Johnson DB, Menzies AM, Davies MA (2018) Association of body-mass index and outcomes in patients with metastatic melanoma treated with targeted therapy, immunotherapy, or chemotherapy: a retrospective, multicohort analysis. Lancet Oncol 19:310–322 McShane LM, Altman DG, Sauerbrei W, Taube SE, Gion M, Clark GM, Statistics Subcommittee of the NCIEORTC Working Group on Cancer Diagnostics (2005a) Reporting recommendations for tumor marker prognostic studies. J Clin Oncol 23:9067–9072 McShane LM, Altman DG, Sauerbrei W, Taube SE, Gion M, Clark GM (2005b) REporting recommendations for tumour MARKer prognostic studies (REMARK). Br J Cancer 93(4):387–391 Miller AJ, Mihm MC Jr (2006) Melanoma. N Engl J Med 355:51–65 Minca EC, Al-Rohil RN, Wang M, Harms PW, Ko JS, Collie AM, Kovalyshyn I, Prieto VG, Tetzlaff MT, Billings SD, Andea AA (2016) Comparison between melanoma gene expression score and fluorescence in situ hybridization for the classification of melanocytic lesions. Mod Pathol 29:832–843 Mitchell PS, Parkin RK, Kroh EM, Fritz BR, Wyman SK, Pogosova-Agadjanyan EL, Peterson A, Noteboom J, O’Briant KC, Allen A, Lin DW, Urban N, Drescher CW, Knudsen BS, Stirewalt DL, Gentleman R, Vessella RL, Nelson PS, Martin DB, Tewari M (2008) Circulating microRNAs as stable blood-based markers for cancer detection. Proc Natl Acad Sci USA 105: 10513–10518 Mocellin S, Del Fiore P, Guarnieri L, Scalerta R, Foletto M, Chiarion V, Pilati P, Nitti D, Lise M, Rossi CR (2004)

D. Schadendorf et al. Molecular detection of circulating tumor cells is an independent prognostic factor in patients with highrisk cutaneous melanoma. Int J Cancer 111:741–745 Mocellin S, Hoon D, Ambrosi A, Nitti D, Rossi CR (2006) The prognostic value of circulating tumor cells in patients with melanoma: a systematic review and meta-analysis. Clin Cancer Res 12:4605–4613 Mocellin S, Zavagno G, Nitti D (2008) The prognostic value of serum S100B in patients with cutaneous melanoma: a meta-analysis. Int J Cancer 123:2370–2376 Mordente A, Meucci E, Martorana GE, Silvestrini A (2015) Cancer biomarkers discovery and validation: state of the art, problems and future perspectives. Adv Exp Med Biol 867:9–26 Mori T, O’Day SJ, Umetani N, Martinez SR, Kitago M, Koyanagi K, Kuo C, Takeshima TL, Milford R, Wang HJ, Vu VD, Nguyen SL, Hoon DS (2005) Predictive utility of circulating methylated DNA in serum of melanoma patients receiving biochemotherapy. J Clin Oncol 23:9351–9358 Mori T, Martinez SR, O’Day SJ, Morton DL, Umetani N, Kitago M, Tanemura A, Nguyen SL, Tran AN, Wang HJ, Hoon DS (2006) Estrogen receptor-alpha methylation predicts melanoma progression. Cancer Res 66:6692–6698 Muller J, Krijgsman O, Tsoi J, Robert L, Hugo W, Song C, Kong X, Possik PA, Cornelissen-Steijger PD, Geukes Foppen MH, Kemper K, Goding CR, McDermott U, Blank C, Haanen J, Graeber TG, Ribas A, Lo RS, Peeper DS (2014) Low MITF/AXL ratio predicts early resistance to multiple targeted drugs in melanoma. Nat Commun 5:5712 Neagu M, Constantin C, Longo C (2015) Chemokines in the melanoma metastasis biomarkers portrait. J Immunoassay Immunochem 36:559–566 Network, The Cancer Genome Atlas (2015) Genomic classification of cutaneous melanoma. Cell 161:1681–1696 Newman AM, Bratman SV, J. To, Wynne JF, Eclov NC, Modlin LA, Liu CL, Neal JW, Wakelee HA, Merritt RE, Shrager JB, Loo BW Jr, Alizadeh AA, Diehn M (2014) An ultrasensitive method for quantitating circulating tumor DNA with broad patient coverage. Nat Med 20:548–554 Nezos A, Msaouel P, Pissimissis N, Lembessis P, Sourla A, Armakolas A, Gogas H, Stratigos AJ, Katsambas AD, Koutsilieris M (2011) Methods of detection of circulating melanoma cells: a comparative overview. Cancer Treat Rev 37:284–290 Paschen A, Sucker A, Hill B, Moll I, Zapatka M, Nguyen XD, Sim GC, Gutmann I, Hassel J, Becker JC, Steinle A, Schadendorf D, Ugurel S (2009) Differential clinical significance of individual NKG2D ligands in melanoma: soluble ULBP2 as an indicator of poor prognosis superior to S100B. Clin Cancer Res 15:5208–5215 Pepe MS, Etzioni R, Feng Z, Potter JD, Thompson ML, Thornquist M, Winget M, Yasui Y (2001) Phases of biomarker development for early detection of cancer. J Natl Cancer Inst 93:1054–1061

Biomarkers for Melanoma Persson L, Hardemark HG, Gustafsson J, Rundstrom G, Mendel-Hartvig I, Esscher T, Pahlman S (1987) S-100 protein and neuron-specific enolase in cerebrospinal fluid and serum: markers of cell damage in human central nervous system. Stroke 18:911–918 Pietra G, Manzini C, Rivara S, Vitale M, Cantoni C, Petretto A, Balsamo M, Conte R, Benelli R, Minghelli S, Solari N, Gualco M, Queirolo P, Moretta L, Mingari MC (2012) Melanoma cells inhibit natural killer cell function by modulating the expression of activating receptors and cytolytic activity. Cancer Res 72:1407–1415 Poste G (2011) Bring on the biomarkers. Nature 469:156–157 Rebmann V, Ugurel S, Tilgen W, Reinhold U, GrosseWilde H (2002) Soluble HLA-DR is a potent predictive indicator of disease progression in serum from earlystage melanoma patients. Int J Cancer 100:580–585 Rothhammer T, Bosserhoff AK (2007) Epigenetic events in malignant melanoma. Pigment Cell Res 20:92–111 Sanchez-Vega F, Mina M, Armenia J, Chatila WK, Luna A, La KC, Dimitriadoy S, Liu DL, Kantheti HS, Saghafinia S, Chakravarty D, Daian F, Gao Q, Bailey MH, Liang WW, Foltz SM, Shmulevich I, Ding L, Heins Z, Ochoa A, Gross B, Gao J, Zhang H, Kundra R, Kandoth C, Bahceci I, Dervishi L, Dogrusoz U, Zhou W, Shen H, Laird PW, Way GP, Greene CS, Liang H, Xiao Y, Wang C, Iavarone A, Berger AH, Bivona TG, Lazar AJ, Hammer GD, Giordano T, Kwong LN, McArthur G, Huang C, Tward AD, Frederick MJ, McCormick F, Meyerson M, Van Allen EM, Cherniack AD, Ciriello G, Sander C, Schultz N (2018) Oncogenic signaling pathways in the Cancer genome atlas. Cell 173:321–37.e10 Sanmamed MF, Carranza-Rua O, Alfaro C, Onate C, Martin-Algarra S, Perez G, Landazuri SF, Gonzalez A, Gross S, Rodriguez I, Munoz-Calleja C, RodriguezRuiz M, Sangro B, Lopez-Picazo JM, Rizzo M, Mazzolini G, Pascual JI, Andueza MP, Perez-Gracia JL, Melero I (2014) Serum interleukin-8 reflects tumor burden and treatment response across malignancies of multiple tissue origins. Clin Cancer Res 20:5697–5707 Sargent DJ, Conley BA, Allegra C, Collette L (2005) Clinical trial designs for predictive marker validation in cancer treatment trials. J Clin Oncol 23:2020–2027 Sarvaiya PJ, Guo D, Ulasov I, Gabikian P, Lesniak MS (2013) Chemokines in tumor progression and metastasis. Oncotarget 4:2171–2185 Schatton T, Murphy GF, Frank NY, Yamaura K, WaagaGasser AM, Gasser M, Zhan Q, Jordan S, Duncan LM, Weishaupt C, Fuhlbrigge RC, Kupper TS, Sayegh MH, Frank MH (2008) Identification of cells initiating human melanomas. Nature 451:345–349 Schatton T, Schutte U, Frank NY, Zhan Q, Hoerning A, Robles SC, Zhou J, Hodi FS, Spagnoli GC, Murphy GF, Frank MH (2010) Modulation of T-cell activation by malignant melanoma initiating cells. Cancer Res 70:697–708

103 Schinke C, Mo Y, Yu Y, Amiri K, Sosman J, Greally J, Verma A (2010) Aberrant DNA methylation in malignant melanoma. Melanoma Res 20:253–265 Schmidt H, Johansen JS, Sjoegren P, Christensen IJ, Sorensen BS, Fode K, Larsen J, von der Maase H (2006) Serum YKL-40 predicts relapse-free and overall survival in patients with American joint committee on Cancer stage I and II melanoma. J Clin Oncol 24:798–804 Schultz ES, Diepgen TL, Von Den Driesch P (1998) Clinical and prognostic relevance of serum S-100 beta protein in malignant melanoma. Br J Dermatol 138:426–430 Schwarzenbach H, Hoon DS, Pantel K (2011) Cell-free nucleic acids as biomarkers in cancer patients. Nat Rev Cancer 11:426–437 Scolyer RA, Shaw HM, Thompson JF, Li LX, Colman MH, Lo SK, McCarthy SW, Palmer AA, Nicoll KD, Dutta B, Slobedman E, Watson GF, Stretch JR (2003) Interobserver reproducibility of histopathologic prognostic variables in primary cutaneous melanomas. Am J Surg Pathol 27:1571–1576 Shoo BA, Sagebiel RW, Kashani-Sabet M (2010) Discordance in the histopathologic diagnosis of melanoma at a melanoma referral center. J Am Acad Dermatol 62:751–756 Sivan S, Suzan F, Rona O, Tamar H, Vivian B, Tamar P, Jacob S, Gal M, Michal L (2012) Serum CEACAM1 correlates with disease progression and survival in malignant melanoma patients. Clin Dev Immunol 2012:290536 Sorber L, Zwaenepoel K, Deschoolmeester V, Roeyen G, Lardon F, Rolfo C, Pauwels P (2017) A comparison of cell-free DNA isolation kits: isolation and quantification of cell-free DNA in plasma. J Mol Diagn 19:162–168 Speeckaert R, Vermaelen K, van Geel N, Autier P, Lambert J, Haspeslagh M, van Gele M, Thielemans K, Neyns B, Roche N, Verbeke N, Deron P, Speeckaert M, Brochez L (2012) Indoleamine 2,3-dioxygenase, a new prognostic marker in sentinel lymph nodes of melanoma patients. Eur J Cancer 48:2004–2011 Spranger S, Bao R, Gajewski TF (2015) Melanoma-intrinsic beta-catenin signalling prevents anti-tumour immunity. Nature 523:231–235 Stagni C, Zamuner C, Elefanti L, Zanin T, Bianco PD, Sommariva A, Fabozzi A, Pigozzo J, Mocellin S, Montesco MC, Chiarion-Sileni V, De Nicolo A, Menin C (2018) BRAF gene copy number and mutant allele frequency correlate with time to progression in metastatic melanoma patients treated with MAPK inhibitors. Mol Cancer Ther 17: 1332–1340 Stark MS, Klein K, Weide B, Haydu LE, Pflugfelder A, Tang YH, Palmer JM, Whiteman DC, Scolyer RA, Mann GJ, Thompson JF, Long GV, Barbour AP, Soyer HP, Garbe C, Herington A, Pollock PM, Hayward NK (2015) The prognostic and predictive value of melanoma-related MicroRNAs using tissue and serum:

104 a MicroRNA expression analysis. EBioMedicine 2:671–680 Tanemura A, Terando AM, Sim MS, van Hoesel AQ, de Maat MF, Morton DL, Hoon DS (2009) CpG island methylator phenotype predicts progression of malignant melanoma. Clin Cancer Res 15:1801–1807 Ugurel S (2005) Serological markers of melanoma. Hautarzt 2:173 Ugurel S, Rappl G, Tilgen W, Reinhold U (2001a) Increased serum concentration of angiogenic factors in malignant melanoma patients correlates with tumor progression and survival. J Clin Oncol 19:577–583 Ugurel S, Rebmann V, Ferrone S, Tilgen W, Grosse-Wilde H, Reinhold U (2001b) Soluble human leukocyte antigen-G serum level is elevated in melanoma patients and is further increased by interferon-alpha immunotherapy. Cancer 92:369–376 Vaquero J, Jordano Q, Lee WM, Blei AT (2003) Serum protein S-100b in acute liver failure: results of the US acute liver failure study group. Liver Transpl 9:887–888 Vickers KC, Palmisano BT, Shoucri BM, Shamburek RD, Remaley AT (2011) MicroRNAs are transported in plasma and delivered to recipient cells by high-density lipoproteins. Nat Cell Biol 13:423–433 Wang H, Lee S, Nigro CL, Lattanzio L, Merlano M, Monteverde M, Matin R, Purdie K, Mladkova N, Bergamaschi D, Harwood C, Syed N, Szlosarek P, Briasoulis E, McHugh A, Thompson A, Evans A, Leigh I, Fleming C, Inman GJ, Hatzimichael E, Proby C, Crook T (2012) NT5E (CD73) is epigenetically regulated in malignant melanoma and associated with metastatic site specificity. Br J Cancer 106:1446–1452 Warton K, Samimi G (2015) Methylation of cell-free circulating DNA in the diagnosis of cancer. Front Mol Biosci 2:13 Weide B, Martens A, Hassel JC, Berking C, Postow MA, Bisschop K, Simeone E, Mangana J, Schilling B, Di Giacomo AM, Brenner N, Kahler K, Heinzerling L, Gutzmer R, Bender A, Gebhardt C, Romano E, Meier F, Martus P, Maio M, Blank C, Schadendorf D, Dummer R, Ascierto PA, Hospers G, Garbe C, Wolchok JD (2016) Baseline biomarkers for outcome of melanoma patients treated with Pembrolizumab. Clin Cancer Res 22:5487–5496 Willingham SB, Volkmer JP, Gentles AJ, Sahoo D, Dalerba P, Mitra SS, Wang J, Contreras-Trujillo H, Martin R, Cohen JD, Lovelace P, Scheeren FA, Chao MP, Weiskopf K, Tang C, Volkmer AK, Naik TJ, Storm TA, Mosley AR, Edris B, Schmid SM, Sun CK, Chua MS, Murillo O, Rajendran P, Cha AC, Chin RK, Kim D, Adorno M, Raveh T, Tseng D, Jaiswal S, Enger PO,

D. Schadendorf et al. Steinberg GK, Li G, So SK, Majeti R, Harsh GR, van de Rijn M, Teng NN, Sunwoo JB, Alizadeh AA, Clarke MF, Weissman IL (2012) The CD47-signal regulatory protein alpha (SIRPa) interaction is a therapeutic target for human solid tumors. Proc Natl Acad Sci USA 109:6662–6667 Wilson BJ, Saab KR, Ma J, Schatton T, Putz P, Zhan Q, Murphy GF, Gasser M, Waaga-Gasser AM, Frank NY, Frank MH (2014) ABCB5 maintains melanoma-initiating cells through a proinflammatory cytokine signaling circuit. Cancer Res 74:4196–4207 Winnepenninckx V, Lazar V, Michiels S, Dessen P, Stas M, Alonso SR, Avril MF, Ortiz Romero PL, Robert T, Balacescu O, Eggermont AM, Lenoir G, Sarasin A, Tursz T, van Den Oord JJ, Spatz A (2006) Gene expression profiling of primary cutaneous melanoma and clinical outcome. J Natl Cancer Inst 98:472–482 Wolchok JD, Chiarion-Sileni V, Gonzalez R, Rutkowski P, Grob JJ, Cowey CL, Lao CD, Wagstaff J, Schadendorf D, Ferrucci PF, Smylie M, Dummer R, Hill A, Hogg D, Haanen J, Carlino MS, Bechter O, Maio M, MarquezRodas I, Guidoboni M, McArthur G, Lebbe C, Ascierto PA, Long GV, Cebon J, Sosman J, Postow MA, Callahan MK, Walker D, Rollin L, Bhore R, Hodi FS, Larkin J (2017) Overall survival with combined Nivolumab and Ipilimumab in advanced melanoma. N Engl J Med 377:1345–1356 Yuan J, Zhou J, Dong Z, Tandon S, Kuk D, Panageas KS, Wong P, Wu X, Naidoo J, Page DB, Wolchok JD, Hodi FS (2014) Pretreatment serum VEGF is associated with clinical response and overall survival in advanced melanoma patients treated with ipilimumab. Cancer Immunol Res 2:127–132 Zager JS, Gastman BR, Leachman S, Gonzalez RC, Fleming MD, Ferris LK, Ho J, Miller AR, Cook RW, Covington KR, Meldi-Plasseraud K, Middlebrook B, Kaminester LH, Greisinger A, Estrada SI, Pariser DM, Cranmer LD, Messina JL, Vetto JT, Wayne JD, Delman KA, Lawson DH, Gerami P (2018) Performance of a prognostic 31-gene expression profile in an independent cohort of 523 cutaneous melanoma patients. BMC Cancer 18:130 Ziani L, Safta-Saadoun TB, Gourbeix J, Cavalcanti A, Robert C, Favre G, Chouaib S, Thiery J (2017) Melanoma-associated fibroblasts decrease tumor cell susceptibility to NK cell-mediated killing through matrix-metalloproteinases secretion. Oncotarget 8:19780–19794 Zuo Q, Liu J, Huang L, Qin Y, Hawley T, Seo C, Merlino G, Yu Y (2018) AXL/AKT axis mediated-resistance to BRAF inhibitor depends on PTEN status in melanoma. Oncogene 37:3275–3289

Part II Diagnosis and Staging

Clinical Presentations of Melanoma Allan C. Halpern , Ashfaq A. Marghoob, Arthur J. Sober, Victoria Mar, and Michael A. Marchetti

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Patterns of Presentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Clinical Assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Patient History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Personal History of Skin Cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Family History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Phototype and Sun Exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Signs and Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Physical Examination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Clinical Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

109 109 110 111 111 111 116 116

Aids to Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Clinical Photography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dermoscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reflectance Confocal Scanning Laser Microscopy (RCM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . Image Analysis for Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Other Techniques: Multispectral Imaging, Electrical Impedance Spectroscopy, Adhesive Patch Molecular Assays, Optical Coherence Tomography, and Ultrasound Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Evolving Paradigms in the Visual Assessment of Skin Lesions . . . . . . . . . . . . . . . . . . . . . . .

127 128 129 131 132

132 134

A. C. Halpern (*) · A. A. Marghoob · M. A. Marchetti Dermatology Service, Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY, USA e-mail: [email protected]; [email protected]; [email protected] A. J. Sober Department of Dermatology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA e-mail: [email protected] V. Mar Victorian Melanoma Service, Alfred Hospital, Melbourne, VIC, Australia e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_9

107

108

A. C. Halpern et al. Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

Abstract

Cutaneous melanoma is unique among cancers in that it can be readily identified through visual examination of the skin surface. In this chapter, we detail patterns of melanoma presentation as well as appropriate clinical assessment to facilitate early diagnosis. The major histogenic types of melanoma are superficial spreading melanoma, nodular melanoma, lentigo maligna melanoma, and acral lentiginous melanoma; each differs in their associations with age, sex, race, anatomic site, ultraviolet exposure, and molecular profile. The cardinal clinical feature of all types of melanoma, however, is change in size, shape, and color, eventually becoming distinctly different from the remainder of a patient’s skin lesions (i.e., the ugly duckling sign). Variant, uncommon clinical presentations of melanoma, such as amelanotic, desmoplastic, and spitzoid types, are summarized. Finally, we outline aids to the diagnosis of melanoma, including established tools, such as photography and dermoscopy, as well as emerging ones like reflectance confocal microscopy, artificial intelligence-based diagnostic systems, electrical impedance spectroscopy, and adhesive patch molecular assays.

Introduction Prompt and accurate clinical assessment of melanoma remains an important strategy to reducing morbidity and mortality associated with this disease. Through increased public and physician awareness and knowledge of melanoma, there is a trend toward diagnosis of disease at an earlier stage with significant improvement in long-term survival (Rigel and Carucci 2000). As a result of progress in early detection and primary prevention, deaths from melanoma have recently decreased in younger cohorts but continue to

increase in those over 55, especially men (Curchin et al. 2018) (see chapter ▶ “Clinical Epidemiology of Melanoma”). Increased detection pressure has been associated with rising incidence of melanoma in situ. Continued improvements in the early clinical recognition of melanoma are needed, especially for high-risk individuals, while simultaneously improving the specificity of diagnosis. This chapter broadly reviews a general approach to the early diagnosis of melanoma with attention to the varying presentations of the different histogenic subtypes. More details on risk factors, screening, and technologic aids to diagnosis can be found in chapters ▶ “Clinical Genetics and Risk Assessment of Melanoma,” ▶ “Melanoma Prevention and Screening,” and ▶ “Dermoscopy/Confocal Microscopy for Melanoma Diagnosis,” respectively.

Patterns of Presentation Several studies have addressed the pattern of melanoma detection and factors that have an impact on delays in diagnosis (Cassileth et al. 1988; Hennrikus et al. 1991; Negin et al. 2003; Oliveria et al. 1999; Richard et al. 2000a; Richard et al. 2000b; Schmid-Wendtner et al. 2002; Temoshok et al. 1984). Most melanomas currently are selfdetected by either the patient or a member of the immediate family (Aviles-Izquierdo et al. 2016; Betti et al. 2003; Brady et al. 2000; Carli et al. 2004c; Fisher et al. 2005; Koh et al. 1992). However, physicians detect approximately 80% of second primary tumors (Fisher et al. 2005). The majority (~88%) of lethal melanomas are found by non-physicians (Aviles-Izquierdo et al. 2016). The major component of delay in patient-detected melanomas is lack of concern (Betti et al. 2003). A personal history of melanoma is more predictive of a thinner Breslow depth at the time the patient is first seen than a family history of melanoma

Clinical Presentations of Melanoma

(Fisher et al. 2005). Women detect a higher percentage of melanomas than men, both in themselves and in their spouses (Koh et al. 1992). Given the importance of melanoma self-detection, public education campaigns aimed at raising awareness of melanoma and increasing knowledge of the early warning signs of melanoma have potential for reducing the melanoma mortality rate (see chapter ▶ “Melanoma Prevention and Screening”). To reduce patient delays in seeking treatment, educational messages should adequately stress the need for prompt referral to a physician once a suspicious pigmented lesion is self-detected. However, it has been noted that melanomas detected by a physician either in the screening or case-finding setting tend to be diagnosed at a thinner Breslow thickness (50) and freckles (Liu et al. 2006). Together these studies and others (Chamberlain and Kelly 2004) suggest that men older than 50 years of age constitute a distinct group with a higher risk of undetected melanoma and should be targeted in special screening programs (Aitken et al. 2006; Geller et al. 2007; Janda et al. 2006).

Clinical Assessment Elements of the clinical encounter relevant to early detection of melanoma are patient history, physical examination, and diagnostic aids.

Patient History The key components of the patient history are questions pertaining to assessment of melanoma risk and questions pertaining to the detection of current melanomas. Risk-related questions include an assessment of family history of melanoma, personal history of skin cancer and/or nevus excision, sun exposure, and phototype. Questions pertaining to the presence of melanoma relate to a history of a changing, worrisome, or symptomatic lesion. Multivariable risk prediction models for melanoma commonly include age, number of nevi, skin phototype, freckling, hair color, and sunburn history, and the few that have been validated show good discrimination (Olsen et al. 2018a; Usher-

110

Smith et al. 2014; Vuong et al. 2014). Integration of genetic determinants of risk into these models (e.g., MC1R genotype and melanoma susceptibility SNPs) may provide some improvement in discrimination, though further validation is required (Cust et al. 2013). An analysis of the American Academy of Dermatology Skin Cancer Screening Program indicates that 5 factors independently increased the likelihood of finding a suspected melanoma in the 362,804 people screened (Goldberg et al. 2007). They are represented by the mnemonic HARMM, which stands for history of previous melanoma (OR = 3.3; 95% CI 2.9–3.8), age greater than 50 years (OR = 1.2; 95% CI 1.1–1.3), regular dermatologist absent (OR = 1.4; 95% CI 1.3–1.5), mole changing (OR = 2.0; 95% CI 1.9–2.2), and male sex (OR = 1.4; 95% CI 1.3–1.5). Individuals at highest risk for melanoma (4–5 of these factors) composed only 5.8% of the total population, yet accounted for 13.6% of presumptive cases of melanoma and were 4.4 times (95% Cl 3.8–5.1) more likely to be diagnosed with suspected melanoma than those at lowest risk (0 or 1 of these factors).

Personal History of Skin Cancer Patients with a personal history of melanoma (Bradford et al. 2010; Chen et al. 2015) or nonmelanoma skin cancer (Wu et al. 2017) are at increased risk for developing subsequent melanomas. Approximately 1–8% of patients with melanoma will develop multiple primary melanomas according to retrospective studies (StamPosthuma et al. 2001). Atypical moles are strongly associated with increased risk of multiple primary melanomas (see chapter ▶ “Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma”) (Marghoob et al. 1996; Titus-Ernstoff et al. 2006). A single institutional series of 4484 cases of melanoma found that 8.6% of patients had 2 or more primary melanomas when they were first seen (Ferrone et al. 2005). Among these patients, 59% had a second primary tumor within 1 year, and 21% had a family history of melanoma

A. C. Halpern et al.

compared with only 12% of patients with a single primary melanoma ( p < 0.001); 38% of patients with multiple primary melanomas had dysplastic nevi compared with 18% of those with a single primary melanoma ( p < 0.001). Patients who had a positive family history of melanoma or dysplastic nevi had an estimated 5-year risk of multiple primary melanomas of 19.1% and 23.7%, respectively. The most striking increase in incidence for the population with multiple primary melanomas was seen for development of a third primary melanoma from the time of the second primary melanoma, which was 15.6% at 1 year and 30.9% at 5 years (Ferrone et al. 2005). Approximately one third of multiple primary melanomas are found concurrently (synchronous) with the diagnosis of the first melanoma, and two thirds are found sequentially (metachronous) during follow-up, with some being diagnosed more than 30 years after the first diagnosis. It stands to reason that a history of melanoma indicates that the person may have a genetic susceptibility to melanoma and/or have had the causative environmental exposure necessary to form melanoma. Germline mutations in CDKN2A, CDK4, and MITF have been associated with both family history of melanoma and development of multiple primary melanomas (Ferrone et al. 2005; Puig et al. 2005; Yokoyama et al. 2011). The genes, environment, and melanoma study identified several other low penetrance susceptibility loci associated with increased risk of developing subsequent melanomas (Gibbs et al. 2015). In patients with multiple cutaneous melanomas, synchronous or subsequent primary melanomas need to be distinguished from epidermotropic metastases, because the prognosis and treatment differ between the two (Abernethy et al. 1994; Gerami et al. 2006; Mehregan et al. 1995; White and Hitchcock 1998). There is conflicting evidence for the effect of multiple primary melanomas on survival given the inherent complexity in estimating survival in this group. The “delayed entry” approach has been advocated to avoid survival bias, and studies using this method have reported poorer survival in patients with multiple primary melanoma independent of other prognostic factors (Rowe et al. 2015).

Clinical Presentations of Melanoma

Family History It has been demonstrated that the validity of the family history of melanoma is poor (Weinstock and Brodsky 1998). This stems, in part, from the erroneous yet common interchangeable use of “melanoma” and “skin cancer.” Therefore patients should be educated in the distinction between melanoma and other types of skin cancer before a history of melanoma is elicited from them. It is advisable to confirm the family history on a follow-up visit once the patient has had the opportunity to specifically question family members, with the added benefit of a greater understanding of the types of skin cancer. Confirmation of family history by pathology report is considered the gold standard. In patients with a positive family history or personal history of melanoma, it is appropriate to recommend screening of other family members. It is estimated that 5–10% of melanoma cases are hereditary, although this varies depending on the background incidence of melanoma in different regions (Leachman et al. 2009). CDKN2A germline mutations are strongly associated with familial melanoma although the penetrance varies by environmental exposures; mutations in CDK4, BAP1, POT1, ACD, TERF2IP, and TERT are rare and account for a small percentage of familial melanoma cases. It is estimated that a mutation in any one of the above genes is implicated in only 50% of melanoma dense kindreds (Read et al. 2016). The likelihood of a CDKN2A mutation being responsible for a familial melanoma cluster increases with number of family members affected, presence of multiple primary melanomas, early age of melanoma diagnosis, and familial cases of pancreatic cancer. In such cases where there is a strong family history (three or more first- or seconddegree relatives) and other predictive factors present, genetic counseling and testing should be discussed (Leachman et al. 2009; Mann).

111

Patients should be questioned about their natural hair color and eye color, as these may be difficult to ascertain on physical examination because of canities and the use of hair dyes and colored contact lenses. A general assessment of occupational and recreational sun exposure, as well as a history of severe sunburn, should be elicited.

Signs and Symptoms Patients should be questioned regarding the presence of any worrisome or changing skin lesions. A history of change is elicited more often in lesions that prove to be melanomas compared with lesions that are benign (Kittler et al. 1999). Specific questioning is often required to elicit a history of symptomatic lesions, for example, itching, bleeding, or lesions that are easily irritated. Questions regarding the presence of birthmarks and moles on unusual anatomic sites often can alert the physician to examine these areas more closely. The cardinal clinical feature of cutaneous melanoma is a pigmented skin lesion that changes visibly over a period of months to years. Sometimes the change is so gradual that the patient is unaware of it. Changes in pigmented lesions that occur over the course of days are typically inflammatory or traumatic in nature. However, as a general rule, any lesion noted to have changed in color, shape, size, or elevation warrants medical attention. Some of the presenting signs are shown in Figs. 1, 2, 3 and 4. Bleeding, itching, Table 1 Classification of skin phototype Type I II III IV

Phototype and Sun Exposure

V

Questions regarding burning tendency and tanning ability should be asked to determine the patient’s phototype as described in Table 1.

VI

Description Always burns; never tans Always burns; sometimes tans Sometimes burns; always tans Rarely burns; always tans Burns and tans after extreme UVexposure Burns and tans after extreme UVexposure

Population affected Ivory white Caucasian (e.g., Celtic) Fair Caucasian Caucasian Olive-skinned Caucasian Dark-skinned Caucasian (Latino, Indian, etc.) Black

112

Fig. 1 (a and b) Melanoma in situ. Note variation in pigment pattern. (c) Lentigo maligna melanoma. Note variation in pigment pattern. (d) Lentigo maligna melanoma. Note highly irregular borders and background of chronic actinic damage. (e) Small melanoma exhibiting variation in color. (f) Melanoma exhibiting irregular

A. C. Halpern et al.

borders. (g) Radial growth phase superficial spreading melanoma (0.64 mm). (h) Intermediate-risk superficial spreading melanoma (1.72 mm) (a, b, c, g, and h courtesy of R.A. Johnson, MD; d and e courtesy of C.M. Balch, MD)

Clinical Presentations of Melanoma

Fig. 2 (a) Advanced superficial spreading melanoma with asymmetry, irregular borders, and variation in color. (b) Melanoma with radial and early vertical growth phases clinically. (c) Melanoma with radial and advanced vertical growth phases. (d) Melanoma with irregular border and

113

blue-black coloration. (e and f) Advanced melanoma with clinical ulceration. (g) Relatively lightly pigmented melanoma. (h) Acral lentiginous melanoma arising in a nevus (a and g courtesy of C.M. Balch, MD; h courtesy of R.A. Johnson, MD)

114

Fig. 3 (a) Advanced melanoma with highly irregular borders and variations in color and pigment pattern. (b) Relatively amelanotic melanoma with radial and vertical growth phases. (c) Halo melanoma. (d) Amelanotic superficial spreading melanoma, 1.2 mm thick. Note multiple

A. C. Halpern et al.

punch biopsy sites. (e) Relatively amelanotic melanoma. (f) Superficial spreading melanoma with regression in upper left corner. (g) Melanoma with central regression. (h) Advanced melanoma with central regression (b, g, and h courtesy of C.M. Balch MD)

Clinical Presentations of Melanoma

Fig. 4 (a) Acral lentiginous melanoma in situ in web space. (b) Advanced acral lentiginous melanoma on plantar surface. (c) Advanced clinically ulcerated acral lentiginous melanoma on plantar surface. (d) Conjunctional melanoma. (e) Penile melanoma. Patient had metastases to groin nodes. (f) Vulvar melanoma. (g) Recurrent

115

melanoma in margin of skin graft. (h) Diffuse melanosis resulting from advanced metastatic melanoma. Note bluish-gray color of skin, gingiva, and nail beds (a courtesy of R.A. Johnson, MD; c, d, and e courtesy of C.M. Balch, MD)

116

tenderness, and ulceration can be associated with cutaneous melanoma. Bleeding and ulceration are typically signs of more advanced local disease. On the other hand, it is not uncommon for patients to report unusual sensations in early melanomas, including melanoma in situ. Although it is often difficult for patients to verbalize the exact nature of the sensation or the cause of their concern, lesions that are a source of concern to a patient should be taken seriously. It is not uncommon for melanomas that defy clinical diagnosis on morphologic grounds to be excised strictly on the basis of patient’s insistence (Andersen and Silvers 1991). Furthermore, the presenting signs and symptoms of melanoma reported by patients differ between young and older patients. Younger patients have been reported to more often have a history of change in color or contour and have signs of itching (Christos et al. 2000), whereas older patients more often have a history of ulceration, which is a poor prognostic sign (Christos et al. 2000).

Physical Examination Total body skin examination serves to ascertain melanoma risk factors, such as mole pattern, mole type, freckles, and so forth, and is essential for early detection of melanoma. In addition, total body skin examination performed by the physician demonstrates to the patient proper technique for skin self-examination. The examination should be performed with the patient fully disrobed and appropriately draped to permit a complete examination while addressing the issues of modesty and patient comfort. Lighting that is sufficiently bright is required and may be facilitated by a light source that can be readily manipulated during the course of the examination. Various poses and positions have been recommended for total body skin examination (Kopf et al. 1995). Regardless of the positions used, a systematic consistent approach is critical to ensure a comprehensive examination. All cutaneous surfaces including intertriginous areas, web spaces, and the scalp should be examined. Nails should be examined after all nail polish has been

A. C. Halpern et al.

removed. Genital, ocular, and mucous membrane examinations should be performed or recommended as part of the patient’s routine gynecologic, ophthalmologic, and dental examinations. When examining the oral cavity, it is important to remove any dentures that could obscure lesions (Dimitrakopoulos et al. 1998). Approximately 80% of melanomas arising in the oral mucosa occurred on the maxillary anterior gingival area, especially on the palatal and alveolar mucosa (Ebenezer 2006; Ulusal et al. 2003). Features to be noted on skin examination include the approximate number of nevi, the presence of atypical/dysplastic nevi, and the presence of actinic damage such as actinic keratoses, dermatoheliosis, solar lentigines, and poikiloderma. The presence of congenital nevi, halo nevi, acral nevi, and scalp nevi should be noted. Some simple measures can aid in the examination of certain anatomic sites and lesions. For the scalp examination, some prefer to use a hair blower, whereas others prefer to use a comb to methodically part the hair. Examination of pigmented lesions of the nails, palms, and soles is facilitated by swabbing the surface with mineral oil or alcohol to render the nail plate or thickened stratum corneum translucent. Wood’s lamp examination can be helpful in assessing the presence of halo nevi or leukoderma or defining the margins of atypical lentiginous lesions (Reyes and Robins 1988). When faced with a highly unusual macular pigmented lesion (Fig. 5), cleansing of the surface with an alcohol swab can prevent unnecessary biopsy of the occasional pseudo-lesion, such as a stain from hair dye or adherent dirt.

Clinical Features The clinical features of melanoma vary by anatomic site and growth pattern; this is also referred to as histogenic type. These growth patterns, in turn, vary in incidence by sex, age, and race (Crombie 1979; Reintgen et al. 1982; Wang et al. 2016) (Table 2). The discovery of various molecular markers has offered the possibility of more detailed subclassification beyond growth

Clinical Presentations of Melanoma

117

Fig. 5 Pseudolesion. This “pigmented lesion” on the scalp was referred for biopsy because of its irregularity and central regression. The “lesion” rubbed off during preparation with an isopropyl alcohol rub

Table 2 Age-adjusted US melanoma incidence rates per 100,000 person-years, stratified by race, age, and gender Non-Hispanic White Superficial spreading melanoma All ages 9.05 10% of melanocytic lesions; they are suggestive of melanoma (Ribero et al. 2016). In addition, a combination of gray dots and gray circles, called annular-granular pattern, is a sign of lentigo maligna (Fig. 10) (Kittler et al. 2016; Soyer et al. 2001a). Pigmented circles are often a sign of malignancy in facial lesions. Incomplete circles, consisting of asymmetrically pigmented follicular openings, and concentric circles are signs of lentigo maligna (Kittler et al. 2016; Schiffner et al. 2000; Soyer et al. 2001a). These circles sometimes connect with each other and form lines that angulate from each other forming polygonal/rhomboidal structures. A blue structureless zone can describe a diffuse blue area, usually covering the whole lesion

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

155

Fig. 5 Irregular or atypical networks. (a) Thickened lines in an atypical nevus. (b) Lines ending abruptly at the edge of the lesion in a melanoma in situ. (c) Smudged or out of focus lines in a melanoma in situ

Fig. 6 Radial lines (streaks) distributed irregularly within an atypical Spitz nevus

as found in benign blue nevi, or a blue-white veil (Fig. 11), which is a blue or blue-gray area overlaid by a whitish, ground-glass-like haze, representing heavily pigmented melanocytes in the upper dermis overlaid by an acanthotic epidermis with compact orthokeratosis (Yélamos et al. 2018a). The presence of a blue-white veil over a raised area should raise suspicion for invasive melanoma, and lesions with structureless blue areas that do not cover the whole lesion are suspicious for melanoma. Blue structureless zones are often associated with thick or multicolored reticular lines, dots, clods, or atypical lines, and blue structureless areas are often palpable. Blue structureless areas are commonly found in melanoma and not commonly found in benign Clark nevi, though it can appear in benign Reed or Spitz nevi (Soyer et al. 2001a; Wolner et al. 2017). Brown or black structureless zones, also called blotches (Fig. 12), are dark brown to black areas

which obscure the visualization of other underlying dermoscopic structures. Blotches represent areas of heavy melanin pigmentation throughout the epidermis or upper dermis or a lamella of pigment in the stratum corneum. Structureless zones can be described as localized, being confined to a small section of the lesion, or diffuse, spreading across most or all of the lesion. Regular structureless zones are symmetrical in shape and are distributed more or less symmetrically throughout the lesion; these suggest a benign lesion. Irregular structureless zones are either asymmetric in shape or are distributed asymmetrically in the lesion, such as a single off-center blotch near the periphery of the lesion or multiple blotches distributed asymmetrically. These irregular zones suggest melanoma. However, areas of structureless, heavy pigmentation are so common in pigmented lesions that their diagnostic significance is limited (Soyer et al. 2001a; Wolner et al. 2017). Hypopigmented structureless areas in an otherwise normal pigmented lesion represent areas in the epidermis and dermis that have little melanin and can be focal, multifocal, or diffuse (Fig. 13). Various kinds of hypopigmentation are found in benign nevi, and some melanomas display irregularly outlined hypopigmented areas. Hypopigmented reticular lines around pigmented clods, also called a negative network (Fig. 14), are usually found in melanomas or Spitz/Reed nevi. The clods can be elongated or laid out in a serpiginous fashion (Kittler et al. 2016; Pizzichetta et al. 2013; Soyer et al. 2001a; Wolner et al. 2017). Scar-like areas are white structureless areas, usually paler than nearby normal skin, and do not have shiny white lines or visible blood vessels (Fig. 15a). Scar-like areas represent areas of regression and

156

K. J. Lee et al.

Fig. 7 Pseudopods. (a) Regular pseudopods around a growing nevus. (b) Pseudopods distributed irregularly at periphery of a melanoma (Breslow thickness 0.3 mm)

Fig. 8 White lines with perpendicularity (shiny white lines or crystalline structures) in melanomas. (a) Breslow thickness 1.6 mm. (b) Breslow thickness 1.4 mm

fibrosis. When white and blue areas appear alongside each other, they can be difficult to distinguish from a blue-white veil for novice dermoscopists, but they are usually not palpable while an area of bluewhite veil is often palpable (Fig. 15b). While white scar-like areas occasionally appear in benign nevi, they are highly specific for melanomas. One exception is a central white structureless area, usually found in dermatofibromas (Fig. 16) (Soyer et al. 2001a; Wolner et al. 2017).

Vascular structures can also be visible with dermoscopy, if noncontact polarized dermoscopy is used or if contact dermoscopy is performed without pressing too hard or by using ultrasound gel (Figs. 17 and 18). Milky red areas, which are light pink structureless areas with occasional presence of vessels, strongly suggest melanoma. Other vascular structures that suggest melanoma are serpentine vessels (linear with multiple bends, also called linear irregular), helical vessels

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

157

Fig. 9 (a) Irregular/asymmetrically distributed clods (globules) in an atypical nevus. (b) Asymmetric peripheral dots and clods in a melanoma (Breslow thickness 0.45 mm)

Fig. 10 Gray dots (peppering) in a melanoma (Breslow thickness 0.5 mm)

(vessels twisted into loops along a central axis, also called corkscrew), and polymorphous vessels (more than one type of vascular structure) (Fig. 19). Linear vessels, which are straight or mildly curved, can be a sign of melanoma when they are irregular with different sizes, shapes, and curves and a haphazard or asymmetrical distribution (Braun et al. 2004; Carrera et al. 2016; Kittler et al. 2016; Soyer et al. 2001a; Wolner et al. 2017).

Common Global Dermoscopic Patterns There are eight common global dermoscopic patterns for melanocytic lesions: reticular, globular, cobblestone, homogenous, starburst, parallel (acral

surfaces), unspecific, and multicomponent. When regular, they are usually indicative of a melanocytic nevus, while if they are irregular, unspecific, or multicomponent, they are often associated with melanoma. A uniform, symmetrical, or regularly distributed pattern is usually a sign of a benign nevus, but abnormalities in several of the usually benign patterns are indicative of melanoma. When evaluating acral skin, generally pigment located on the furrows (parallel furrow pattern) is suggestive of a nevus, whereas pigment located on the ridges (parallel ridge pattern) is suggestive of melanoma. Reticular pattern, also called a pigment network, is a network or grid of brown lines, commonly found in benign acquired melanocytic nevi, solar lentigines, and thin melanomas (Fig. 20). Despite its presence in such a broad range of lesions, alterations in the thickness, color, or uniformity of the lines can indicate a malignant lesion (irregular pigment pattern). Regular, light to dark brown lines that fade toward the periphery of the lesion are typical of a benign nevus, while thickened, gray, brown, or black lines distributed irregularly and ending abruptly at the edge of the lesion are indicative of melanoma. Previously, a facial feature presenting as round meshes corresponding to follicular openings was called a pseudonetwork (Fig. 21) (Soyer et al. 2001a); Globular pattern consists of numerous small, round to oval brown and black clods, of varying sizes, or dots, no bigger than the diameter of a terminal hair (Kittler et al. 2016). This pattern is

158

K. J. Lee et al.

Fig. 11 Blue-white veil in melanomas. (a) Breslow thickness 1.4 mm. (b) Breslow thickness 0.7 mm

Fig. 12 Structureless black zone (blotch/lamella) at the center of a nevus

Fig. 13 Multifocal hypopigmentation in an atypical nevus

also common in benign nevi and can appear in combination with reticular lines (Soyer et al. 2001a). Circumferential brown clods, a rim of small clods around the edge of a lesion, indicate a growing melanocytic lesion (Fig. 22). Cobblestone pattern consists of large polygonal clods packed closely together, resembling cobblestones (Fig. 23). This pattern is common in benign nevi such as papillomatous dermal nevi or congenital nevi (Soyer et al. 2001a). Homogenous pattern is characterized by a structureless zone of diffuse brown, pink, black, or blue pigmentation in the absence of other

structures. It is common in a variety of lesions, such as blue nevi, BAP1-inactivated melanocytic tumors, and nodular or metastatic melanomas (Fig. 24) (Soyer et al. 2001a; Yélamos et al. 2018c). Starburst pattern consists of circumferential pseudopods or circumferential radial streaks arranged symmetrically around the perimeter of the lesion (Fig. 25). It is characteristic of benign Spitz/Reed nevi but can also appear in spitzoid melanomas (Lallas et al. 2017; Soyer et al. 2001a). Parallel furrow patterns are thin parallel lines of pigment on the acral surfaces of the feet or

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

159

Fig. 14 Hypopigmented reticular lines around brown clods (negative network) in melanomas. (a) Breslow thickness 0.6 mm. (b) Breslow thickness 0.7 mm

Fig. 15 Regression structures. (a) Central, irregularly shaped white structureless area (scar-like area) in a melanoma. (b) Blue-gray structureless zones in a melanoma (Breslow thickness 0.3 mm)

hands, arranged along the cristae or sulci, or more rarely at right angles to them (Fig. 26). Thin lines of pigment in the furrows are common in acral nevi, but thick lines of pigment on the ridges are more suggestive of melanoma (Soyer et al. 2001a). Unspecific lesions have none of the above patterns and are often but not always associated with melanoma (Soyer et al. 2001a). Many lesions have a multicomponent or complex pattern, consisting of three or more distinctive dermoscopic patterns within the same lesion

(Fig. 27). For example, a lesion with a cluster of clods, a zone of reticular lines, and a structureless zone of pigment would be considered multicomponent. This pattern is often found in melanoma and basal cell carcinoma and less frequently in benign nevi (Soyer et al. 2001a). Global dermoscopic patterns are influenced by a number of patient-specific factors that may mislead dermoscopists. Benign nevi can change with age: in children, globular or structureless patterns are common, and during adolescence, symmetrically growing nevi with circumferential brown clods are

160

K. J. Lee et al.

common, while in adults, reticular lines are the most common pattern. Nevi are also prone to change during pregnancy, with reversible changes in color, thickness of reticular lines, new clods, or more prominent vasculature. Exposure to UV can also cause reversible changes in size, the darkness of pigment, erythema, and the development of irregular clods or thickened lines. Skin type

influences pigment distribution in nevi, with type I skin tending to light brown nevi with central hypopigmentation, types II and III light to dark brown with multifocal pigmentation, and type IV to dark brown nevi with central hyperpigmentation. Finally, even benign nevi in melanoma patients are more likely to have mixed dermoscopic pattern, such as reticular-globular or globular-structureless, than more uniform nevi (Zalaudek et al. 2009).

Anatomical Site Considerations

Fig. 16 Central dermatofibroma

white

structureless

zone

in

a

Fig. 17 Vascular structures visible with dermoscopy. (a) Curved (comma-like). (b) Dotted. (c) Serpentine (linear irregular). (d) Helical (corkscrew). (e) Polymorphous.

Site-specific anatomical structures influence dermoscopic features in the trunk and extremities, face, palms and soles, nails, and mucosal surfaces, in turn influencing the dermoscopic presentation of in situ and early invasive melanoma (Breslow index 0.75 mm) have usually destroyed the site-specific anatomic structures of the skin, so dermoscopic features of these melanomas are usually independent of the anatomic location (Soyer et al. 2001b). On the trunk, arms, legs, and dorsal surfaces of the hands and feet, more than 70% of melanomas have a multicomponent pattern or reticular lines that are thick or varied in color (atypical pigment network). Irregular forms of dots, clods, radial

(f) Pink structureless clods (milky red globules). (Provided by Ralph P. Braun CC BY-NC 4.0)

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

161

Fig. 18 Vascular structures suggestive of melanoma. (a) Milky red area in a melanoma (Breslow thickness 1.3 mm). (b) Serpentine vessels, here shown in a mucoid cyst, are common in melanoma

Fig. 20 Regular reticular or network pattern in a nevus, featuring reticular lines of similar color and thickness, fading toward the periphery of the lesion Fig. 19 Polymorphous vessels, where a lesion displays several types of vascular structures, are suggestive of malignancy. Here a basal cell carcinoma includes linear, serpentine, curved, and coiled vessels

lines, or structureless zones appear in 50–70% of melanomas, as do regression structures. Blue and white structureless zones (blue-white veil and scarlike depigmentation) are not common features on

the trunk and extremities, but do occur in up to 30% of melanomas (Soyer et al. 2001b). On the face, irregular dots and clods occur in 30–50% of melanomas, as does irregular brown or black structureless zones after the very early stage of melanoma. A brown structureless area interrupted by follicular openings (pseudonetwork) is always present in early invasive facial

162

K. J. Lee et al.

circles around the follicular openings appear in 50–70% of early facial melanomas. Gray structureless zones are formed by the confluence of gray dots and circles around the follicular openings, and angulated or polygonal lines are gray-brown pigment in lines with obvious angles around the follicular openings. These appear in 50–70% of facial melanomas. Concentric pigmented circles appearing within the follicles, gray circles within the follicles, and asymmetric or incomplete circles of pigment around the follicular opening can also indicate melanoma. As the melanoma progresses, structureless zones that obliterate the follicles can also appear (Soyer et al. 2001b; Wolner et al. 2017). On the acral surfaces, a thick parallel line on the ridges (parallel ridge pattern), as opposed to thin parallel lines in the furrows (parallel furrow pattern), is a very common feature, appearing in over 70% of melanomas. Here the pigmentation follows the cristae superficiales rather than the sulci superficiales, which is a common sign of benign acral nevi. Irregular forms of dots, clods, and radial lines also appear in 50–70% of melanomas on the acral surfaces. Blue structureless zones (blue-white veil) is an uncommon feature of acral melanomas but does occur in up to 30% of cases. Subcorneal hemorrhage can appear as a Fig. 21 Structureless zone interrupted by follicular openclinically concerning jet-black macule, but ings (pseudonetwork pattern) in a facial nevus melanomas. It is usually atypical, appearing as gray dots and circles (annular-granular structures), gray structureless zones, or angulated or polygonal lines (rhomboid structures). Gray or blue-gray dots and

Fig. 22 Globular pattern. (a) Central brown clods (globules) in a nevus. (b) Circumferential clods (peripheral rim of globules) in a growing nevus

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

dermoscopic examination usually reveals reddish pigmentation surrounded by small reddish dots (Soyer et al. 2001b). Melanoma of the nail apparatus appears dermoscopically as multiple brown and black lines, with irregular thickness and arrangement. There may also be interruptions in the pigment bands, and bands are sometimes not parallel (Wolner et al. 2017). Benign nevi of the nail apparatus usually feature a regular band-like

Fig. 23 Cobblestone version

163

pattern, with parallel bands (Zalaudek et al. 2009). Hemorrhage under the nails may present as a clinically concerning round to oval, sharply circumscribed black area. Dermoscopically, these are distinguished from melanomas by dark red or red-black pigmentation surrounded by small reddish dots that are not visible clinically (Soyer et al. 2001b). Melanoma of the mucosal surfaces, incorporating the lips, mouth, nasal cavity, and anal and genital surfaces, features structureless zones and gray color in the early stages, progressing to multiple colors and patterns, particularly blue, white, or gray (Wolner et al. 2017). Benign nevi here usually have a globular mixed pattern (Zalaudek et al. 2009), and benign melanoses usually have a pattern of parallel linear or curvilinear brown streaks over a diffuse pigmentation. Benign melanosis sometimes shares features of melanoma, which can only be ruled out by histopathological analysis (Soyer et al. 2001b). Once a melanoma has progressed to an intermediate or thick stage, anatomic-specific features are less evident. In melanomas with a Breslow index of >0.75 mm, regardless of anatomical site, blue structureless zones are a very common feature, appearing in more than 70% of these thicker melanomas. Irregular forms of dots, clods, radial lines, and structureless zones are

Fig. 24 Homogenous pattern. (a) Blue-black homogenous pattern in a melanoma (Breslow thickness 3.7 mm). (b) Blue homogenous pattern in a plantar nevus

164

K. J. Lee et al.

Fig. 25 Starburst pattern. (a) With radial lines (streaks) in an atypical Spitz/Reed nevus. (b) With pseudopods in a Spitz/ Reed nevus

Fig. 26 Parallel furrow pattern of thin parallel lines in the furrows of acral nevi

Fig. 27 Multicomponent pattern in (a) a melanoma (Breslow thickness 1.3 mm) and (b) a nevus

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

also found in 50–70% of these melanomas, and thick or multicolored reticular lines (atypical pigment networks) or polymorphous vascular patterns are found in 30–50%. Blue-gray structures and white structureless zones (regression structures) are less common but still appear in up to 30% of intermediate and thick melanomas (Soyer et al. 2001b).

Featureless Melanomas Although most melanomas display at least some degree of asymmetry of pattern, color, and structure, a subset of early melanomas are featureless. Most of these early melanomas can be correctly identified by observing their growth dynamics or detecting an increased or atypical vasculature.

Diagnostic Algorithms for Dermoscopy While initially only well-trained dermatologists were able to gain increased sensitivity and specificity with dermoscopy (Kittler et al. 2002), the development of the two-step algorithm and other checklists to detect malignancy has enabled primary care physicians to effectively use dermoscopes with a brief online training course (Zalaudek et al. 2006) (see Table 1). However, the more complex method of pattern analysis remains the preferred method of experienced dermoscopists. See Table 2 for a comparison of the sensitivity and specificity of scoring methods.

Two-Step Algorithm The two-step algorithm first classifies the lesion as melanocytic or non-melanocytic (Fig. 28). Pigmented features that indicate a melanocytic lesion are reticular lines on non-glabrous skin, brown to black dots or clods, radial lines, homogeneous blue zones, or parallel lines of pigment on the acral surfaces. In the absence of pigmented features, users look for specific criteria that can identify the lesion as non-melanocytic.

165

Features indicating a non-melanocytic lesion are curved, parallel thin brown lines (fingerprint-like structures); curved thick lines (cerebriform patterns); radial lines connected to a common base (leaflike structures); white dots or clods (milia-like cysts); brown, yellow, or black clods (comedo-like openings); irregular blue-gray clods or structureless zones; large clustered blue clods (blue-gray ovoid shapes); concentric clods (spoke wheel-like areas); branched (arborizing) vessels; red-blue clods (lacunae); red-blue to red-black structureless zones; and ulceration. Featureless lesions that do not show any melanocytic or non-melanocytic lesion structures require special attention. Because it is not uncommon to encounter amelanotic and hypomelanotic melanomas that are structureless, all so-called featureless lesions should be viewed with extreme suspicion, and assessed alongside melanocytic lesions, especially if the lesion exhibits irregular linear or dotted blood vessels, both of which are commonly seen in melanoma (Argenziano et al. 2003; Soyer et al. 2001b). The second step of the two-step algorithm involves the assessment of melanocytic and featureless lesions with a checklist of dermoscopic features to differentiate benign nevi from melanoma. Pattern analysis is typically used by experienced dermoscopists, while novices in dermoscopy may find one of the score-based algorithms described below useful in assessing melanocytic lesions. A study by the International Dermoscopy Society, where 130 participants scored 477 lesions for the presence or absence of specific dermoscopic features found in various score-based algorithms, found that the Menzies method had the highest sensitivity (95.1%) but the lowest specificity (24.8%); the other algorithms had similar levels of diagnostic accuracy (68.9–77.9%) (Carrera et al. 2016).

Three-Point Checklist With the three-point checklist, users score the lesion on the presence of three criteria:

166

K. J. Lee et al.

Table 1 Diagnostic algorithms and checklists for dermoscopy. These methods apply only after a lesion has been assessed to be melanocytic Method Components Three-point 1. Asymmetry in color or structure checklista 2. Irregular reticular lines with thickened lines and irregular holes 3. Blue or white pigmentation ABCD or Asymmetry: in 0, 1, or 2 axes, in contours, color, or ABCDstructure EFG ruleb, c Border: abrupt cutoff of pigment pattern at the periphery in 0–8 sectors Colors: presence of up to six colors (white, red, light or dark brown, blue-gray, black) Dermoscopic structures: presence of reticular lines, structureless zones, branched lines, dots, or clods For nodular melanomas: elevated, firm, and growing for at least 1 month Negative features Menzies scoringd 1. Symmetry of pattern: symmetry of pattern across all axes through the center of the lesion; symmetry of shape is not required 2. Single color: black, gray, blue, dark brown, tan, or red Positive features 1. Blue structureless zone (blue-white veil): irregular structureless area of confluent blue pigmentation with an overlying white “groundglass” haze, not occupying the whole lesion or associated with red-blue clods 2. Multiple brown dots: focal areas of dots, not clods 3. Pseudopods: bulbous, often kinked projections at the edge of the lesion, connected to the lesion body or pigment network and not distributed regularly or symmetrically around the lesion 4. Radial segmental lines (radial streaming): finger-like extensions at the edge of the lesion, not distributed regularly or symmetrically around the lesion 5. White structureless zone (scar-like depigmentation): areas of white, distinct, irregular extensions, not confused with areas of hypopigmentation or depigmentation caused by simple loss of melanin 6. Peripheral black dots or clods: found at or near the edge of the lesion 7. Five to six colors: gray, black, blue, dark brown, tan, or red; white is not scored as a color 8. Multiple blue or gray dots: foci of multiple blue or gray dots, not clods, often described as “pepperlike” 9. Thick reticular lines: network made of irregular thick “cords,” often seen focally thicker

Scoring 1 1 1 0–2 0–8 1–6 1–5

Formula and interpretation 2 or more is indicative of a melanoma or BCC

(A score  1.3) + (B score  0.1) + (C score  0.5) + (D score  0.5) < 4.75 = benign nevus 4.75–5.45 = suspicious for melanoma >5.45 = melanoma

Lesions with both of these features are not melanomas

Lesions with one or more of these features are suspicious for melanoma

(continued)

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

167

Table 1 (continued) Method Revised seven-point checkliste

CASH scoring systemf

Components 1. Reticular lines that are thick or that vary in color: a combination of two or more types of pigment network, for example, with different colors or line thickness, distributed asymmetrically within the lesion 2. Blue structureless zone (blue-white veil): an area of irregular, confluent structureless blue pigmentation with an overlying whitish area resembling a “ground-glass” film; the pigmentation must not occupy the whole lesion 3. Atypical vascular pattern: serpentine or dotted vessels or pink structureless areas, not clearly combined with regression structures 4. Irregular radial lines (streaks): more than three brown or black bulbous or finger-like projections, not clearly combined with a pigment network and asymmetrically distributed around the edge of the lesion 5. Irregular structureless zones (blotches): structureless brown or gray areas distributed asymmetrically within the lesion 6. Irregular dots/clods: more than three dots or clods, black or brown, distributed asymmetrically within the lesion 7. Regression structures: scar-like white structureless zones or blue pepper-like dots (granules) Color: Light brown Dark brown Red Black White Blue Architectural disorder: None/mild Moderate Marked Symmetry in shape and dermoscopic structures: Biaxial symmetry Monoaxial symmetry Biaxial asymmetry Homogeneity/heterogeneity based on number of dermoscopic structures: Dot/clods Radial lines (streaks)/pseudopods Blue structureless zone (blue-white veil) Regression structures: white or blue structureless zones or blue or gray dots and circles Irregular dark structureless zones (blotches) Polymorphous blood vessels

Scoring 1

Formula and interpretation 1 = excision recommended

1

1

1

1

1

1

1 point for each color

2 mm at diagnosis (Kalkhoran et al. 2010). Histologically, NM is characterized by vertical growth without evidence of an associated radial growth beyond the width of three rete ridges beyond the invasive component

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

(Clark et al. 1969). Because its epidemiological and morphological features differ significantly from other forms of melanoma, some postulate that NM may originate in the dermis and only gives rise to clinically recognizable features on the skin when it gains enough tumor volume (Zalaudek et al. 2008). A high level of suspicion is required to recognize early forms of NM; thus, any clinically equivocal growing nodule should always be immediately excised (Moscarella et al. 2017). Fast growing NM more commonly affects men aged >50 years who lack known melanoma risk factors such as multiple nevi, freckles, or signs of sun damage (Chamberlain et al. 2002; Lipsker et al. 2007; Murray et al. 2005), and people with the highest risk for NM are consistently underrepresented during skin cancer screening programs (Hubner et al. 2017). The majority of tumors develop rapidly on previously unaffected skin and are mainly patient detected (Halpern et al. 2014), with patients reporting the tumor “suddenly appearing” or “bulging out” of healthy skin (Chamberlain et al. 2003; Warycha et al. 2008). Clinically, NM usually appears as a symmetric, red to pink or gray-blue/black plaque or nodule with a diameter 6 mm, so the EFG rule and CCC rule have been developed to aid clinical diagnosis of NM. The EFG rule stands for Elevation, Firmness on palpation, and continuous Growth for more than 1 month (Kelly et al. 2003), while the CCC rule stands for Color (uniform red to blue or black), Contour (roundish), and Change (rapid growth) (Moynihan 1994).

Dermoscopic Features of Nodular Melanoma Dermoscopy of pigmented NM often features blue and black colors, with a combination of

179

structureless blue areas and black dots, clods, or structureless areas covering at least 10% of the lesion surface. This blue-black rule enhances the discrimination of NM from other benign nodular tumors, such as angiokeratoma, blue nevus, seborrheic keratosis, or hemangioma (Argenziano et al. 2011b). Although blue color is seen in many benign lesions, the combination with black dots and structureless areas is rarely present in benign lesions. Black color may be predictive for a high risk of ulceration (Longo et al. 2013b). For changing heavily pigmented skin lesions, the blue-black rule represents a highly effective clue for the diagnosis of pigmented NM (Fig. 30). Nonpigmented and hypopigmented NM are also common (Fig. 31) (Pizzichetta et al. 2017). In the case of amelanotic or hypomelanotic NM (AHNM), diagnosis relies on vascular patterns (Zalaudek et al. 2010). While a true amelanotic melanoma lacks any pigmentation, AHNM is characterized by residual or light pigmentation (Menzies et al. 2008). The presence of brown, gray, or blue pigment, especially when seen on the base of a pink nodule, raises the suspicion of melanoma. In the case of amelanotic melanoma, pink structureless zones (milky red areas), short perpendicular white lines (white shiny streaks; only seen under polarized dermoscopy), and a polymorphous vascular pattern are the only diagnostic clues (Pizzichetta et al. 2017; Zalaudek et al. 2010). The most frequent combination of vessel types in melanoma is dotted, linear, and coiled. In contrast to basal cell carcinoma or well-differentiated squamous cell carcinoma, which is characterized by larger vessels, NM more frequently exhibits microvessels of small diameter and length. Pink clods (milky red globules) often show a central serpentine (linear irregular) or helical (corkscrew) vessel, which probably represents neo-angiogenesis (Zalaudek et al. 2010). Thin AHNM is more difficult to diagnose than thick AHNM, because in thin tumors, vascular polymorphism is less evident (Menzies et al. 2008; Pizzichetta et al. 2017; Zalaudek et al. 2010). Short, perpendicular white lines (also known as

180

K. J. Lee et al.

Fig. 30 Blue-black rule in nodular melanomas

Fig. 31 This hypomelanotic nodular melanoma features polymorphous vessels, pink structureless zones, and small pigmented structures at its base (7 o’clock)

shiny white lines, white thick lines, chrysalis or crystalline structures) are seen only under polarized light dermoscopy but represent an important criterion for diagnosis, especially when associated with polymorphic vessels or pink areas (Di Stefani et al. 2010). Although perpendicular white lines are not very specific for the diagnosis of NM, they are rarely seen in benign skin tumors (Navarrete-Dechent et al. 2016).

Even with these dermoscopic features in mind, NM may be missed. For this reason, specific management rules have been introduced, combining clinical and dermoscopic criteria (Lallas et al. 2013). Any nodular lesion that cannot be confidently diagnosed as benign requires immediate excision. Dermoscopic follow-up of a doubtful nodular lesion is strongly discouraged, because the rapid vertical growth of NM means any delay can worsen prognosis (Moscarella et al. 2017). In particular, pink lesions without a clear clinical and dermoscopic diagnosis should be treated with caution (Moscarella et al. 2017). Other nodular lesions that should always be excised are those with polymorphous vessels and any ulcerated nonpigmented nodule (Fig. 32) (Moscarella et al. 2017; Zalaudek et al. 2010). Suspected pyogenic granuloma may also require excision as it has overlapping clinical and dermoscopic features with nodular melanoma and can often only be distinguished with histopathology. Destructive methods such laser therapy and/or liquid nitrogen are strongly discouraged as they destroy tissue without the possibility of a histopathological examination.

Reflectance Confocal Microscopy Features of Nodular Melanoma RCM has significantly improved the early diagnosis of NM. NM usually lacks some of the confocal features of superficial spreading melanoma,

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

181

pigmented melanocytes under a significantly thinned but not ulcerated epidermis. Accordingly, black color may be even predictive of ulceration, which is a known negative prognostic criterion. The dermoscopic criterion of perpendicular white lines (shiny white streaks) appears to correlate to dermal fibrosis or collagen bundles under RCM (Pellacani et al. 2007a; Segura et al. 2008). Other RCM features relatively common in NM include the presence of enlarged deep vessels, which dermoscopically correspond to polymorphous vessels, and the presence of plump bright cells in the dermis which correspond to peppering on dermoscopy and to melanophages in histology (Segura et al. 2008; Waddell et al. 2018). Fig. 32 Ulcerated nodules, such as this pigmented ulcerated nodular melanoma, are highly suspicious for malignancy

such as epidermal disarrangement and pagetoid spreading, but has characteristic confocal features of its own (Fig. 33). In NM, massive proliferation in the dermis obliterates the typical papillary architecture, thus effacing the normal DEJ contours with no visible dermal papillae (Waddell et al. 2018). The basal layer and upper dermis show pleomorphic cells with bright cytoplasm and dark nuclei, while amorphous, hyporeflective nests, called cerebriform nests, are found in the deeper dermis (Segura et al. 2008). Cerebriform nests are very specific for NM although they are not always identified (Guitera et al. 2012; Pellacani et al. 2005). The dermoscopic blueblack rule has also been correlated with confocal features in NM (Longo et al. 2013a). Dermoscopic black color is associated with two different patterns: large black structureless areas (blotches) and irregular black dots/clods. Black structureless areas correspond to large, confluent areas of upwardly migrating melanocyte nests and pagetoid cells in the epidermis, whereas black dots/clods correspond to separate areas of upwardly migrating nests and pagetoid cells. Black color results not only from epidermal melanin or hemoglobin (in the case of ulceration) but also from a dense dermal proliferation of

Lentigo Maligna and Lentigo Maligna Melanoma Lentigo maligna (LM) is an in situ melanoma and lentigo maligna melanoma (LMM) an invasive melanoma, typically found on chronically sunexposed skin. LM/LMM appears mostly on the head and neck after the fourth decade of life, as a slow-growing asymmetrical macule with irregular borders. It is difficult to differentiate, even dermoscopically, from solar lentigo and early seborrheic keratosis, since all three types of lesions can present with typical melanoma features of multiple colors, asymmetry, irregular borders, and large size. Clinically it can also mimic a melanocytic nevus, lichen planus-like keratosis, and pigmented actinic keratosis (Lallas et al. 2014; Tanaka et al. 2011). Diagnosis is also more difficult because LM/LMM typically lack most classical dermoscopic signs of melanoma, due to the unusual structure of skin on the face with a matrix of hair and sweat follicles and flattened dermo-epidermal junction and rete ridges (Stolz et al. 2002).

Dermoscopic Features of Lentigo Maligna and Lentigo Maligna Melanoma The Schiffner progression model for LM/LMM describes the evolution of dermoscopic features with the increasing spread of melanoma cells (Fig. 34). The early signs of slate-gray dots,

182

K. J. Lee et al.

Fig. 33 Confocal characteristics of nodular melanoma. (a) Normal epidermis. (b) Atypia at the basal layer. (c) Cerebriform nests with plump, bright cells. (d) Cerebriform nests with vessels

increasing in size to clods, are formed by aggregates of melanin-loaded macrophages. Short lines are then formed by sheets of melanoma cells in the epidermis or upper dermis, and asymmetrical circles of pigmentation around hair follicles, often incomplete circles (crescent-shaped), form as melanoma cells descend unevenly around the hair follicle. As melanoma cells proliferate within the hair follicle and invade the adjacent dermis, streaks join up to form angulated or polygonal lines (zigzag pattern or rhomboid structures). Angulated/polygonal lines increase in thickness

until they form structureless zones with hair follicles still evident, followed by structureless zones with hair follicles obliterated (Fig. 35) (Schiffner et al. 2000). The most widely accepted diagnostic features of LM/LMM were proposed by Stolz et al., who listed primary criteria of dots and circles (annulargranular pattern), slate-gray dots and clods, incomplete pigmented circles around follicular openings, asymmetric changes over time, and the absence of seborrheic keratosis features such as yellow-brown opaque structureless areas, horn

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

Fig. 34 The Schiffner progression model for lentigo maligna and lentigo maligna melanoma. (a) Dots and then clods are formed by aggregates of melanin-loaded macrophages. Short lines and asymmetrical or incomplete circles of pigmentation around hair follicles are then formed by sheets of melanoma cells in the epidermis or

183

upper dermis. (b) Lines join up to form angulated or polygonal lines, which increase in thickness until they form structureless zones with hair follicles still evident, (c) followed by structureless zones with hair follicles obliterated. (Provided by Ralph P. Braun. CC BY-NC 4.0)

Fig. 35 (a) Brown and gray annular-granular structures joining up into incomplete circles around follicles in a lentigo maligna melanoma. (b) Angulated lines in a lentigo maligna melanoma

clods (pseudocysts), thin, curved parallel brown lines (fingerprint-like structures), and a sharply demarcated scalloped border (moth-eaten border) (Stolz et al. 2002). However, other researchers have added dermoscopic features to this list that may improve specificity and sensitivity. Schiffner et al. found that a combination of incomplete circles, dark brown or black polygonal lines (rhomboid structures), and slate-gray dots or clods gave a sensitivity of 89% and a specificity of 96% for LM/LMM in a series of 87 patients with pigmented facial lesions (Schiffner et al. 2000). In addition to the Stolz criteria, Pralong et

al. found that 125 cases of histopathologically confirmed LM/LMM frequently displayed increased vascular network density (58%), red polygonal lines (rhomboid structures) (40%), and target-like patterns with a dark dot in the center of a hyperpigmented hair follicle (41%) (Pralong et al. 2012). Annessi et al. proposed an algorithm considering irregularly distributed brown clods, a necklace pigment pattern consisting of a fragmented pigment lines with small dots on the lines, reticular lines with irregularly thickened lines, and thick brown or bluegray lines (ribbon-like structures). In a study of

184

167 doubtful pigmented lesions on the head, the presence of one or more of these dermoscopic features had a sensitivity of 99% and specificity of 83.9% for LM/LMM (Annessi et al. 2017). Areas of intense pigmentation and darkening on dermoscopy, where the lesion appears darker under dermoscopy than under clinical view, are another sign of LM/LMM. Many of these signs can occur individually in benign facial lesions; however, thick curved lines (cerebriform pattern); an opaque yellow-brown structureless zone; yellow or white clods corresponding to keratin plugs, scales, pseudocysts, white dots, and clods (milialike cysts); and yellow, brown, or black clods (comedo-like openings) are usually signs of a benign lesion (Table 5) (Annessi et al. 2017; Bollea-Garlatti et al. 2016; Lallas et al. 2016; Peris et al. 2016; Tanaka et al. 2011). An amelanotic LM/LMM can feature a pink or red homogenous area, pink or red structureless zone interrupted by follicular openings, dotted or irregular linear vessels, whitish structureless areas, or whitish radial lines, with or without traces of pigment. These structures can help distinguish it from clinical mimics such as actinic keratosis or superficial basal cell carcinoma (Giacomel et al. 2014). Although they occur mostly on the head and neck, 10% of LM/LMM can occur in other heavily sun-damaged areas of the body (Weigert and Stolz 2007). These lesions can mimic superficial spreading melanomas but often retain gray coloring, polygonal lines (rhomboid structures), angulated lines (zigzag lines), incomplete circles, obliteration of follicular openings, concentric circles, blue-gray dots and circles, or increased vasculature (Bollea-Garlatti et al. 2016; Jaimes et al. 2015; Martinez-Leborans et al. 2016; TiodorovicZivkovic et al. 2015).

Reflectance Confocal Microscopy Features of Lentigo Maligna and Lentigo Maligna Melanoma As with dermoscopy, LM/LMM features on RCM differ from other melanoma subtypes due to the anatomy of the face and the folliculotropism that occurs in LM/LMM. On RCM, the epidermis of LM/LMM is atrophic and characterized by the

K. J. Lee et al. Table 5 Features of lesions of the head and neck: lentigo maligna, solar lentigo, and seborrheic keratosis. Several features suggestive of lentigo maligna may also appear in benign lesions Dermoscopic appearance Features suggestive of lentigo maligna Gray dots and circles Gray dots arranged around (annular-granular pattern) a follicle Slate-gray dots or clods Irregularly distributed, slate-gray, round to oval, well-circumscribed aggregations of pigmentladen melanophages, over 0.1 mm in diameter Incomplete circles Dark brown, thick pigmentation distributed asymmetrically around the follicular openings, often as crescents Dark brown or black Thick brown or black lines angulated or polygonal between follicles, lines intersecting to form rhomboid or polygon shapes Red angulated or Similarly shaped to brown/ polygonal lines black angulated/polygonal lines, formed by increased vasculature between follicular openings Increased density of Thickened red lines of vasculature vessels Central brown dots A dark dot in the center of a (target-like or targetoid) hyperpigmented hair follicle Brown clods Irregularly distributed, brown, round to oval, wellcircumscribed pigment aggregations, over 0.1 mm in diameter Necklace lines Fragments of thin lines with small dark brown globules on the lines, resembling a necklace Reticular lines that are Fragments of pigment thick or that vary in color network with areas of hyperpigmentation and irregularly thickened lines and irregular meshes Dark brown or blue-gray Dark brown or blue-gray, thick lines (ribbon-like thick, linear, ribbon-like structures) structures that do not fill the whole interfollicular space; they may intersect to form zigzags or rhomboids Black structures Any type of black structure (continued)

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis Table 5 (continued) Dermoscopic appearance Darkening at dermoscopy The lesion appears darker under dermoscopic examination than under clinical examination Features specific to benign solar lentigo or seborrheic keratosis Thick, curved lines Brain-like pattern of (cerebriform pattern) brown, regularly thickened curved lines Thin, brown, curved A network of fine, light parallel lines (fingerprint- brown lines resembling a like pattern) fingerprint Opaque yellow-brown Yellow-brown, structureless zone homogenous, structureless pigmentation with an opaque surface Horn pseudocysts or Irregularly distributed and white dots and clods variously sized white or (milia-like cysts) yellow structures made of intraepithelial horn cysts Sharply demarcated, Irregular borders with scalloped border (mothragged large or small eaten border) concave indentations, similar to moth-eaten fragment Sharply demarcated The edge of the lesion is border (jelly sign) sharply demarcated, resembling the edge of a jelly

presence of large or round pagetoid melanocytes that may or may not disarray the epidermis (Fig. 36a). Dendritic pagetoid cells can also be seen in LM/LMM and can correspond to Langerhans cells, activated melanocytes, or atypical melanocytes. Therefore, the careful assessment of dendritic cells in lesions located on sun-exposed skin is crucial. When these dendritic cells are located at the edges of a fully fledged LM/LMM, especially if they are large, they can correspond to the trailing edge of a LM/LMM (Champin et al. 2014; Yélamos et al. 2017). This is very important when mapping the margins of LM/LMM prior to treatment. Therefore, it is key to evaluate the RCM features from the center to the periphery to put these findings in the right context (Guitera et al. 2013; Yélamos et al. 2017). In 2010, Guitera et al. described an algorithm to diagnose LM/LMM in which major and minor

185

criteria were taken into account (Table 6). A score of 2 points has a sensitivity of 85% and specificity of 76% for LM/LMM. However, this score was developed using the wide-probe VS1500 microscope at the center of the lesion. This score has also been validated with the handheld VS3000 microscope (Menge et al. 2016), with the addition of the presence of isolated atypical cells at the periphery of a lesion (Champin et al. 2014; Yélamos et al. 2017). Since the face tends to have a flattened DEJ with minimal elongation of the rete ridges, in RCM the visualization of edged papillae in normal skin can be more difficult. Hence, in LM/LMM the papillae are poorly defined (Guitera et al. 2010). The presence of large atypical melanocytes, both dendritic and round, surrounding the hair follicles is a key RCM feature for LM/LMM (Fig. 36b) and corresponds to the angulated lines, incomplete circles, and other signs of folliculotropism identified with dermoscopy. Initially, these atypical cells can be dispersed as individual cells, along the outer root sheath epithelium of the hair follicles and the basal layer of the epidermis. As the lesion progresses, the melanocytes become continuous along the basal layer and then develop elongated junctional nests of dendritic cells connected to the hair follicles (Hibler et al. 2017). These structures sometimes adopt the shape of a medusa head when located radially around a hair follicle (Fig. 36c) or adopt a mitochondrial-like appearance when located in parallel (Fig. 36d) (Gamo et al. 2016; Waddell et al. 2018). Another dermoscopic sign of folliculotropism, the circle within the circle, has been recently described to correspond on RCM to pigmented keratinocytes as well as pigmented atypical melanocytes, which results in increased pigmentation of the follicular epithelium as well as increased pigmentation in the rete ridges surrounding the hair follicle (Navarrete-Dechent et al. 2018). When evaluating the upper dermis of LM/ LMM, one can identify increased curled fibers corresponding to solar elastosis (Longo and Pellacani 2016), plump bright cells corresponding to dermal melanophages, as well as atypical nucleated cells within the papillae (Waddell et al. 2018). The latter are generally a sign of dermal invasion (Hibler et al. 2017).

186

K. J. Lee et al.

Fig. 36 Confocal characteristics of lentigo maligna and lentigo maligna melanoma. (a) Pleomorphic atypical epidermal cells, both round and dendritic. (b) Atypical cells at

the dermo-epidermal junction, perifollicular dendritic cells, plump cells, and bright cells. (c) Medusa head. (d) Mitochondria-like structures

Spitzoid Melanomas

pink-red nodules on the face or extremities of children but can also be found on adults. However, pigmented Spitz/Reed nevi are more common, presenting as well-circumscribed brown-black pigmented macules or plaques that are sometimes verrucous, composed of sharply circumscribed nests of pigmented spindle cells, melanophages, and dendritic melanocytes (Ferrara et al. 2013; Soyer et al. 2001b).

Spitzoid lesions exhibit particular unusual dermoscopic characteristics, and exist on a morphobiologic spectrum from frankly benign to benign with atypical features to frankly malignant, possibly reflecting an accumulation of genetic mutations within the lesion. Classical Spitz/Reed nevi present as well-circumscribed,

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis Table 6 Lentigo maligna scoring criteria, with additional features relevant to lentigo maligna and lentigo maligna melanoma Author Guitera et al. (2010)

Champin et al. (2014) Yélamos et al. (2017)

Criteria Major criteria: Non-edged papillae (+2) Round large pagetoid cells (+2) Minor criteria: Nucleated cells in dermal papillae (+1) Atypical cells at the DEJ (+1) Follicular localization of atypical cells (+1) Broadened honeycomb pattern (1) Single large round or large dendritic cell Atypical dendritic cell (any size) continuing from the trailing edge

Scoring 2

Interpretation Score 2 indicates a melanoma

2

1

1

1

1

Should be considered part of the tumor Should be considered part of the tumor

Dermoscopic Features of Spitzoid Melanomas Three main dermoscopic patterns are characteristic of Spitz/Reed nevi. The majority of Spitz/Reed nevi are pigmented: 51% have a starburst pattern, with a structureless blue-black center and symmetrically distributed radial lines or pseudopods at the periphery. These lines correspond to nests of spindle-shaped melanocytes arranged densely along the dermo-epidermal junction, often with melanophages in the papillary dermis below the nests of melanocytes (Lallas et al. 2017; Soyer et al. 2001b). Seventeen percent have reticular hypopigmented lines around brown clods (negative network); in other nevi, this pattern is considered to be strongly suggestive of melanoma. Here the pigmented clods correspond to pigmented nests of

187

spindle or large epithelioid cells. Other pigmented Spitz/Reed nevi have a globular, homogenous, reticular, or multicomponent pattern (Lallas et al. 2017; Soyer et al. 2001b). Multicomponent pattern consists of irregular or asymmetrically distributed clods, radial lines, or superficial reticular lines and may also have blue structureless zones (blue-white veil); these lesions are very suspicious for melanoma but on histopathological examination may be benign (Ferrara et al. 2013; Soyer et al. 2001b). Nineteen percent of Spitz/Reed nevi are nonpigmented and have a pattern of regularly distributed, monomorphic dotted vessels; a negative network can also exist in these lesions, with the depigmented reticular white lines surrounding a blood vessel. In nodular or raised lesions, the vessels may be seen as larger red clods or coiled, helical (corkscrew), or looped (hairpin) vessels (Lallas et al. 2017). Management of spitzoid lesions is complicated by the fact the some spitzoid melanomas are indistinguishable dermoscopically from Spitz/Reed nevi (Soyer et al. 2001b). Consensus guidelines by the International Dermoscopy Society (Lallas et al. 2017) state that spitzoid lesions should be managed with both the age of the patient and the dermoscopic features of the lesion in mind. Dermoscopically asymmetric lesions with spitzoid characteristics, whether flat or nodular, are suggestive of melanoma and should be excised, regardless of the patient’s age. In addition, any new spitzoid lesion appearing after the age of 12, even when symmetrical, should be excised or closely monitored for dermoscopic changes. For patients under 12 years of age, flat, symmetrical spitzoid lesions should be monitored and excised if there are asymmetric changes; lesions with symmetric changes should be followed up until there has been no new changes or growth for 6 months.

Reflectance Confocal Microscopy Features of Spitzoid Melanomas Since Spitz/Reed nevus and spitzoid melanomas can present clinically with the same patterns, RCM has been used to help distinguish spitzoid lesions sharing the same dermoscopic patterns. Initially, Pellacani et al. described Spitz/Reed

188

nevi that presented on RCM with a sharp border, junctional nests, and melanophages (Pellacani et al. 2007b). Later, Guida et al. performed a retrospective study evaluating the RCM features of 34 Spitz/Reed nevi and spitzoid melanomas sharing the same dermoscopic patterns (Guida et al. 2016). They concluded that the features more commonly associated with spitzoid melanoma were (1) the presence of very marked cell pleomorphism in the epidermis, (2) very marked pleomorphism within nests, and (3) atypical cells widespread throughout the DEJ of the entire lesion. However, RCM results when assessing spitzoid lesions should be treated with caution since important histologic features to distinguish Spitz/Reed nevi from melanomas, such as dermal maturation or the present of deep mitoses, cannot be evaluated with RCM.

Desmoplastic Melanomas Desmoplastic melanoma (DM) is a type of invasive melanoma characterized by a fibrocollagenous stroma with sparsely distributed spindle cells. They are easily mistaken clinically for benign lesions or other skin cancer types, and there is an overlying melanoma in situ component or atypical melanocytic hyperplasia in three quarters of desmoplastic melanomas. DM are commonly located in the head/neck region, and the in situ component is usually lentigo maligna. Clinically, multiple colors are common, particularly pink, red, brown, or white, and DM frequently have either a papular or nodular component or are entirely nodular. Poorly defined borders are another common feature (Jaimes et al. 2013; Maher et al. 2017).

Dermoscopic Features of Desmoplastic Melanomas Common dermoscopic patterns in DM include typical and atypical reticular and globular patterns and a structureless pattern interrupted by follicular openings (pseudonetwork pattern) on the face. White lines perpendicular to each (crystalline structures) other appear in 80% of DM, and atypical and polymorphous vascular structures are

K. J. Lee et al.

present in over 80% of DM, including dotted, serpentine (linear irregular), or coiled (glomerular) vessels and pink structureless zones. Other common dermoscopic features of DM are gray dots (peppering), gray dots with gray circles (annular-granular pattern), incomplete circles (asymmetrically pigmented follicular openings), blue structureless zones (blue-white veil), irregularly distributed clods, reticular lines that vary in thickness or color (atypical network), white structureless zones (scar-like areas), off-center dark structureless zones, and peripheral light brown structureless zones. Negative network, radial lines (streaks), polygonal lines, and structureless zones with obliterated follicles are occasionally seen in DM (Jaimes et al. 2013; Maher et al. 2017).

Reflectance Confocal Microscopy Features of Desmoplastic Melanomas Confocal features that are common in superficial spreading melanomas are also found in DM. Pagetoid cells, cellular atypia, and nucleated cells in the dermis are all found in the majority of DM. A study of 37 DM found that the Modena algorithm (Pellacani et al. 2007b) had a 97% sensitivity for DM. However, several confocal features are DM-specific. Spindle cells in the superficial dermis, inflammation in the dermis, and vertical vessels in the papillae are more common in DM. Round pagetoid cells are less common in DM than in other invasive melanomas (Maher et al. 2017).

Conclusion The development of noninvasive diagnostic tools such as dermoscopy and RCM, and the corresponding improvement in diagnostic accuracy, has marked an important step forward in melanoma diagnosis. Dermoscopy in particular has been shown to be useful to primary care physicians and dermatologists after a short training course, while RCM allows very detailed examination of equivocal lesions in sensitive areas and mapping the edges of lesions prior to excision, reducing the number of excisions required.

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis

Ongoing research with these tools has revealed in detail the most common dermoscopic and confocal features of melanoma while also determining the features of difficult-to-diagnose lesions such as nodular melanoma, spitzoid melanomas, or lesions on unusual body sites such as facial or acral skin.

Cross-References ▶ Acral Lentiginous Melanoma ▶ Classification and Histopathology of Melanoma ▶ Clinical Presentations of Melanoma ▶ Melanoma Prevention and Screening ▶ Mucosal Melanoma

References Alarcon I, Carrera C, Alos L, Palou J, Malvehy J, Puig S (2014a) In vivo reflectance confocal microscopy to monitor the response of lentigo maligna to imiquimod. J Am Acad Dermatol 71:49–55. https://doi.org/10.1016/ j.jaad.2014.02.043 Alarcon I, Carrera C, Palou J, Alos L, Malvehy J, Puig S (2014b) Impact of in vivo reflectance confocal microscopy on the number needed to treat melanoma in doubtful lesions. Br J Dermatol 170:802–808. https://doi.org/ 10.1111/bjd.12678 Annessi G, Bono R, Abeni D (2017) Correlation between digital epiluminescence microscopy parameters and histopathological changes in lentigo maligna and solar lentigo: a dermoscopic index for the diagnosis of lentigo maligna. J Am Acad Dermatol 76:234–243. https://doi.org/10.1016/j.jaad.2016.08.032 Argenziano G, Soyer HP (2001) Dermoscopy of pigmented skin lesions – a valuable tool for early diagnosis of melanoma. Lancet Oncol 2:443–449 Argenziano G, Fabbrocini G, Carli P, De Giorgi V, Sammarco E, Delfino M (1998) Epiluminescence microscopy for the diagnosis of doubtful melanocytic skin lesions. Comparison of the ABCD rule of dermatoscopy and a new 7-point checklist based on pattern analysis. Arch Dermatol 134:1563–1570 Argenziano G et al (2003) Dermoscopy of pigmented skin lesions: results of a consensus meeting via the Internet. J Am Acad Dermatol 48:679–693. https://doi.org/ 10.1067/mjd.2003.281 Argenziano G et al (2006) Dermoscopy improves accuracy of primary care physicians to triage lesions suggestive of skin cancer. J Clin Oncol: Off J Am Soc Clin Oncol 24:1877–1882. https://doi.org/10.1200/jco.2005. 05.0864

189

Argenziano G et al (2011a) Seven-point checklist of dermoscopy revisited. Br J Dermatol 164:785–790. https://doi.org/10.1111/j.1365-2133.2010.10194.x Argenziano G et al (2011b) Blue-black rule: a simple dermoscopic clue to recognize pigmented nodular melanoma. Br J Dermatol 165:1251–1255. https://doi.org/ 10.1111/j.1365-2133.2011.10621.x Argenziano G et al (2012) Accuracy in melanoma detection: a 10-year multicenter survey. J Am Acad Dermatol 67:54–59. https://doi.org/10.1016/j.jaad.2011.07.019 Bollea-Garlatti LA, Galimberti GN, Galimberti RL (2016) Lentigo Maligna: keys to dermoscopic diagnosis. Actas Dermo-sifiliogr 107:489–497. https://doi.org/10.1016/ j.ad.2016.01.001 Borsari S et al (2016) Clinical indications for use of reflectance confocal microscopy for skin cancer diagnosis. JAMA Dermatol. https://doi.org/10.1001/jamadermatol. 2016.1188 Braga JC, Scope A, Klaz I, Mecca P, Gonzalez S, Rabinovitz H, Marghoob AA (2009) The significance of reflectance confocal microscopy in the assessment of solitary pink skin lesions. J Am Acad Dermatol 61:230–241. https:// doi.org/10.1016/j.jaad.2009.02.036 Braun RP, Saurat JH, French LE (2004) Dermoscopy of pigmented lesions: a valuable tool in the diagnosis of melanoma. Swiss Med Wkly 134:83–90. Buhl T et al (2012) Integrating static and dynamic features of melanoma: the DynaMel algorithm. J Am Acad Dermatol 66:27–36. https://doi.org/10.1016/j.jaad.2010.09.731 Carrera C, Marghoob AA (2016) Discriminating nevi from melanomas: clues and pitfalls. Dermatol Clin 34:395–409. https://doi.org/10.1016/j.det.2016.05.003 Carrera C et al (2016) Validity and reliability of dermoscopic criteria used to differentiate nevi from melanoma: a web-based international dermoscopy society study. JAMA Dermatol 152:798–806. https:// doi.org/10.1001/jamadermatol.2016.0624 Carrera C et al (2017a) Dermoscopic clues for diagnosing melanomas that resemble seborrheic keratosis. JAMA Dermatol 153:544–551. https://doi.org/10.1001/jamader matol.2017.0129 Carrera C et al (2017b) Dermoscopic clues for diagnosing melanomas that resemble seborrheic keratosis. JAMA Dermatol. https://doi.org/10.1001/jamadermatol. 2017.0129 Chamberlain AJ, Kelly JW (2007) Use of dermoscopy in Australia. Med J Aust 187:252–253 Chamberlain AJ, Fritschi L, Giles GG, Dowling JP, Kelly JW (2002) Nodular type and older age as the most significant associations of thick melanoma in Victoria, Australia. Arch Dermatol 138:609–614 Chamberlain AJ, Fritschi L, Kelly JW (2003) Nodular melanoma: patients’ perceptions of presenting features and implications for earlier detection. J Am Acad Dermatol 48:694–701. https://doi.org/10.1067/mjd.2003.216 Champin J et al (2014) In vivo reflectance confocal microscopy to optimize the spaghetti technique for defining surgical margins of lentigo maligna. Dermatol Surg: Off Publ Am Soc Dermatol Surg

190 [et al] 40:247–256. https://doi.org/10.1111/dsu. 12432 Cheng HM, Lo S, Scolyer R, Meekings A, Carlos G, Guitera P (2016) Accuracy of optical coherence tomography for the diagnosis of superficial basal cell carcinoma: a prospective, consecutive, cohort study of 168 cases. Br J Dermatol 175:1290–1300. https://doi.org/ 10.1111/bjd.14714 Clark WH Jr, From L, Bernardino EA, Mihm MC (1969) The histogenesis and biologic behavior of primary human malignant melanomas of the skin. Cancer Res 29:705–727 Di Stefani A, Campbell TM, Malvehy J, Massone C, Soyer HP, Hofmann-Wellenhof R (2010) Shiny white streaks: an additional dermoscopic finding in melanomas viewed using contact polarised dermoscopy. Australas J Dermatol 51:295–298. https://doi.org/10.1111/j.14400960.2010.00672.x Edwards SJ, Mavranezouli I, Osei-Assibey G, Marceniuk G, Wakefield V, Karner C (2016) VivaScope(R) 1500 and 3000 systems for detecting and monitoring skin lesions: a systematic review and economic evaluation. Health Technol Assess (Winchester, England) 20:1–260. https://doi.org/10.3310/hta20580 Esteva A, Kuprel B, Novoa RA, Ko J, Swetter SM, Blau HM, Thrun S (2017) Dermatologist-level classification of skin cancer with deep neural networks. Nature 542:115–118. https://doi.org/10.1038/nature21056 Ferrara G, Gianotti R, Cavicchini S, Salviato T, Zalaudek I, Argenziano G (2013) Spitz nevus, Spitz tumor, and spitzoid melanoma: a comprehensive clinicopathologic overview. Dermatol Clin 31:589–598, viii. https://doi. org/10.1016/j.det.2013.06.012 Forsea AM, Tschandl P, Zalaudek I, Del Marmol V, Soyer HP, Argenziano G, Geller AC (2017) The impact of dermoscopy on melanoma detection in the practice of dermatologists in Europe: results of a pan-European survey. J Eur Acad Dermatol Venereol: JEADV 31:1148–1156. https://doi.org/10.1111/jdv.14129 Fraga-Braghiroli NA, Stephens A, Grossman D, Rabinovitz H, Castro RP, Scope A (2014) Use of handheld reflectance confocal microscopy for in vivo diagnosis of solitary facial papules: a case series. J Eur Acad Dermatol Venereol: JEADV 28:933–942. https://doi. org/10.1111/jdv.12218 Gambichler T et al (2015a) High-definition optical coherence tomography of melanocytic skin lesions. J Biophotonics 8:681–686. https://doi.org/10.1002/jbio.201400085 Gambichler T, Schmid-Wendtner MH, Plura I, Kampilafkos P, Stucker M, Berking C, Maier T (2015b) A multicentre pilot study investigating high-definition optical coherence tomography in the differentiation of cutaneous melanoma and melanocytic naevi. J Eur Acad Dermatol Venereol: JEADV 29:537–541. https://doi.org/10.1111/ jdv.12621 Gamo R, Pampin A, Floristan U (2016) Reflectance confocal microscopy in Lentigo Maligna. Actas Dermosifiliogr 107:830–835. https://doi.org/10.1016/j.ad.2016.07.012

K. J. Lee et al. Giacomel J, Lallas A, Zalaudek I, Argenziano G (2014) Dermoscopic “signature” pattern of pigmented and nonpigmented lentigo maligna. J Am Acad Dermatol 70: e33–e35. https://doi.org/10.1016/j.jaad.2013.08.045 Gualdi G, Venturini M, Zanca A, Calzavara-Pinton PG, Pellacani G (2016) Pre-surgical basal cell carcinoma margin definition: the SMART approach. J Eur Acad Dermatol Venereol: JEADV 30:474–476. https://doi. org/10.1111/jdv.12858 Guida S et al (2016) Spitz naevi and melanomas with similar dermoscopic patterns: can confocal microscopy differentiate? Br J Dermatol 174:610–616. https://doi. org/10.1111/bjd.14286 Guitera P et al (2010) The impact of in vivo reflectance confocal microscopy on the diagnostic accuracy of lentigo maligna and equivocal pigmented and nonpigmented macules of the face. J Invest Dermatol 130:2080–2091. https://doi.org/10.1038/jid.2010.84 Guitera P, Menzies SW, Longo C, Cesinaro AM, Scolyer RA, Pellacani G (2012) In vivo confocal microscopy for diagnosis of melanoma and basal cell carcinoma using a two-step method: analysis of 710 consecutive clinically equivocal cases. J Invest Dermatol 132:2386–2394. https://doi.org/10.1038/jid.2012.172 Guitera P et al (2013) Improving management and patient care in lentigo maligna by mapping with in vivo confocal microscopy. JAMA Dermatol 149:692–698. https://doi.org/10.1001/jamadermatol.2013.2301 Guitera P et al (2016) Dermoscopy and in vivo confocal microscopy are complimentary techniques for the diagnosis of difficult amelanotic and light colored skin lesions. Br J Dermatol. https://doi.org/10.1111/bjd.14749 Haenssle HA et al (2015) Melanoma thickness: the role of patients’ characteristics, risk indicators and patterns of diagnosis. J Eur Acad Dermatol Venereol: JEADV 29:102–108. https://doi.org/10.1111/jdv.12471 Haenssle HA et al (2018) Man against machine: diagnostic performance of a deep learning convolutional neural network for dermoscopic melanoma recognition in comparison to 58 dermatologists. Ann Oncol: Off J Eur Soc Med Oncol 29:1836–1842. https://doi.org/ 10.1093/annonc/mdy166 Halpern AC, Marchetti MA, Marghoob AA (2014) Melanoma surveillance in “high-risk” individuals. JAMA Dermatol 150:815–816. https://doi.org/10.1001/jama dermatol.2014.513 Henning JS, Dusza SW, Wang SQ, Marghoob AA, Rabinovitz HS, Polsky D, Kopf AW (2007) The CASH (color, architecture, symmetry, and homogeneity) algorithm for dermoscopy. J Am Acad Dermatol 56:45–52. https://doi.org/10.1016/j.jaad.2006.09.003 Hibler BP, Cordova M, Wong RJ, Rossi AM (2015) Intraoperative real-time reflectance confocal microscopy for guiding surgical margins of lentigo maligna melanoma. Dermatol Surg: Off Publ Am Soc Dermatol Surg [et al] 41:980–983. https://doi.org/10.1097/dss. 0000000000000401 Hibler B, Yélamos O, Cordova M, Sierra H, Rajadhyaksha M, Nehal K, Rossi A (2017) Handheld reflectance

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis confocal microscopy to aid in the management of complex facial lentigo maligna. Cutis 99:346–352 Hofmann-Wellenhof R, Blum A, Wolf IH, Piccolo D, Kerl H, Garbe C, Soyer HP (2001) Dermoscopic classification of atypical melanocytic nevi (Clark nevi). Arch Dermatol 137:1575–1580 Hubner J et al (2017) Interval cancers after skin cancer screening: incidence, tumour characteristics and risk factors for cutaneous melanoma. Br J Cancer 116:253–259. https://doi.org/10.1038/bjc.2016.390 Jaimes N et al (2013) Clinical and dermoscopic characteristics of desmoplastic melanomas. JAMA Dermatol 149:413–421. https://doi.org/10.1001/jamadermatol.20 13.2248 Jaimes N et al (2015) Clinical and dermoscopic characteristics of melanomas on nonfacial chronically sun-damaged skin. J Am Acad Dermatol 72:1027–1035. https://doi.org/ 10.1016/j.jaad.2015.02.1117 Kalkhoran S, Milne O, Zalaudek I, Puig S, Malvehy J, Kelly JW, Marghoob AA (2010) Historical, clinical, and dermoscopic characteristics of thin nodular melanoma. Arch Dermatol 146:311–318. https://doi.org/ 10.1001/archdermatol.2009.369 Kelly JW, Chamberlain AJ, Staples MP, McAvoy B (2003) Nodular melanoma. No longer as simple as ABC. Aust Fam Physician 32:706–709 Kittler H, Pehamberger H, Wolff K, Binder M (2002) Diagnostic accuracy of dermoscopy. Lancet Oncol 3:159–165 Kittler H et al (2016) Standardization of terminology in dermoscopy/dermatoscopy: results of the third consensus conference of the International Society of Dermoscopy. J Am Acad Dermatol 74:1093–1106. https://doi.org/10.1016/j.jaad.2015.12.038 Kose K et al (2017) Automated video-mosaicking approach for confocal microscopic imaging in vivo: an approach to address challenges in imaging living tissue and extend field of view. Sci Rep 7:10759. https://doi.org/10.1038/s41598-017-11072-9 Lallas A et al (2013) Management rules to detect melanoma. Dermatology (Basel, Switzerland) 226:52–60. https://doi.org/10.1159/000346645 Lallas A, Argenziano G, Moscarella E, Longo C, Simonetti V, Zalaudek I (2014) Diagnosis and management of facial pigmented macules. Clin Dermatol 32:94–100. https://doi.org/10.1016/j.clindermatol.2013.05.030 Lallas A et al (2016) Dermoscopic clues to differentiate facial lentigo maligna from pigmented actinic keratosis. Br J Dermatol 174:1079–1085. https://doi.org/ 10.1111/bjd.14355 Lallas A et al (2017) Update on dermoscopy of Spitz/Reed naevi and management guidelines by the International Dermoscopy Society. Br J Dermatol 177:645–655. https://doi.org/10.1111/bjd.15339 Lipsker D, Engel F, Cribier B, Velten M, Hedelin G (2007) Trends in melanoma epidemiology suggest three different types of melanoma. Br J Dermatol 157:338–343. https://doi.org/10.1111/j.1365-2133. 2007.08029.x

191

Liu W, Dowling JP, Murray WK, McArthur GA, Thompson JF, Wolfe R, Kelly JW (2006) Rate of growth in melanomas: characteristics and associations of rapidly growing melanomas. Arch Dermatol 142:1551–1558. https://doi.org/10.1001/archderm.142.12.1551 Longo C, Pellacani G (2016) Melanomas. Dermatol Clin 34:411–419. https://doi.org/10.1016/j.det.2016.05.004 Longo C et al (2013a) Is confocal microscopy a valuable tool in diagnosing nodular lesions? A study of 140 cases. Br J Dermatol 169:58–67. https://doi.org/ 10.1111/bjd.12259 Longo C et al (2013b) Can noninvasive imaging tools potentially predict the risk of ulceration in invasive melanomas showing blue and black colors? Melanoma Res 23:125–131. https://doi.org/10.1097/CMR.0b013 e32835d90b8 Maher NG, Solinas A, Scolyer RA, Puig S, Pellacani G, Guitera P (2017) Detection of desmoplastic melanoma with dermoscopy and reflectance confocal microscopy. J Eur Acad Dermatol Venereol: JEADV 31:2016–2024. https://doi.org/10.1111/jdv.14381 Malvehy J, Puig S (2007) Melanoma of the trunk and limbs including superficial and nodular melanoma. In: Soyer HP, Argenziano G, Hofmann-Wellenhof R, Johr R (eds) Color atlas of melanocytic lesions of the skin. Springer, Berlin, pp 237–259 Malvehy J et al (2014) Clinical performance of the Nevisense system in cutaneous melanoma detection: an international, multicentre, prospective and blinded clinical trial on efficacy and safety. Br J Dermatol 171:1099–1107. https://doi.org/10.1111/bjd.13121 Mar V, Roberts H, Wolfe R, English DR, Kelly JW (2013) Nodular melanoma: a distinct clinical entity and the largest contributor to melanoma deaths in Victoria, Australia. J Am Acad Dermatol 68:568–575. https:// doi.org/10.1016/j.jaad.2012.09.047 Marchetti MA et al (2018) Results of the 2016 international skin imaging collaboration international symposium on biomedical imaging challenge: comparison of the accuracy of computer algorithms to dermatologists for the diagnosis of melanoma from dermoscopic images. J Am Acad Dermatol 78:270–277 e271. https://doi. org/10.1016/j.jaad.2017.08.016 Martinez-Leborans L, Garcias-Ladaria J, Oliver-Martinez V, Alegre de Miquel V (2016) Extrafacial Lentigo Maligna: a report on 14 cases and a review of the literature. Actas Dermo-sifiliogr 107:e57–e63. https:// doi.org/10.1016/j.ad.2015.10.018 Menge TD, Hibler BP, Cordova MA, Nehal KS, Rossi AM (2016) Concordance of handheld reflectance confocal microscopy (RCM) with histopathology in the diagnosis of lentigo maligna (LM): a prospective study. J Am Acad Dermatol. https://doi.org/10.1016/j.jaad.2015.12.045 Menzies SW, Ingvar C, Crotty KA, McCarthy WH (1996) Frequency and morphologic characteristics of invasive melanomas lacking specific surface microscopic features. Arch Dermatol 132:1178–1182 Menzies SW et al (2008) Dermoscopic evaluation of amelanotic and hypomelanotic melanoma. Arch

192 Dermatol 144:1120–1127. https://doi.org/10.1001/arch derm.144.9.1120 Monheit G et al (2011) The performance of MelaFind: a prospective multicenter study. Arch Dermatol 147:188–194. https://doi.org/10.1001/archdermatol.2 010.302 Moscarella E et al (2017) Performance of the “if in doubt, cut it out” rule for the management of nodular melanoma. Dermatol Pract Concept 7:1–5. https://doi.org/ 10.5826/dpc.0703a01 Moynihan GD (1994) The 3 Cs of melanoma: time for a change? J Am Acad Dermatol 30:510–511 Murray CS, Stockton DL, Doherty VR (2005) Thick melanoma: the challenge persists. Br J Dermatol 152:104–109. https://doi.org/10.1111/j.1365-2133.200 5.06409.x Nachbar F et al (1994) The ABCD rule of dermatoscopy. High prospective value in the diagnosis of doubtful melanocytic skin lesions. J Am Acad Dermatol 30:551–559 Navarrete-Dechent C, Bajaj S, Marchetti MA, Rabinovitz H, Dusza SW, Marghoob AA (2016) Association of shiny white blotches and strands with nonpigmented basal cell carcinoma: evaluation of an additional dermoscopic diagnostic criterion. JAMA Dermatol 152:546–552. https://doi.org/10.1001/jama dermatol.2015.5731 Navarrete-Dechent C, Liopyris K, Cordova M, Busam K, Marghoob A, Chen CJ (2018) Reflectance confocal microscopic and En face histopathologic correlation of the Dermoscopic “circle within a circle” in Lentigo Maligna. JAMA Dermatol. Published online 25 July 2018. https://doi.org/10.1001/jamadermatol.2018.2216 Pehamberger H, Steiner A, Wolff K (1987) In vivo epiluminescence microscopy of pigmented skin lesions. I. Pattern analysis of pigmented skin lesions. J Am Acad Dermatol 17:571–583 Pellacani G, Cesinaro AM, Seidenari S (2005) Reflectancemode confocal microscopy of pigmented skin lesions – improvement in melanoma diagnostic specificity. J Am Acad Dermatol 53:979–985. https://doi.org/10.1016/j. jaad.2005.08.022 Pellacani G, Bassoli S, Longo C, Cesinaro AM, Seidenari S (2007a) Diving into the blue: in vivo microscopic characterization of the dermoscopic blue hue. J Am Acad Dermatol 57:96–104. https://doi.org/10.1016/j. jaad.2006.11.006 Pellacani G, Guitera P, Longo C, Avramidis M, Seidenari S, Menzies S (2007b) The impact of in vivo reflectance confocal microscopy for the diagnostic accuracy of melanoma and equivocal melanocytic lesions. J Invest Dermatol 127:2759–2765. https://doi.org/10.1038/sj. jid.5700993 Peris K, Maiorino A, Di Stefani A, Longo C, Piana S, Argenziano G (2016) Brown globules in lentigo maligna (LM): a useful dermoscopic clue. J Am Acad Dermatol 75:429–430. https://doi.org/10.1016/j.jaad.2016.02.1231 Pizzichetta MA et al (2013) Negative pigment network: an additional dermoscopic feature for the diagnosis of

K. J. Lee et al. melanoma. J Am Acad Dermatol 68:552–559. https:// doi.org/10.1016/j.jaad.2012.08.012 Pizzichetta MA et al (2015) Pigmented nodular melanoma: the predictive value of dermoscopic features using multivariate analysis. Br J Dermatol 173:106–114. https://doi.org/10.1111/bjd.13861 Pizzichetta MA et al (2017) Dermoscopic diagnosis of amelanotic/hypomelanotic melanoma. Br J Dermatol 177:538–540. https://doi.org/10.1111/bjd.15093 Pralong P, Bathelier E, Dalle S, Poulalhon N, Debarbieux S, Thomas L (2012) Dermoscopy of lentigo maligna melanoma: report of 125 cases. Br J Dermatol 167:280–287. https://doi.org/10.1111/j.1365-2133.2012.10932.x Que SK, Grant-Kels JM, Longo C, Pellacani G (2016) Basics of confocal microscopy and the complexity of diagnosing skin tumors: new imaging tools in clinical practice, diagnostic workflows, cost-estimate, and new trends. Dermatol Clin 34:367–375. https://doi.org/ 10.1016/j.det.2016.05.001 Rajadhyaksha M, Grossman M, Esterowitz D, Webb RH, Anderson RR (1995) In vivo confocal scanning laser microscopy of human skin: melanin provides strong contrast. J Invest Dermatol 104:946–952 Rajadhyaksha M, Gonzalez S, Zavislan JM, Anderson RR, Webb RH (1999) In vivo confocal scanning laser microscopy of human skin II: advances in instrumentation and comparison with histology. J Invest Dermatol 113:293–303. https://doi.org/10.1046/j.1523-1747.1999. 00690.x Rajadhyaksha M, Marghoob A, Rossi A, Halpern AC, Nehal KS (2016) Reflectance confocal microscopy of skin in vivo: from bench to bedside. Lasers Surg Med. https://doi.org/10.1002/lsm.22600 Rayner JE, Laino AM, Nufer KL, Adams L, Raphael AP, Menzies SW, Soyer HP (2018) Clinical perspective of 3D total body photography for early detection and screening of melanoma. Front Med 5:152. https://doi. org/10.3389/fmed.2018.00152 Ribero S, Moscarella E, Ferrara G, Piana S, Argenziano G, Longo C (2016) Regression in cutaneous melanoma: a comprehensive review from diagnosis to prognosis. J Eur Acad Dermatol Venereol : JEADV 30:2030–2037. https://doi.org/10.1111/jdv.13815 Rogers T, Marino ML, Dusza SW, Bajaj S, Usatine RP, Marchetti MA, Marghoob AA (2016) A clinical aid for detecting skin cancer: the triage amalgamated dermoscopic algorithm (TADA). J Am Board Fam Med : JABFM 29:694–701. https://doi.org/10.3122/ jabfm.2016.06.160079 Rosendahl C, Cameron A, McColl I, Wilkinson D (2012) Dermatoscopy in routine practice – ‘chaos and clues’. Aust Fam Physician 41:482–487 Salerni G et al (2012a) Characterization of 1152 lesions excised over 10 years using total-body photography and digital dermatoscopy in the surveillance of patients at high risk for melanoma. J Am Acad Dermatol 67:836–845. https://doi.org/10.1016/j.jaad.2012.01.028 Salerni G et al (2012b) Benefits of total body photography and digital dermatoscopy (“two-step method of digital

Dermoscopy/Confocal Microscopy for Melanoma Diagnosis follow-up”) in the early diagnosis of melanoma in patients at high risk for melanoma. J Am Acad Dermatol 67:e17–e27. https://doi.org/10.1016/j.jaad.2011.04.008 Salerni G, Teran T, Puig S, Malvehy J, Zalaudek I, Argenziano G, Kittler H (2013) Meta-analysis of digital dermoscopy follow-up of melanocytic skin lesions: a study on behalf of the international dermoscopy society. J Eur Acad Dermatol Venereol: JEADV 27:805–814. https://doi.org/10.1111/jdv.12032 Salerni G, Teran T, Alonso C, Fernandez-Bussy R (2014) The role of dermoscopy and digital dermoscopy follow-up in the clinical diagnosis of melanoma: clinical and dermoscopic features of 99 consecutive primary melanomas. Dermatol Pract Conceptual 4:39–46. https://doi.org/10.5826/dpc.0404a07 Schiffner R, Schiffner-Rohe J, Vogt T, Landthaler M, Wlotzke U, Cognetta AB, Stolz W (2000) Improvement of early recognition of lentigo maligna using dermatoscopy. J Am Acad Dermatol 42:25–32 Scope A et al (2006) Predominant dermoscopic patterns observed among nevi. J Cutan Med Surg 10:170–174. https://doi.org/10.2310/7750.2006.00045 Scope A et al (2008) The “ugly duckling” sign: agreement between observers. Arch Dermatol 144:58–64. https:// doi.org/10.1001/archdermatol.2007.15 Segura S et al (2008) In vivo microscopic features of nodular melanomas: dermoscopy, confocal microscopy, and histopathologic correlates. Arch Dermatol 144:1311–1320. https://doi.org/10.1001/archderm.144.10.1311 Segura S, Puig S, Carrera C, Palou J, Malvehy J (2009) Development of a two-step method for the diagnosis of melanoma by reflectance confocal microscopy. J Am Acad Dermatol 61:216–229. https://doi.org/10.1016/j. jaad.2009.02.014 Soyer HP, Argenziano G, Chimenti S, Ruocco V (2001a) Dermoscopy of pigmented skin lesions. Eur J Dermatol: EJD 11:270–276; quiz 277 Soyer HP, Argenziano G, Ruocco V, Chimenti S (2001b) Dermoscopy of pigmented skin lesions (part II). Eur J Dermatol: EJD 11:483–498 Stolz W, Schiffner R, Burgdorf WH (2002) Dermatoscopy for facial pigmented skin lesions. Clin Dermatol 20:276–278 Tanaka M, Sawada M, Kobayashi K (2011) Key points in dermoscopic differentiation between lentigo maligna and solar lentigo. J Dermatol 38:53–58. https://doi. org/10.1111/j.1346-8138.2010.01132.x Tiodorovic-Zivkovic D et al (2015) Age, gender, and topography influence the clinical and dermoscopic appearance of lentigo maligna. J Am Acad Dermatol 72:801–808. https://doi.org/10.1016/j.jaad.2015.01.030 Ulrich M et al (2016) Dynamic optical coherence tomography in dermatology. Dermatology (Basel, Switzerland) 232:298–311. https://doi.org/10.1159/000444706 Vestergaard ME, Macaskill P, Holt PE, Menzies SW (2008) Dermoscopy compared with naked eye examination for the diagnosis of primary melanoma: a meta-analysis of studies performed in a clinical setting. Brit J Dermatol

193

159:669–676. https://doi.org/10.1111/j.1365-2133.200 8.08713.x Waddell A, Star P, Guitera P (2018) Advances in the use of reflectance confocal microscopy in melanoma. Melanoma Manag. Epub ahead of print. https://doi.org/ 10.2217/mmt-2018-0001 Wang SQ, Marghoob AA, Scope A (2012) Principles of dermoscopy and dermoscopic equipment. In: Marghoob AA, Malvehy J, Braun RP (eds) Atlas of dermoscopy, 2nd edn. Informa Healthcare, London, pp 3–9 Wang W, Zhao J, Short M, Zeng H (2015) Real-time in vivo cancer diagnosis using Raman spectroscopy. J Biophotonics 8:527–545. https://doi.org/10.1002/ jbio.201400026 Warycha MA et al (2008) Changes in the presentation of nodular and superficial spreading melanomas over 35 years. Cancer 113:3341–3348. https://doi.org/10.1002/ cncr.23955 Weigert U, Stolz W (2007) Melanoma of the face. In: Soyer HP, Argenziano G, Hofmann-Wellenhof R, Johr R (eds) Color atlas of melanocytic lesions of the skin. Spinger, Berlin, pp 233–236 Welzel J, Schuh S (2017) Noninvasive diagnosis in dermatology. J Dtsch Dermatol Ges = J Ger Soc Dermatol: JDDG 15:999–1016. https://doi.org/10.1111/ddg.13347 Wolner ZJ, Yelamos O, Liopyris K, Rogers T, Marchetti MA, Marghoob AA (2017) Enhancing skin cancer diagnosis with Dermoscopy. Dermatol Clin 35:417–437. https://doi.org/10.1016/j.det.2017.06.003 Xiong YQ, Ma SJ, Mo Y, Huo ST, Wen YQ, Chen Q (2017) Comparison of dermoscopy and reflectance confocal microscopy for the diagnosis of malignant skin tumours: a meta-analysis. J Cancer Res Clin Oncol 143:1627–1635. https://doi.org/10.1007/s00432-0172391-9 Yélamos O et al (2017) Correlation of handheld reflectance confocal microscopy with radial video mosaicing for margin mapping of lentigo maligna and lentigo maligna melanoma. JAMA Dermatol 153:1278–1284. https:// doi.org/10.1001/jamadermatol.2017.3114 Yélamos O, Braun R, Liopyris K, Wolner ZJ, Kerl K, Gerami P, Marghoob A (2018a) Dermoscopy/ dermatoscopy and dermatopathology correlates of cutaneous neoplasms. J Am Acad Dermatol 80:341–363. https://doi.org/10.1016/j.jaad.2018.07.073 Yélamos O et al (2018b) Impact of reflectance confocal microscopy on the physician’s diagnostic confidence and decision making of equivocal skin tumors: outcomes after 1 year follow-up Poster 6903 presented at the 2018 American Academy Meeting Yélamos O et al (2018c) Clinical and dermoscopic features of cutaneous BAP1 inactivated melanocytic tumors: results of a multicenter case-control study by the International Dermoscopy Society (IDS) Presented at the World Dermoscopy Congress, Thessaloniki 2018 Zalaudek I et al (2006) Three-point checklist of dermoscopy: an open internet study. Br J Dermatol 154:431–437. https://doi.org/10.1111/j.1365-2133.200 5.06983.x

194 Zalaudek I et al (2008) Three roots of melanoma. Arch Dermatol 144:1375–1379. https://doi.org/10.1001/ archderm.144.10.1375 Zalaudek I, Docimo G, Argenziano G (2009) Using dermoscopic criteria and patient-related factors for the management of pigmented melanocytic nevi. Arch Dermatol 145:816–826. https://doi.org/10.1001/archder matol.2009.115 Zalaudek I, Kreusch J, Giacomel J, Ferrara G, Catricala C, Argenziano G (2010) How to diagnose nonpigmented skin tumors: a review of vascular structures seen with dermoscopy: part I. Melanocytic skin tumors. J Am Acad Dermatol 63:361–374; quiz 375–366. https:// doi.org/10.1016/j.jaad.2009.11.698

K. J. Lee et al. Zhao J, Lui H, Kalia S, Zeng H (2015) Real-time Raman spectroscopy for automatic in vivo skin cancer detection: an independent validation. Anal Bioanal Chem 407:8373–8379. https://doi.org/10.1007/s00216-0158914-9 Zhao J, Zeng H, Kalia S, Lui H (2016) Wavenumber selection based analysis in Raman spectroscopy improves skin cancer diagnostic specificity. Analyst 141:1034–1043. https://doi.org/10.1039/ c5an02073e Zhao J, Zeng H, Kalia S, Lui H (2017) Using Raman spectroscopy to detect and diagnose skin cancer in vivo. Dermatol Clin 35:495–504. https://doi.org/ 10.1016/j.det.2017.06.010

Biopsy of Suspected Melanoma Noah Smith, Timothy M. Johnson, John W. Kelly, Arthur J. Sober, and Christopher Bichakjian

Contents Prebiopsy Lesion Assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196 Biopsy Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Excisional Biopsy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Incisional Biopsy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Fine-Needle Aspiration and Core Biopsy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Frozen Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biopsy of the Nail Unit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biopsy of Mucosa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

196 197 199 201 201 202 202

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

Abstract

A biopsy should be performed with relative promptness on any lesion that is clinically suspicious for melanoma, as early diagnosis is the key to a favorable prognosis. The gold standard for melanoma diagnosis is histologic

N. Smith · T. M. Johnson · C. Bichakjian (*) Department of Dermatology, University of Michigan, Ann Arbor, MI, USA e-mail: [email protected]; [email protected]. edu; [email protected] J. W. Kelly Victorian Melanoma Service, The Alfred Hospital, Melbourne, VIC, Australia e-mail: [email protected]; [email protected] A. J. Sober Department of Dermatology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_10

assessment of a tissue biopsy specimen by a dermatopathologist or pathologist with experience in interpreting melanocytic lesions. The primary goal of a biopsy is to obtain a tissue sample that is sufficient to allow the pathologist to render an accurate histologic diagnosis and assess key prognostic features for microstaging, particularly tumor thickness, if the lesion is indeed melanoma. An accurate diagnosis and tumor microstaging in turn facilitates therapeutic planning with appropriate and expeditious treatment. In this chapter, we will describe important considerations when performing a biopsy of a lesion suspicious for melanoma, commonly used biopsy techniques, and their applications to maximize biopsy accuracy in the initial diagnosis of melanoma.

195

196

Prebiopsy Lesion Assessment Assessment of the degree of suspicion of a lesion begins with a history and physical examination (Negin et al. 2003). For clinical features of and risk factors for melanoma, see also chapters ▶ “Clinical Presentations of Melanoma” and ▶ “Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma.” Difference is particularly important when one is evaluating a lesion. A difference or changes in color, size, shape, or elevation and persistent itching in a lesion are the most common early features of melanoma (Schwartz et al. 2002). Additionally, pigmented skin lesions tend to share a common overall family resemblance in appearance, with a small number of nevus patterns, within each patient (Wazaefi et al. 2013). An individual lesion that does not belong to that family and is different from other lesions should be approached with a high index of suspicion (ugly duckling sign) (Scope et al. 2008). The skin examination should be conducted under good lighting and ideally include the entire skin surface. A dermatoscope or magnifying loupes may be helpful for clinically equivocal lesions. A photograph of the lesion prior to biopsy should be strongly considered to document the appearance, size, and precise location or orientation in the event that future treatment is indicated. Palpation of the regional lymph nodes should be performed if melanoma is considered a possibility. A biopsy is a minor procedure, and thus a brief pertinent history should be obtained before the biopsy, including allergies (particularly to local anesthetics), whether the patient is taking anticoagulants, the patient’s infection risk, cardiac history (arrhythmias, presence of pacemaker and/or defibrillator), seizure history, history of vasovagal reaction, and so forth. The single most important factor for primary staging, survival, and clinical management in localized melanoma is tumor (Breslow) thickness (Balch et al. 2001a, b). Thus, a superficial shave biopsy is not appropriate if melanoma is suspected, because tumor thickness cannot be accurately determined if the lesion is transected at the deep margin (Fig. 1a). The shave biopsy

N. Smith et al.

technique may be acceptable when performed to below the anticipated plane of the lesion (Fig. 1b). Shave biopsy is sometimes preferable for evaluating epidermal or superficial dermal processes when melanoma is not suspected and is routinely used to biopsy nonmelanoma cutaneous lesions.

Biopsy Techniques An initial consideration is whether an incisional biopsy of a melanoma could disrupt melanoma cells that would then metastasize. In 1985, Lederman and Sober (Lederman and Sober 1985) provided data that demonstrated no adverse effect on prognosis that could be attributed to partial biopsy of cutaneous melanoma (incisional or punch biopsy versus excisional biopsy) after correcting for thickness. Subsequent analyses have not demonstrated any independent effect of the type of biopsy on outcome (Bong et al. 2002; Martin et al. 2005; Molenkamp et al. 2007; Stell et al. 2007). Thus, the most important principle is to promptly establish the diagnosis of melanoma by whatever type of biopsy is appropriate for the specific lesion at hand, followed promptly by definitive therapy based upon the specific characteristics of that melanoma (Sober and Balch 2007). A second consideration concerns the adequacy of each type of biopsy. There is no question that a total excisional biopsy with narrow margins provides the pathologist with the ideal specimen for establishing a correct diagnosis of melanoma and accurately microstaging the tumor (Bichakjian et al. 2011; Ng et al. 2003). Importantly, partial biopsy may be inadequate for establishing a diagnosis of cutaneous melanoma under the following circumstances: • For pathologists who evaluate specimens by pattern recognition, the assessment of important features by this method, such as asymmetry, may be precluded by a partial biopsy. Partial punch biopsy of melanoma is associated with high false-negative diagnosis rates compared with shave and excisional biopsy (Ng et al. 2010). The majority result from the

Biopsy of Suspected Melanoma

197

Fig. 1 Considerations when employing shave biopsy for suspected melanoma. (a) The tumor thickness of a melanoma cannot be accurately determined when shave biopsy results in transection at the deep margin. Care must be

taken to shave below the anticipated plane of the lesion in an attempt to avoid transecting the lesion at the deep margin. (b) Shave biopsy below the plane of the lesion will allow for accurate determination of the tumor thickness

abovementioned limitations in histopathologic interpretation of a small, partial, punch sample and fewer from sampling error (Ng et al. 2010). • For difficult or borderline lesions, such as Spitz tumors or nevoid and desmoplastic melanomas, having the entire specimen to evaluate may help improve the accuracy of the diagnosis, in contrast to having only a portion of the tumor to evaluate. • In situations where the lesion is heterogeneous (e.g., melanoma arising in a nevus), the biopsy of one portion may reflect a benign lesion, whereas the biopsy of another portion may indicate the presence of a melanoma (Fig. 2b).

be common with shave biopsy of invasive melanoma in recent studies, ranging from 9% to 68% (Egnatios et al. 2011; Mills et al. 2013; Mir et al. 2013; Zager et al. 2011). Base transection compromises the ability to correctly stage these patients and make a proper decision with regard to management, including indications for a sentinel node biopsy (Karimipour et al. 2005; Stell et al. 2007).

An additional issue of importance is the frequency with which partial biopsy may not capture the deepest component of the melanoma, resulting in underestimation of tumor thickness, potentially leading to undertreatment. In two studies, punch biopsy specimens had a higher frequency of positive margins, whereas shave biopsy specimens had a significantly higher percentage of positive deep margins than punch or excisional specimens (Karimipour et al. 2005; Stell et al. 2007). It is important to note that shave biopsy samples are the most frequent type of specimen submitted to dermatopathology laboratories nationally for the evaluation of pigmented lesions. Transection of the tumor base has been shown to

Excisional Biopsy An excisional biopsy of the entire clinically apparent lesion, with a narrow 1–2 mm margin of adjacent normal-appearing skin, is the biopsy technique of choice when melanoma is suspected (Bichakjian et al. 2011; Houghton et al. 2006; Karimipour et al. 2005). An excisional biopsy optimally provides the dermatopathologist with the entire lesion for histologic evaluation. The excision is most commonly performed in a fusiform elliptical shape with the deep margin into the subcutaneous fat (Fig. 3a–c). M-plasties may be used for diagnostic excisional biopsies to reduce the length of the scar and preserve local lymphatics, which may facilitate the accuracy of potential future lymphoscintigraphy and sentinel lymph node biopsy. Injection of radiocolloid and/ or blue dye at the tips of a long ellipse, particularly

198

N. Smith et al.

Fig. 2 Considerations when employing punch biopsy for suspected melanoma. (a) An alternative method for excising small lesions. A punch biopsy tool may be used to completely excise a small melanocytic lesion. (b) Partial punch biopsy of a heterogeneous lesion (for example, melanoma arising within a nevus) may not be representative of the entire lesion and may fail to capture the

histologically concerning aspect of the lesion. (c) Partial punch biopsy of a melanoma may be representative of the tumor thickness. (d) In some cases, partial punch biopsy of a melanoma may fail to capture the thickest component of the lesion, which could result in inadequate treatment if microstaging and treatment are based upon the tumor thickness of the partial punch biopsy

a long transverse scar on an extremity or in an ambiguous drainage location such as the trunk, may impede accurate localization of the true sentinel lymph node. The long axis of the ellipse, in general, should be oriented vertically on an extremity to preserve optimal lymphatic mapping and decrease the need for a skin graft after subsequent wide local excision. For lesions on the head, neck, and trunk, orientation along relaxed skin tension lines is typically preferred to maximize accuracy of sentinel lymph node biopsy and facilitate future wide local excision with a minimal likelihood of the need for skin grafting. Some lesions are small enough to completely excise with a punch instrument, most commonly a

6 mm punch (Figs. 2a and 4). Most patients and physicians prefer suturing of punch biopsy wounds; however, healing by granulation is an option, particularly for punch biopsies that are 4 mm or less (Christenson et al. 2005). Alternatively, a deep shave biopsy or saucerization may be performed with the use of a scalpel or razor blade to below the anticipated plane of the lesion in an attempt to avoid transecting the lesion at the deep margin (Fig. 1b) (Bichakjian et al. 2011; Swanson et al. 2002). As it is difficult to accurately estimate tumor depth prior to biopsy, intentional use of deep shave biopsy or saucerization for diagnostic excision of possible melanoma should be confined to lesions

Biopsy of Suspected Melanoma

199

Fig. 3 Technique of excisional biopsy for melanoma. (a) The lesion is infiltrated with local anesthetic, which is administered around but not into the lesion itself. (b) The entire lesion is excised including a narrow rim (1–2 mm) of normal-appearing skin and subcutaneous fat. (c) The wound is closed with sutures

that appear to be in situ or with a low index of suspicion for invasive melanoma on clinical examination. No sutures are used for a deep shave biopsy or saucerization, and the wound heals by granulation. This technique is quick, easy, and requires less training and equipment. The excisional biopsy, the first part of a twostep process, provides a diagnosis and accurate microstaging of the melanoma. The second step involves definitive therapeutic excision with a wider margin excision, with or without sentinel lymph node biopsy, depending on histopathologic parameters of the lesion and patient characteristics (Houghton et al. 2006; Karimipour et al. 2005). Excisional biopsy with wide margins should be avoided in order to facilitate the greatest accuracy of a subsequent sentinel lymph node biopsy, if indicated, and to minimize the morbidity and scarring resulting from the procedure in the event that the lesion is determined to be benign.

Incisional Biopsy Incisional biopsy involves subtotal removal of a portion of a lesion. Similar to excisional biopsy, the three general incisional biopsy techniques

Fig. 4 Technique of punch biopsy. (a) After local anesthesia, most commonly a 4–6 mm punch is placed over the desired portion of the tumor and a core of tissue is cut by pressing and rotating the punch through the skin. (b) The base of the specimen is cut to the adipose tissue, often with the use of scissors. (c) Cross section illustrating the proper depth of the biopsy wound. It should extend to the underlying subcutaneous adipose tissue to avoid transecting the lesion

include the following: fusiform ellipse, punch (see Figs. 2b–d and 4), and deep saucerization. Incisional biopsy of a suspicious lesion is a reasonable alternative when there is a low suspicion for melanoma, or the lesion is too large to excise because of its anatomic location, time constraints, the experience level of the surgeon, or concerns for cosmesis. Regardless of the type of biopsy performed, when melanoma is in the differential diagnosis, the depth of the biopsy should be

200

N. Smith et al.

to below the anticipated plane of the lesion to minimize the likelihood of transecting the lesion at the deep margin (Bichakjian et al. 2011; Karimipour et al. 2005). When an incisional biopsy is performed, the darkest or most indurated or elevated regions of the lesion are usually the preferred areas to sample, with a clear understanding that these regions may not correlate with the thickest portion of the lesion (Karimipour et al. 2005) (Figs. 5, 6, and 7). If one area within a large lesion appears darkest and thickest, an ellipse encompassing this area may be most appropriate. A second type of incisional biopsy is a deep shave biopsy or saucerization in which the majority of the lesion is essentially scooped from the skin. Alternatively, a 4- to 6-mm punch biopsy sample may be taken from within the lesion. Usually a 6-mm punch trephine is used so most of the lesion can be removed to the level of the adipose, and there is less chance of fragmentation of the lesion, which may occur with a small punch biopsy trephine. If no particular region of the lesion is most suspicious for the thickest portion

Fig. 6 An example of the importance of accurate microstaging before definitive surgical treatment. A punch biopsy specimen was initially obtained from the most elevated portion of the lesion centrally, which revealed melanoma 0.70 mm in thickness, with positive peripheral margins. Complete excision with narrow margins for microstaging revealed residual melanoma, 1.40 mm in thickness. Of interest, the most “clinically suspicious” elevated central pigmented component was marked by an experienced clinician, but only revealed melanoma 0.62 mm in thickness. Definitive wide local excision and sentinel lymph node biopsy (negative) were subsequently performed

Fig. 5 Initial small shave biopsy at the periphery of a pink lesion on the back. Clinically, basal cell carcinoma was suspected, but a biopsy revealed a small cell melanoma with a nevoid pattern, 0.64 mm in thickness with extension to the peripheral margins. Complete excision with narrow margins for microstaging revealed residual melanoma arising in association with an intradermal nevus, 1.30 mm in thickness. Definitive wide local excision and sentinel lymph node biopsy (negative) were subsequently performed

Fig. 7 Initially, two punch biopsy specimens were obtained by an experienced clinician from the most suspicious portions of a neck lesion; these revealed a melanoma 0.75 mm in thickness and melanoma in situ, respectively, to the peripheral margins. Complete excision with narrow margins for microstaging revealed residual melanoma, 2.47 mm in thickness. It is interesting to note that the most suspicious component of this excision was marked by a different experienced clinician, and revealed only melanoma in situ. Definitive wide local excision and sentinel lymph node biopsy (negative) were subsequently performed

Biopsy of Suspected Melanoma

of the tumor, or if several such areas exist, a long, narrow ellipse that traverses the diameter of the lesion and samples all clinically suspect areas may be considered (Pardasani et al. 2000). In this instance, optimal tissue embedding of the specimen so that the pathologist may examine the entire sample along its longitudinal long axis is ideal. Alternatively, multiple punch biopsies may be performed to sample the most clinically suspect areas within the lesion. Communication between the surgeon and the pathologist for clinical correlation is beneficial, particularly when only a portion of a lesion concerning for melanoma is submitted. The potential for sampling error must be recognized when a melanoma biopsy is performed by incisional rather than excisional technique. While partial sampling of a melanoma may accurately represent the thickest portion of the tumor (Fig. 2c), several studies have documented that incisional biopsies obtained by experienced clinicians do not always contain the deepest portion of a melanoma because of sampling error (Karimipour et al. 2005; MacyRoberts and Ackerman 1982; Somach et al. 1996) (see Figs. 2d, 6, and 7). One large prospective study compared melanoma microstaging of initial incisional biopsies of less than 50% of a lesion with microstaging obtained on subsequent complete excision of the residual lesion. Excision of the remaining clinical lesion performed after incisional biopsy resulted in upstaging among 21% of patients. Ten percent of patients became candidates for sentinel lymph node biopsy only after excision of the residual lesion, using the criterion of 1 mm or more in thickness as an indication (Karimipour et al. 2005). Thus when an incisional biopsy confirms a superficially invasive melanoma (e.g., 1 mm in thickness found that 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0

patients with clinically negative lymph nodes had a 5-year survival of 69.8% (Fig. 1) (Dessureault et al. 2001). Patients whose nodes were pathologically negative after elective lymph node dissection had significantly better survival (77.7%), but about a quarter of these “node-negative” patients died of melanoma. Enhanced staging by more intensive nodal evaluation following SLN biopsy identified a node-negative group who achieved 5-year survival of 90.5%. The greatly improved discrimination of favorable outcome groups afforded by SLN staging reflects the pathologists’ ability to examine the SLN more thoroughly than is possible when dealing with the multiple nodes retrieved by dissecting the entire basin. Greater accuracy is attributable to step sectioning and extensive immunohistochemistry, performed on one or at most a few SLNs, which would be impractical if required for the multiple nodes of a full dissection specimen. While modern imaging has steadily improved, even the best current imaging techniques fail to detect a significant number of nodal metastases that are readily detected by SLN biopsy. Nodal ultrasound is currently the most sensitive imaging modality for evaluating regional lymph nodes. Ultrasound characteristics associated with nodal metastases include a length to width ratio of 90% for ultrasound in detecting SLN metastases, but these results have not been able to be duplicated elsewhere, even with dedicated ultrasonographers and substantial experience (Testori et al. 2005; Thompson et al. 2011). For the melanoma centers that used ultrasound as part of the screening phase of the second Multicenter Selective Lymphadenectomy Trial (MSLT-II), ultrasound detected only 8% of sentinel node metastases. Sensitivities are better in certain groups, including patients with thicker primary tumors (Chai et al. 2012). The high operator-dependency of ultrasound may be responsible for the low sensitivity of the technique across multiple centers. The most significant challenge for ultrasound is the very small size of metastases currently identified in SLNs. In MSLT-II, the median size of SLN metastasis was 4.0 mm >4.0 mm >4.0 mm

Not applicable Unknown or unspecified Without ulceration With ulceration With or without ulceration Unknown or unspecified Without ulceration With ulceration Unknown or unspecified Without ulceration With ulceration Unknown or unspecified Without ulceration With ulceration

a

Used with permission of the American Joint Committee on Cancer (AJCC), Chicago, Illinois. The original and primary source for this information is the AJCC Cancer Staging Manual, Eighth Edition (2017), published by Springer International Publishing (Gershenwald et al. 2017b)

6.

7.

8.

9.

satellites, or in-transit metastases now categorized as N1c, N2c, or N3c based on number of tumor-involved regional lymph nodes, if any (Table 2). The site of distant metastases remains the primary component of the M category: non-visceral (distant cutaneous, subcutaneous, nodal), M1a; lung, M1b; visceral disease not involving the central nervous system (CNS) visceral, M1c; and a new M1d designation for metastases involving the CNS (Table 3). LDH remains a strong adverse predictor of survival, although elevated LDH no longer automatically categorizes a patient as M1c. Rather, descriptors have been added to each M1 subcategory to designate serum lactate dehydrogenase (LDH) level. Pathological (but not clinical) stage IA has been revised to include T1bN0M0 (formerly pathological stage IB) (Table 4). There are now four (increased from three) prognostic stage III subgroups based on N and T category criteria (Table 5).

Prognostic Factors and Staging of Primary Melanoma (AJCC Stages I and II) In the eighth edition, AJCC melanoma staging system, primary melanomas without evidence of nodal or distant metastases are characterized as stage I or II and are stratified by primary tumor characteristics (i.e., T category criteria) (Tables 1 and 4, Figs. 1 and 2). The overall prognosis for patients with localized melanoma without any nodal or distant metastatic disease is generally favorable. Based on the 2017 AJCC Melanoma Expert Panel analysis of a large, contemporary, international melanoma database, the overall 5year melanoma-specific survival (MSS) rates for patients with stages I and II disease are 98% and 90% (Fig. 2), respectively (Gershenwald et al. 2017a). Multiple known clinical and pathological factors were included in the AJCC analysis; the significance of these factors as well as others is detailed below. Emphasis was placed on factors included in the multiple covariate analyses and that contribute to the heterogeneous outcomes of

Melanoma Prognosis and Staging

275

Table 2 Eighth edition AJCC N category (regional metastasis) criteriaa Extent of regional lymph node and/or lymphatic metastasis N category NX

N0 N1 N1a N1b N1c N2 N2a N2b N2c N3

N3a N3b N3c

Number of tumor-involved regional lymph nodes Regional nodes not assessed (e.g., SLN biopsy not performed, regional nodes previously removed for another reason) Exception: Pathological N category is not required for T1 melanomas, use cN. No regional metastases detected 1 tumor-involved node or in-transit, satellite, and/or microsatellite metastases with no tumor-involved nodes 1 clinically occult (i.e., detected by SLN biopsy) 1 clinically detected No regional lymph node disease 2 or 3 tumor-involved nodes or in-transit, satellite, and/or microsatellite metastases with 1 tumor-involved node 2 or 3 clinically occult (i.e., detected by SLN biopsy) 2 or 3, at least 1 of which was clinically detected 1 clinically occult or clinically detected 4 tumor-involved nodes or in-transit, satellite, and/or microsatellite metastases with 2 tumor-involved nodes or any number of matted nodes without or with in-transit, satellite, and/or microsatellite metastases 4 clinically occult (i.e., detected by SLN biopsy) 4, at least 1 of which was clinically detected or the presence of any number of matted nodes 2 clinically occult or clinically detected and/or presence of any number of matted nodes

Presence of in-transit, satellite, and/or microsatellite metastases No

No

No No Yes

No No Yes

No No Yes

a

Used with permission of the American Joint Committee on Cancer (AJCC), Chicago, Illinois. The original and primary source for this information is the AJCC Cancer Staging Manual, Eighth Edition (2017), published by Springer International Publishing (Gershenwald et al. 2017b)

patients with stages I and II melanoma, whose 5year MSS range from 82% to 99% (Fig. 2). In the eighth edition AJCC analyses, MSS for all T categories were higher in the eighth edition than those reported in the seventh edition (for instance, 10-year MSS in the seventh edition was 93% and 39% for T1a N0 and T4b N0 melanomas, respectively, Fig. 1) (Balch et al. 2009a) and reflects, at least in part, the requirement that patients with T2-T4 primary melanoma had to have SLN biopsy to be included in the survival analyses that informed the eighth edition. As such, patients who in the past would have been classified as clinically node negative (cN0) were now more accurately staged at initial presentation, as they were only included in stage I or II if their SLNs were uninvolved by tumor, whereas in the seventh edition not all patients with T2-T4 melanoma included in the survival analyses underwent

SLN biopsy (Balch et al. 2009a; Gershenwald et al. 2017a, b). In the most recent AJCC analyses, primary tumor thickness and ulceration were identified as factors predictive of melanoma-specific survival (MSS) with 5- and 10-year MSS of 99% and 98% in patients with T1a N0 melanomas (tumor thickness 4.0 mm, ulcerated) (Fig. 1) (Gershenwald et al. 2017a, b).

Primary (Breslow) Tumor Thickness The AJCC classification of primary lesions is based primarily on microscopic assessment of the measured tumor thickness. This microstaging methodology, first defined by Breslow (1970),

276

E. Z. Keung et al.

Table 3 Eighth edition AJCC M category (distant metastasis) criteriaa M category M0 M1 M1a M1a(0) M1a(1) M1b M1b(0) M1b(1) M1c M1c(0) M1c(1) M1d M1d(0) M1d(1)

M criteria Anatomic site No evidence of distant metastasis Evidence of distant metastasis Distant metastasis to skin, soft tissue including muscles, and/or nonregional lymph node Distant metastasis to lung with or without M1a sites of disease

Distant metastasis to non-CNS visceral sites with or without M1a or M1b sites of disease Distant metastasis to CNS with or without M1a, M1b, or M1c sites of disease

LDH level Not applicable Not recorded or unspecified Not elevated Elevated Not recorded or unspecified Not elevated Elevated Not recorded or unspecified Not elevated Elevated Not recorded or unspecified Not elevated Elevated

a

Used with permission of the American Joint Committee on Cancer (AJCC), Chicago, Illinois. The original and primary source for this information is the AJCC Cancer Staging Manual, Eighth Edition (2017), published by Springer International Publishing (Gershenwald et al. 2017b)

determines the thickness of the lesion in millimeters. Using an ocular micrometer to measure its total vertical height, Breslow first established the correlation between tumor thickness and survival in 1970, when he used this measurement as a surrogate for tumor volume. This was initially confirmed in multivariate analyses by Balch et al. (1978) and Eldh et al. (1978). In a database of 1786 patients treated at the University of Alabama and at the University of Sydney, tumor thickness was found to be the most statistically significant prognostic variable using both single-factor and multifactorial analysis (Balch et al. 1982). Numerous subsequent studies, including the AJCC collaborative melanoma database analyses, have confirmed that tumor thickness is an accurate, quantitative, reproducible, and important primary prognostic factor for survival in stages I and II melanoma (Balch et al. 1982, 2001a, b, 2009a; Häffner et al. 1992; Thörn et al. 1994; Haddad et al. 1999; Eriksson et al. 2015; Gershenwald et al. 2017a, b; Lyth et al. 2017). Survival estimates for stages I and II melanoma patients decline as tumor thickness increases (Table 1 and Fig. 1) with a possible clinically important subgroup “breakpoint” in the region of 0.7–0.8 mm (Breslow 1970; Gimotty et al. 2007;

Green et al. 2012; Lo et al. 2018). In the most recent AJCC analyses of the T1 cohort (Gershenwald et al. 2017a), the impact on outcome of a 0.8 mm tumor thickness threshold, as well as mitotic rate (as a dichotomous variable, 1 mm in diameter, because of the relatively worse prognosis of this patient subgroup. Based on the currently available evidence, the AJCC Melanoma Expert Panel

recommends that, as a minimum, the single largest maximum dimension (measured in millimeters to the nearest 0.1 mm using an ocular micrometer) of the largest discrete metastatic melanoma

Melanoma Prognosis and Staging

deposit in sentinel nodes be recorded in pathology reports (Gershenwald et al. 2017b).

Extranodal Extension The presence of extranodal extension of tumor (also termed extranodal spread or extracapsular extension) is defined as the presence of a nodal metastasis extending through the lymph node capsule into adjacent tissues. Extranodal extension of tumor usually occurs in association with large clinically detected nodal metastases that demonstrate gross effacement of normal nodal architecture but may occasionally be seen with smaller lymph node metastases. Although not included as an N category criterion in the eighth edition, it is recommended that extranodal extension of tumor be recorded for future analyses (Gershenwald et al. 2017b; Crookes et al. 2017).

Non-Nodal Locoregional Metastases (Microsatellite, Satellite, and In-Transit Metastases) The presence or absence of microsatellite, satellite, or in-transit metastases, regardless of the number of such lesions, is a component of the N category in the eighth edition (Table 2 and Fig. 5c) (Gershenwald et al. 2017a, b). These are thought to represent metastases that have occurred as a consequence of intralymphatic or possibly angiotropic tumor spread (Van Es et al. 2008; Wilmott et al. 2012) and they portend a relatively poor prognosis (Buzaid et al. 1997; Rao et al. 2002; Read et al. 2015). In the eighth edition, microsatellites are defined as any foci of metastatic tumor cells in the skin or subcutis adjacent or deep to but discontinuous from the primary tumor (Gershenwald et al. 2017b). They are not clinically apparent and are diagnosed histologically. Satellite metastases are defined as any foci of clinically evident cutaneous and/or subcutaneous metastases occurring within 2 cm of but discontinuous from the primary melanoma (Gershenwald et al. 2017b). In-transit metastases are classically and empirically defined as clinically evident cutaneous and/or subcutaneous

289

metastases occurring >2 cm from the primary melanoma in the region between the primary melanoma and the regional lymph node basin (Gershenwald et al. 2017b). In the eighth edition AJCC international melanoma database, there was no significant difference in survival outcome between patients with these entities in univariate analysis and thus they are grouped together for staging purposes (Gershenwald et al. 2017a). If microsatellites, satellites, or in-transit metastases are present, the patient is categorized as N1c, N2c, or N3c based on the number of involved nodes (Table 2). Patients who undergo systemic treatment after needle biopsy of a clinically detected node or SLN biopsy are clinically staged as cN1 or greater at the time of diagnosis. Patients with clinically detected regional lymph node metastases are subcategorized as N1b, N2b, N2c, N3b, or N3c based on the number of nodes involved and whether there are also microsatellites, satellites, or in-transit disease at diagnosis.

Metastatic Melanoma to Lymph Node(s) from an Unknown Primary Site Although the majority of patients with metastatic melanoma initially present with a known primary, there is no identifiable primary tumor and no history of a primary melanoma in 2–9% of patients who present with evidence of melanoma nodal metastasis (Cormier et al. 2006; Lee et al. 2008). The axilla is the most frequent site of involved nodes when a primary site is not identified (47%), followed by the neck (29%) and groin (24%) (Gershenwald et al. 2009). Cormier et al. (Cormier et al. 2006) conducted a retrospective analysis of consecutive patients with melanoma and an unknown primary (MUP) (n = 71) who underwent surgical resection for metastasis to regional lymph nodes and were classified by N category after lymph node dissection was performed. In multivariate analyses, MUP was identified as a favorable prognostic factor for overall survival, supporting previous reports that demonstrated similar survival rates in patients with an unknown primary site compared to patients matched according to nodal metastasis

290

having a known primary site. The favorable survival profile in patients with MUP was also noted in studies of 262 patients with MUP in a study by the group at the John Wayne Cancer Institute (Lee et al. 2008), and 287 patients with MUP treated at Melanoma Institute Australia (van Der Ploeg et al. 2014a). The survival profile of patients with MUP suggests that patients with MUP have a natural history that is relatively similar to if not better than that of many patients with stage III disease from a primary site (Cormier et al. 2006; Lee et al. 2008; Rutkowski et al. 2010; Prens et al. 2011; Weide et al. 2013; de Waal et al. 2013; van Der Ploeg et al. 2014a; Bae et al. 2015; Gershenwald et al. 2017a). The inability to identify the primary tumor in patients with metastatic melanoma to lymph nodes after a thorough search does not appear to be an adverse prognostic factor. Thus, if an appropriate staging work-up does not reveal any other sites of metastases, these patients are considered AJCC stage III and should be considered for a surgical approach as well as for stage III neoadjuvant and adjuvant therapy protocols (Wargo et al. 2015; Gershenwald et al. 2017b). Of note, patients who present with skin or subcutaneous metastases without a known primary should also be presumed to be regional (stage III instead of stage IV) if appropriate staging work-up does not reveal sites of distant metastasis and pathology review by an experienced melanoma pathologist confirms that the lesion is not a variant of a primary melanoma (Gershenwald et al. 2017b).

Prognostic Factors and Staging of Patients with Distant Metastatic Melanoma (Stage IV) Patients with stage IV melanoma generally have a poorer prognosis than stage I-III melanoma. Historically, the median survival from the time of initial diagnosis of distant metastasis was 6–7.5 months and the 5-year survival was less than 10% (Barth et al. 1995; Manola et al. 2000; Unger et al. 2001; Gershenwald et al. 2009; Balch et al. 2009a). However, since the introduction of the seventh edition of the AJCC staging system, the treatment landscape

E. Z. Keung et al.

for patients with stage IV melanoma has improved dramatically, with a subset of patients achieving durable antitumor responses and dramatic survival gains (Tio et al. 2018; Vosoughi et al. 2018). Thus, the AJCC Melanoma Expert Panel concluded that for the eighth edition it was premature to embark on a broad-based analytic initiative based on new data from patients treated in recent years (Gershenwald et al. 2017b). Rather, the legacy seventh edition AJCC stage IV international melanoma database and analyses were used for the eighth edition as the primary data source, supplemented by published contemporary clinical trial data. In the eighth edition, no M stage subgroups were proposed, although revisions to the M category have been implemented as described below (Tables 3 and 4).

Site of Distant Metastasis In the eighth edition, patients with distant metastasis to the skin, subcutaneous tissue, muscle, or distant lymph nodes are categorized as M1a (Table 3) and have a relatively better prognosis than patients with distant metastases located in any other anatomic sites (Gershenwald et al. 2017a, b; Abdel-Rahman 2018). Those with lung metastasis have an intermediate prognosis and are categorized as M1b. Patients with metastasis to any other visceral sites (excluding the central nervous system [CNS]) have a relatively worse prognosis and are categorized as M1c. Melanoma patients with CNS disease generally have the worst prognosis of any of the M categories (Staudt et al. 2010; Davies et al. 2011; AbdelRahman 2018). Given the poor prognosis associated with the development of CNS metastases in melanoma patients, this subgroup of patients have been excluded from some clinical trials while in other studies the presence of CNS disease has been used as a criterion for protocol inclusion and/or stratification (Hodi et al. 2010; Robert et al. 2011, 2015a, b; Falchook et al. 2012; Flaherty et al. 2012; Long et al. 2012; Margolin et al. 2012; Larkin et al. 2014; Goldberg et al. 2016). Thus, in the eighth edition, a separate M category (M1d) was added to categorize patients with CNS metastases.

Melanoma Prognosis and Staging

LDH Level Previously, in the seventh edition of the AJCC staging system, patients with an elevated LDH level were classified as having M1c disease. In the eighth edition of the AJCC staging system, however, LDH no longer defines M1c disease. Instead, each M subcategory is modified based on whether LDH is elevated (designated as a suffix “(0)” for “not elevated” and “(1)” for “elevated” level). Patients with elevated serum LDH have been observed to have worse survival (Balch et al. 2009a; Kelderman et al. 2014; Petrelli et al. 2015; Menzies et al. 2015; Sullivan 2016; Gershenwald et al. 2017b). LDH remains a clinically significant factor associated with response (Diem et al. 2016), progression-free survival, MSS, and overall survival in the contemporary treatment era of targeted and immune therapies (Kelderman et al. 2014; Long et al. 2016a, b; Frauchiger et al. 2016).

Other Factors Other predictors of survival in patients with stage IV disease include number of metastatic sites, number of metastases, distant metastatic disease burden, performance status, and mutational status (Barth et al. 1995; Eton et al. 1998; Manola et al. 2000; Unger et al. 2001) (see also chapter on ▶ “Molecular Pathology and Genomics of Melanoma”). Elucidation and evaluation of other potential predictors of outcomes in patients with metastatic melanoma is an area of active investigation.

291

melanoma with the highest T category is assigned a suffix “(m)” or a suffix “(number)” to designate the number of primary melanomas (Gershenwald et al. 2017b). In the uncommon scenario where patients harbor regional nodal metastases and have multiple primary melanomas draining to the same regional nodal basin, the primary tumor with the highest T category is assigned as the originating primary tumor with respect to the nodal metastases. If distant metastases are present, the primary tumor with the highest N category (or highest T category if N0) is assigned as the origin of the distant metastases.

Staging Patients After Systemic or Radiation Therapy Although surgery has historically been the mainstay of treatment for patients diagnosed with melanoma, application of neoadjuvant therapy in patients with clinically detected stage III and resectable stage IV melanoma is increasingly being explored (see chapter ▶ “Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients”). The eighth edition introduces the concept of post therapy or post neoadjuvant therapy classification – yTNM – that includes T, N, and M categorization after systemic or radiation therapy intended as definitive therapy (ycTNM) or after neoadjuvant therapy followed by planned surgery (ypTNM) (Gershenwald et al. 2017b; Gress et al. 2017). The “y” classification will be useful in characterizing patients who have “up front” systemic treatment after diagnosis and allow comparison with clinical stages assigned to patients before the start of neoadjuvant or definitive therapy.

Additional Staging Recommendations Patients with Multiple Primary Melanomas In the eighth edition AJCC staging system (Gershenwald et al. 2017b), when a patient presents with multiple primary cutaneous melanomas, each is considered a different primary site and is categorized separately. When there are multiple synchronous primary melanomas, the

Staging Patients at Recurrence The eighth edition also includes a classification schema for patients who recur – rTNM – that is further divided into “r-clinical” (rcTNM) and “r-pathological” (rpTNM) stages. Although this additional classification system is in early evolution at this time, it has potential to facilitate better characterization of an individual patient’s extent

292

of disease along their melanoma continuum (Gershenwald et al. 2017a, b; Gress et al. 2017).

Conclusions A thorough understanding of prognostic factors and staging of cutaneous melanoma is crucial for patient assessment, treatment planning and sequencing, in the development of surveillance strategies, and for clinical trial design and analysis. For the eighth edition AJCC melanoma staging system for cutaneous melanoma, a new international database was created and clinically relevant revisions were made to create a standardized and contemporary staging system that facilitates patient risk stratification and guide treatment. Undoubtedly, melanoma staging will continue to evolve to reflect continued advances in our understanding of melanoma biology, prognostic factors, and treatments. Acknowledgments The authors gratefully acknowledge the members of the eighth Edition AJCC Melanoma Expert Panel and collaborating institutions that contributed clinicopathological data to the International Melanoma Database and Discovery Platform (IMDDP), developed at The University of Texas MD Anderson Cancer Center, that was used to inform revisions to the AJCC melanoma staging system.

References Abdel-Rahman O (2018) Clinical correlates and prognostic value of different metastatic sites in patients with malignant melanoma of the skin: a SEER database analysis. J Dermatolog Treat 29:176–181 Andtbacka RHI, Gershenwald JE (2009) Role of sentinel lymph node biopsy in patients with thin melanoma. J Natl Compr Cancer Netw 7:308–317 Averbook BJ, Lee SJ, Delman KA et al (2013) Pediatric melanoma: analysis of an international registry. Cancer 119:4012–4019 Azimi F, Scolyer RA, Rumcheva P et al (2012) Tumorinfiltrating lymphocyte grade is an independent predictor of sentinel lymph node status and survival in patients with cutaneous melanoma. J Clin Oncol 30:2678–2683 Azzola MF, Shaw HM, Thompson JF et al (2003) Tumor mitotic rate is a more powerful prognostic indicator than ulceration in patients with primary cutaneous melanoma: an analysis of 3661 patients from a single center. Cancer 97:1488–1498

E. Z. Keung et al. Bae JM, Choi YY, Kim DS et al (2015) Metastatic melanomas of unknown primary show better prognosis than those of known primary: a systematic review and metaanalysis of observational studies. J Am Acad Dermatol 72:59–70 Balch CM (2002) Melanoma of the skin. In: Greene F, Page D, Fleming I et al (eds) AJCC cancer staging manual, 6th edn. Springer, New York Balch CM, Murad TM, Soong SJ et al (1978) A multifactorial analysis of melanoma: prognostic histopathological features comparing Clark’s and Breslow’s staging methods. Ann Surg 188:732–742 Balch CM, Soong SJ, Milton GW et al (1982) A comparison of prognostic factors and surgical results in 1,786 patients with localized (stage I) melanoma treated in Alabama, USA, and New South Wales, Australia. Ann Surg 196:677–684 Balch CM, Soong S, Ross MI et al (2000) Long-term results of a multi-institutional randomized trial comparing prognostic factors and surgical results for intermediate thickness melanomas (1.0 to 4.0 mm). Intergroup melanoma surgical trial. Ann Surg Oncol 7:87–97 Balch CM, Buzaid AC, Soong SJ et al (2001a) Final version of the American Joint Committee on Cancer staging system for cutaneous melanoma. J Clin Oncol 19:3635–3648 Balch CM, Soong SJ, Gershenwald JE et al (2001b) Prognostic factors analysis of 17,600 melanoma patients: validation of the American Joint Committee on Cancer melanoma staging system. J Clin Oncol 19:3622–3634 Balch CM, Gershenwald JE, Soong SJ et al (2009a) Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol 27:6199–6206 Balch CM, Soong SJ, Gershenwald JE et al (2009b) Melanoma of the skin. In: Edge S, Byrd D, Compton C et al (eds) AJCC Cancer staging manual, 7th edn. Springer, New York, pp 325–344 Balch CM, Gershenwald JE, Soong SJ et al (2010) Multivariate analysis of prognostic factors among 2,313 patients with stage III melanoma: comparison of nodal micrometastases versus macrometastases. J Clin Oncol 28:2452–2459 Balch CM, Soong S, Gershenwald JE et al (2013) Age as a prognostic factor in patients with localized melanoma and regional metastases. Ann Surg Oncol 20: 3961–3968 Balch CM, Thompson JF, Gershenwald JE et al (2014) Age as a predictor of sentinel node metastasis among patients with localized melanoma: an inverse correlation of melanoma mortality and incidence of sentinel node metastasis among young and old patients. Ann Surg Oncol 21:1075–1081 Barth A, Wanek LA, Morton DL (1995) Prognostic factors in 1,521 melanoma patients with distant metastases. J Am Coll Surg 181:193–201 Bhutiani N, Egger ME, Stromberg AJ et al (2018) A model for predicting low probability of nonsentinel lymph node positivity in melanoma patients with a single positive sentinel lymph node. J Surg Oncol 118:922–927

Melanoma Prognosis and Staging Blessing K, McLaren KM, McLean A, Davidson P (1990) Thin malignant melanomas (less than 1.5 mm) with metastasis: a histological study and survival analysis. Histopathology 17:389–395 Breslow A (1970) Thickness, cross-sectional areas and depth of invasion in the prognosis of cutaneous melanoma. Ann Surg 172:902–908 Breslow A (1977) Problems in the measurement of tumor thickness and level of invasion in cutaneous melanoma. Hum Pathol 8:1–2 Busam KJ (2004) The prognostic importance of tumor mitotic rate for patients with primary cutaneous melanoma. Ann Surg Oncol 11:360–361 Buzaid AC, Ross MI, Balch CM et al (1997) Critical analysis of the current American Joint Committee on Cancer staging system for cutaneous melanoma and proposal of a new staging system. J Clin Oncol 15:1039–1051 Byrom L, Olsen CM, Knight L et al (2015) Does pregnancy after a diagnosis of melanoma affect prognosis? Systematic review and meta-analysis. Dermatol Surg 41:875–882 Caini S, Boniol M, Tosti G et al (2014) Vitamin D and melanoma and non-melanoma skin cancer risk and prognosis: a comprehensive review and meta-analysis. Eur J Cancer 50:2649–2658 Cascinelli N, Belli F, Santinami M et al (2000) Sentinel lymph node biopsy in cutaneous melanoma: the WHO melanoma program experience. Ann Surg Oncol 7:469–474 Cavanaugh-Hussey MW, Mu EW, Kang S et al (2015) Older age is associated with a higher incidence of melanoma death but a lower incidence of sentinel lymph node metastasis in the SEER databases (20032011). Ann Surg Oncol 22:2120–2126 Clark WH, From L, Bernardino EA, Mihm MC (1969) The histogenesis and biologic behavior of primary human malignant melanomas of the skin. Cancer Res 29:705–727 Clark WH, Elder DE, Guerry D et al (1989) Model predicting survival in stage I melanoma based on tumor progression. J Natl Cancer Inst 81:1893–1904 Cochran AJ, Wen D-R, Huang R-R et al (2004) Prediction of metastatic melanoma in nonsentinel nodes and clinical outcome based on the primary melanoma and the sentinel node. Mod Pathol 17:747–755 Cormier JN, Xing Y, Feng L et al (2006) Metastatic melanoma to lymph nodes in patients with unknown primary sites. Cancer 106:2012–2020 Crocetti E, Fancelli L, Manneschi G et al (2016) Melanoma survival: sex does matter, but we do not know how. Eur J Cancer Prev 25:404–409 Crookes TR, Scolyer RA, Lo S et al (2017) Extranodal spread is associated with recurrence and poor survival in stage III cutaneous melanoma patients. Ann Surg Oncol 24:1378–1385 Davies MA, Liu P, McIntyre S et al (2011) Prognostic factors for survival in melanoma patients with brain metastases. Cancer 117:1687–1696 de Waal AC, Aben KKH, van Rossum MM, Kiemeney LALM (2013) Melanoma of unknown primary origin:

293 a population-based study in the Netherlands. Eur J Cancer 49:676–683 Dewar DJ, Newell B, Green MA et al (2004) The microanatomic location of metastatic melanoma in sentinel lymph nodes predicts nonsentinel lymph node involvement. J Clin Oncol 22:3345–3349 Diem S, Kasenda B, Spain L et al (2016) Serum lactate dehydrogenase as an early marker for outcome in patients treated with anti-PD-1 therapy in metastatic melanoma. Br J Cancer 114:256–261 Duprat JP, Brechtbülh ER, Costa de Sá B et al (2016) Absence of tumor-infiltrating lymphocyte is a reproducible predictive factor for sentinel lymph node metastasis: a multicenter database study by the Brazilian melanoma group. PLoS One 11:e0148160 Egger ME, Bower MR, Czyszczon IA et al (2014) Comparison of sentinel lymph node micrometastatic tumor burden measurements in melanoma. J Am Coll Surg 218:519–528 Eggermont AMM, Chiarion-Sileni V, Grob J-J et al (2016) Prolonged survival in stage III melanoma with Ipilimumab adjuvant therapy. N Engl J Med 375:1845–1855 Eldh J, Boeryd B, Peterson LE (1978) Prognostic factors in cutaneous malignant melanoma in stage I. a clinical, morphological and multivariate analysis. Scand J Plast Reconstr Surg 12:243–255 Eriksson H, Frohm-Nilsson M, Järås J et al (2015) Prognostic factors in localized invasive primary cutaneous malignant melanoma: results of a large populationbased study. Br J Dermatol 172:175–186 Eton O, Legha SS, Moon TE et al (1998) Prognostic factors for survival of patients treated systemically for disseminated melanoma. J Clin Oncol 16:1103–1111 Falchook GS, Long GV, Kurzrock R et al (2012) Dabrafenib in patients with melanoma, untreated brain metastases, and other solid tumours: a phase 1 dose-escalation trial. Lancet 379:1893–1901 Fang S, Sui D, Wang Y et al (2016) Association of vitamin D levels with outcome in patients with melanoma after adjustment for C-reactive protein. J Clin Oncol 34:1741–1747 Faries MB, Thompson JF, Cochran AJ et al (2017) Completion dissection or observation for sentinel-node metastasis in melanoma. N Engl J Med 376:2211–2222 Fink AM, Weihsengruber F, Duschek N et al (2011) Value of micromorphometric criteria of sentinel lymph node metastases in predicting further nonsentinel lymph node metastases in patients with melanoma. Melanoma Res 21:139–143 Flaherty KT, Infante JR, Daud A et al (2012) Combined BRAF and MEK inhibition in melanoma with BRAF V600 mutations. N Engl J Med 367:1694–1703 Francischetto T, Spector N, Neto Rezende JF et al (2010) Influence of sentinel lymph node tumor burden on survival in melanoma. Ann Surg Oncol 17:1152–1158 Frankel TL, Griffith KA, Lowe L et al (2008) Do micromorphometric features of metastatic deposits within sentinel nodes predict nonsentinel lymph node involvement in melanoma? Ann Surg Oncol 15:2403–2411

294 Frauchiger AL, Mangana J, Rechsteiner M et al (2016) Prognostic relevance of lactate dehydrogenase and serum S100 levels in stage IV melanoma with known BRAF mutation status. Br J Dermatol 174:823–830 Garbe C, Eigentler TK, Bauer J et al (2016) Mitotic rate in primary melanoma: interobserver and intraobserver reliability, analyzed using H&E sections and immunohistochemistry. J Dtsch Dermatol Ges 14:910–915 Gershenwald JE, Colome MI, Lee JE et al (1998) Patterns of recurrence following a negative sentinel lymph node biopsy in 243 patients with stage I or II melanoma. J Clin Oncol 16:2253–2260 Gershenwald JE, Thompson W, Mansfield PF et al (1999) Multi-institutional melanoma lymphatic mapping experience: the prognostic value of sentinel lymph node status in 612 stage I or II melanoma patients. J Clin Oncol 17:976–983 Gershenwald JE, Andtbacka RHI, Prieto VG et al (2008) Microscopic tumor burden in sentinel lymph nodes predicts synchronous nonsentinel lymph node involvement in patients with melanoma. J Clin Oncol 26:4296–4303 Gershenwald J, Balch CM, Soong S, Al E (2009) Prognostic factors and natural history of melanoma. In: Balch CM, Soong S, Sober AJ et al (eds) Cutaneous Melanoma, 5th edn. Quality Medical Publishing, St Louis Gershenwald JE, Scolyer RA, Hess KR et al (2017a) Melanoma staging: evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67:472–492 Gershenwald JE, Scolyer RA, Hess KR et al (2017b) Melanoma of the skin. In: Amin M, Edge SB, Greene FL et al (eds) AJCC cancer staging manual, 8th edn. Springer, New York, pp 563–585 Gimotty PA, Elder DE, Fraker DL et al (2007) Identification of high-risk patients among those diagnosed with thin cutaneous melanomas. J Clin Oncol 25:1129–1134 Goldberg SB, Gettinger SN, Mahajan A et al (2016) Pembrolizumab for patients with melanoma or nonsmall-cell lung cancer and untreated brain metastases: early analysis of a non-randomised, open-label, phase 2 trial. Lancet Oncol 17:976–983 Green AC, Baade P, Coory M et al (2012) Populationbased 20-year survival among people diagnosed with thin melanomas in Queensland, Australia. J Clin Oncol 30:1462–1467 Gress DM, Edge SB, Greene FL et al (2017) Principles of Cancer staging. In: Amin M, Edge SB, Greene FL et al (eds) AJCC cancer staging manual, 8th edn. Springer, New York, pp 3–30 Gualano MR, Osella-Abate S, Scaioli G et al (2018) Prognostic role of histological regression in primary cutaneous melanoma: a systematic review and metaanalysis. Br J Dermatol 178:357–362 Guitart J, Lowe L, Piepkorn M et al (2002) Histological characteristics of metastasizing thin melanomas: a case-control study of 43 cases. Arch Dermatol 138:603–608

E. Z. Keung et al. Haddad FF, Stall A, Messina J et al (1999) The progression of melanoma nodal metastasis is dependent on tumor thickness of the primary lesion. Ann Surg Oncol 6:144–149 Häffner AC, Garbe C, Burg G et al (1992) The prognosis of primary and metastasising melanoma. An evaluation of the TNM classification in 2,495 patients. Br J Cancer 66:856–861 Han D, Zager JS, Shyr Y et al (2013) Clinicopathologic predictors of sentinel lymph node metastasis in thin melanoma. J Clin Oncol 31:4387–4393 Helsing P, Robsahm TE, Vos L et al (2016) Cutaneous head and neck melanoma (CHNM): a population-based study of the prognostic impact of tumor location. J Am Acad Dermatol 75:975–982.e2 Hodi FS, O’Day SJ, McDermott DF et al (2010) Improved survival with ipilimumab in patients with metastatic melanoma. N Engl J Med 363:711–723 Huismans AM, Haydu LE, Shannon KF et al (2014) Primary melanoma location on the scalp is an important risk factor for brain metastasis: a study of 1,687 patients with cutaneous head and neck melanomas. Ann Surg Oncol 21:3985–3991 In ‘t Hout FE, Haydu LE, Murali R et al (2012) Prognostic importance of the extent of ulceration in patients with clinically localized cutaneous melanoma. Ann Surg 255:1165–1170 Jones MS, Lee J, Stern SL, Faries MB (2017) Is pregnancyassociated melanoma associated with adverse outcomes? J Am Coll Surg 225:149–158 Kashani-Sabet M, Sagebiel RW, Ferreira CM et al (2001) Vascular involvement in the prognosis of primary cutaneous melanoma. Arch Dermatol 137:1169–1173 Kelderman S, Heemskerk B, Van Tinteren H et al (2014) Lactate dehydrogenase as a selection criterion for ipilimumab treatment in metastatic melanoma. Cancer Immunol Immunother 63:449–458 Kelly JW, Sagebiel RW, Blois MS (1985) Regression in malignant melanoma. A histologic feature without independent prognostic significance. Cancer 56:2287–2291 Kyrgidis A, Lallas A, Moscarella E et al (2017) Does pregnancy influence melanoma prognosis? A metaanalysis. Melanoma Res 27:289–299 Lachiewicz AM, Berwick M, Wiggins CL, Thomas NE (2008) Survival differences between patients with scalp or neck melanoma and those with melanoma of other sites in the surveillance, epidemiology, and end results (SEER) program. Arch Dermatol 144:515–521 Larkin J, Ascierto PA, Dréno B et al (2014) Combined vemurafenib and cobimetinib in BRAF-mutated melanoma. N Engl J Med 371:1867–1876 Lee CC, Faries MB, Wanek LA, Morton DL (2008) Improved survival after lymphadenectomy for nodal metastasis from an unknown primary melanoma. J Clin Oncol 26:535–541 Leiter U, Stadler R, Mauch C et al (2016) Complete lymph node dissection versus no dissection in patients with sentinel lymph node biopsy positive melanoma

Melanoma Prognosis and Staging (DeCOG-SLT): a multicentre, randomised, phase 3 trial. Lancet Oncol 17:757–767 Leiter UM, Stadler R, Mauch C et al (2018) Final analysis of DECOG-SLT trial: survival outcomes of complete lymph node dissection in melanoma patients with positive sentinel node. J Clin Oncol 36: suppl;abstr 9501 Lim A, Shayan R, Varigos G (2018) High serum vitamin D level correlates with better prognostic indicators in primary melanoma: a pilot study. Australas J Dermatol 59(3):182–187. Epub 2017 Mar 23 Lo SN, Scolyer RA, Thompson JF (2018) Long-term survival of patients with thin (T1) cutaneous melanomas: a Breslow thickness cut point of 0.8 mm separates higher-risk and lower-risk tumors. Ann Surg Oncol 25:894–902 Long GV, Trefzer U, Davies MA et al (2012) Dabrafenib in patients with Val600Glu or Val600Lys BRAF-mutant melanoma metastatic to the brain (BREAK-MB): a multicentre, open-label, phase 2 trial. Lancet Oncol 13:1087–1095 Long GV, Grob J-J, Nathan P et al (2016a) Factors predictive of response, disease progression, and overall survival after dabrafenib and trametinib combination treatment: a pooled analysis of individual patient data from randomised trials. Lancet Oncol 17:1743–1754 Long GV, Weber JS, Infante JR et al (2016b) Overall survival and durable responses in patients with BRAF V600-mutant metastatic melanoma receiving Dabrafenib combined with Trametinib. J Clin Oncol 34:871–878 Long GV, Hauschild A, Santinami M et al (2017) Adjuvant Dabrafenib plus Trametinib in stage III BRAF-mutated melanoma. N Engl J Med 377:1813–1823 Lyth J, Falk M, Maroti M et al (2017) Prognostic risk factors of first recurrence in patients with primary stages I-II cutaneous malignant melanoma – from the population-based Swedish melanoma register. J Eur Acad Dermatol Venereol 31:1468–1474 Mandalà M, Galli F, Cattaneo L et al (2017) Mitotic rate correlates with sentinel lymph node status and outcome in cutaneous melanoma greater than 1 millimeter in thickness: a multi-institutional study of 1524 cases. J Am Acad Dermatol 76:264–273 Manola J, Atkins M, Ibrahim J, Kirkwood J (2000) Prognostic factors in metastatic melanoma: a pooled analysis of eastern cooperative oncology group trials. J Clin Oncol 18:3782–3793 Margolin K, Ernstoff MS, Hamid O et al (2012) Ipilimumab in patients with melanoma and brain metastases: an open-label, phase 2 trial. Lancet Oncol 13:459–465 Menzies AM, Wilmott JS, Drummond M et al (2015) Clinicopathologic features associated with efficacy and long-term survival in metastatic melanoma patients treated with BRAF or combined BRAF and MEK inhibitors. Cancer 121:3826–3835 Moreno-Ramírez D, Ojeda-Vila T, Ríos-Martín JJ et al (2015) Role of age and sex in the diagnosis of early-

295 stage malignant melanoma: a cross-sectional study. Acta Derm Venereol 95:940–942 Morton DL, Thompson JF, Cochran AJ et al (2006) Sentinel-node biopsy or nodal observation in melanoma. N Engl J Med 355:1307–1317 Morton DL, Thompson JF, Cochran AJ et al (2014) Final trial report of sentinel-node biopsy versus nodal observation in melanoma. N Engl J Med 370:599–609 Murali R, Hughes MT, Fitzgerald P et al (2009) Interobserver variation in the histopathologic reporting of key prognostic parameters, particularly Clark level, affects pathologic staging of primary cutaneous melanoma. Ann Surg 249:641–647 Murali R, Desilva C, Thompson JF, Scolyer RA (2010) Non-sentinel node risk score (N-SNORE): a scoring system for accurately stratifying risk of non-sentinel node positivity in patients with cutaneous melanoma with positive sentinel lymph nodes. J Clin Oncol 28:4441–4449 Nagarajan P, Curry JL, Ning J et al (2017) Tumor thickness and mitotic rate robustly predict melanoma-specific survival in patients with primary vulvar melanoma: a retrospective review of 100 cases. Clin Cancer Res 23:2093–2104 Nagore E, Oliver V, Botella-Estrada R et al (2005) Prognostic factors in localized invasive cutaneous melanoma: high value of mitotic rate, vascular invasion and microscopic satellitosis. Melanoma Res 15:169–177 National Comprehensive Cancer Network. Version 1.2019. Date Accessed 16 Feb 2019. Melanoma Paladugu RR, Yonemoto RH (1983) Biologic behavior of thin malignant melanomas with regressive changes. Arch Surg 118:41–44 Pawlik TM, Ross MI, Thompson JF et al (2005) The risk of in-transit melanoma metastasis depends on tumor biology and not the surgical approach to regional lymph nodes. J Clin Oncol 23:4588–4590 Petrelli F, Cabiddu M, Coinu A et al (2015) Prognostic role of lactate dehydrogenase in solid tumors: a systematic review and meta-analysis of 76 studies. Acta Oncol 54:961–970 Prade M, Sancho-Garnier H, Cesarini JP, Cochran A (1980) Difficulties encountered in the application of Clark classification and the Breslow thickness measurement in cutaneous malignant melanoma. Int J Cancer 26:159–163 Prens SP, van der Ploeg APT, van Akkooi ACJ et al (2011) Outcome after therapeutic lymph node dissection in patients with unknown primary melanoma site. Ann Surg Oncol 18:3586–3592 Quaglino P, Marenco F, Osella-Abate S et al (2010) Vitiligo is an independent favourable prognostic factor in stage III and IV metastatic melanoma patients: results from a single-institution hospital-based observational cohort study. Ann Oncol 21:409–414 Ranieri JM, Wagner JD, Azuaje R et al (2002) Prognostic importance of lymph node tumor burden in melanoma

296 patients staged by sentinel node biopsy. Ann Surg Oncol 9:975–981 Rao UNM, Ibrahim J, Flaherty LE et al (2002) Implications of microscopic satellites of the primary and extracapsular lymph node spread in patients with high-risk melanoma: pathologic corollary of eastern cooperative oncology group trial E1690. J Clin Oncol 20:2053–2057 Read RL, Haydu L, Saw RPM et al (2015) In-transit melanoma metastases: incidence, prognosis, and the role of lymphadenectomy. Ann Surg Oncol 22:475–481 Rees MJ, Liao H, Spillane J et al (2016) Localized melanoma in older patients, the impact of increasing age and comorbid medical conditions. Eur J Surg Oncol 42:1359–1366 Ribero S, Osella-Abate S, Sanlorenzo M et al (2013) Favourable prognostic role of regression of primary melanoma in AJCC stage I-II patients. Br J Dermatol 169:1240–1245 Ribero S, Gualano MR, Osella-Abate S et al (2015) Association of histologic regression in primary melanoma with sentinel lymph node status: a systematic review and meta-analysis. JAMA Dermatol 151:1301–1307 Ribero S, Moscarella E, Ferrara G et al (2016) Regression in cutaneous melanoma: a comprehensive review from diagnosis to prognosis. J Eur Acad Dermatol Venereol 30:2030–2037 Ribero S, Stucci LS, Marra E et al (2018) Effect of age on melanoma risk, prognosis and treatment response. Acta Derm Venereol 98:624–629 Robert C, Thomas L, Bondarenko I et al (2011) Ipilimumab plus Dacarbazine for previously untreated metastatic melanoma. N Engl J Med 364:2517–2526 Robert C, Long GV, Brady B et al (2015a) Nivolumab in previously untreated melanoma without BRAF mutation. N Engl J Med 372:320–330 Robert C, Schachter J, Long GV et al (2015b) Pembrolizumab versus Ipilimumab in Advanced Melanoma. N Engl J Med 372:2521–2532 Ronan SG, Eng AM, Briele HA et al (1987) Thin malignant melanomas with regression and metastases. Arch Dermatol 123:1326–1330 Rousseau DL, Ross MI, Johnson MM et al (2003) Revised American Joint Committee on Cancer staging criteria accurately predict sentinel lymph node positivity in clinically node-negative melanoma patients. Ann Surg Oncol 10:569–574 Rutkowski P, Nowecki ZI, Dziewirski W et al (2010) Melanoma without a detectable primary site with metastases to lymph nodes. Dermatol Surg 36:868–876 Schatton T, Scolyer RA, Thompson JF, Mihm MC (2014) Tumor-infiltrating lymphocytes and their significance in melanoma prognosis. Methods Mol Biol 1102:287–324 Schmoeckel C, Bockelbrink A, Bockelbrink H, BraunFalco O (1983) Low- and high-risk malignant melanoma–II. Multivariate analyses for a prognostic classification. Eur J Cancer Clin Oncol 19:237–243 Schultz S, Kane M, Roush R et al (1990) Time to recurrence varies inversely with thickness in clinical stage

E. Z. Keung et al. 1 cutaneous melanoma. Surg Gynecol Obstet 171:393–397 Scolyer RA, Li L-XL, McCarthy SW et al (2004) Micromorphometric features of positive sentinel lymph nodes predict involvement of nonsentinel nodes in patients with melanoma. Am J Clin Pathol 122:532–539 Sinnamon AJ, Sharon CE, Song Y et al (2018) The prognostic significance of tumor-infiltrating lymphocytes for primary melanoma varies by sex. J Am Acad Dermatol 79:245–251 Spatz A, Cook MG, Elder DE et al (2003) Interobserver reproducibility of ulceration assessment in primary cutaneous melanomas. Eur J Cancer 39:1861–1865 Speijers MJ, Bastiaannet E, Sloot S et al (2015) Tumor mitotic rate added to the equation: melanoma prognostic factors changed? : a single-institution database study on the prognostic value of tumor mitotic rate for sentinel lymph node status and survival of cutaneous melanoma patients. Ann Surg Oncol 22:2978–2987 Starz H, Balda BR, Krämer KU et al (2001) A micromorphometry-based concept for routine classification of sentinel lymph node metastases and its clinical relevance for patients with melanoma. Cancer 91:2110–2121 Staudt M, Lasithiotakis K, Leiter U et al (2010) Determinants of survival in patients with brain metastases from cutaneous melanoma. Br J Cancer 102:1213–1218 Storr SJ, Safuan S, Mitra A et al (2012) Objective assessment of blood and lymphatic vessel invasion and association with macrophage infiltration in cutaneous melanoma. Mod Pathol 25:493–504 Sullivan RJ (2016) Serum lactate dehydrogenase is a more useful biomarker of prognosis than serum S100B in patients with BRAF-mutant melanoma. Br J Dermatol 174:716–717 Tas F, Erturk K (2017) Scalp melanoma is associated with high mitotic rate and is a poor prognostic factor for recurrence and outcome. Melanoma Res 27:387–390 Taylor RC, Patel A, Panageas KS et al (2007) Tumorinfiltrating lymphocytes predict sentinel lymph node positivity in patients with cutaneous melanoma. J Clin Oncol 25:869–875 Thomas NE, Busam KJ, From L et al (2013) Tumor-infiltrating lymphocyte grade in primary melanomas is independently associated with melanoma-specific survival in the population-based genes, environment and melanoma study. J Clin Oncol 31:4252–4259 Thompson JF, Soong SJ, Balch CM et al (2011) Prognostic significance of mitotic rate in localized primary cutaneous melanoma: an analysis of patients in the multiinstitutional American Joint Committee on Cancer melanoma staging database. J Clin Oncol 29:2199–2205 Thörn M, Pontén F, Bergström R et al (1994) Clinical and histopathologic predictors of survival in patients with malignant melanoma: a population-based study in Sweden. J Natl Cancer Inst 86:761–769 Timerman D, McEnery-Stonelake M, Joyce CJ et al (2017) Vitamin D deficiency is associated with a worse prognosis in metastatic melanoma. Oncotarget 8:6873–6882

Melanoma Prognosis and Staging Tio M, Wang X, Carlino MS et al (2018) Survival and prognostic factors for patients with melanoma brain metastases in the era of modern systemic therapy. Pigment Cell Melanoma Res 31(4):509–515. Epub 2018 Jan 10 Unger JM, Flaherty LE, Liu PY et al (2001) Gender and other survival predictors in patients with metastatic melanoma on Southwest Oncology Group trials. Cancer 91:1148–1155 van Akkooi ACJ, Nowecki ZI, Voit C et al (2008) Sentinel node tumor burden according to the Rotterdam criteria is the most important prognostic factor for survival in melanoma patients: a multicenter study in 388 patients with positive sentinel nodes. Ann Surg 248:949–955 van der Ploeg APT, van Akkooi ACJ, Rutkowski P et al (2011) Prognosis in patients with sentinel node-positive melanoma is accurately defined by the combined Rotterdam tumor load and Dewar topography criteria. J Clin Oncol 29:2206–2214 Van Der Ploeg APT, Van Akkooi ACJ, Haydu LE et al (2014a) The prognostic significance of sentinel node tumour burden in melanoma patients: an international, multicenter study of 1539 sentinel node-positive melanoma patients. Eur J Cancer 50:111–120 van Der Ploeg APT, Haydu LE, Spillane AJ et al (2014b) Melanoma patients with an unknown primary tumor site have a better outcome than those with a known primary following therapeutic lymph node dissection for macroscopic (clinically palpable) nodal disease. Ann Surg Oncol 21:3108–3116 Van Es SL, Colman M, Thompson JF et al (2008) Angiotropism is an independent predictor of local recurrence and in-transit metastasis in primary cutaneous melanoma. Am J Surg Pathol 32:1396–1403 Varey AHR, Goumas C, Hong AM et al (2017) Neurotropic melanoma: an analysis of the clinicopathological

297 features, management strategies and survival outcomes for 671 patients treated at a tertiary referral center. Mod Pathol 30:1538–1550 Vosoughi E, Lee JM, Miller JR et al (2018) Survival and clinical outcomes of patients with melanoma brain metastasis in the era of checkpoint inhibitors and targeted therapies. BMC Cancer 18:490 Wargo JA, Amaria RN, Ross MI et al (2015) Neoadjuvant BRAF (dabrafenib) and MEK (trametinib) inhibition for high-risk resectable stage III and IV melanoma. J Clin Oncol 33(Suppl):TPS9091 Wat H, Senthilselvan A, Salopek TG (2016) A retrospective, multicenter analysis of the predictive value of mitotic rate for sentinel lymph node (SLN) positivity in thin melanomas. J Am Acad Dermatol 74:94–101 Weide B, Faller C, Elsässer M et al (2013) Melanoma patients with unknown primary site or nodal recurrence after initial diagnosis have a favourable survival compared to those with synchronous lymph node metastasis and primary tumour. PLoS One 8:e66953 Wilmott J, Haydu L, Bagot M et al (2012) Angiotropism is an independent predictor of microscopic satellites in primary cutaneous melanoma. Histopathology 61:889–898 Wong SL, Faries MB, Kennedy EB et al (2018) Sentinel lymph node biopsy and management of regional lymph nodes in melanoma: American Society of Clinical Oncology and Society of Surgical Oncology clinical practice guideline update. J Clin Oncol 36:399–413 Wyatt C, Lucas RM, Hurst C, Kimlin MG (2015) Vitamin D deficiency at melanoma diagnosis is associated with higher Breslow thickness. PLoS One 10:e0126394 Xie C, Pan Y, McLean C et al (2017) Impact of scalp location on survival in head and neck melanoma: a retrospective cohort study. J Am Acad Dermatol 76:494–498.e2

Models for Predicting Melanoma Outcome Lauren E. Haydu, Phyllis A. Gimotty, Daniel G. Coit, John F. Thompson, and Jeffrey E. Gershenwald

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Prediction Tools and Statistical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Personalized Prognosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Clinical Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Link Between Prediction Tools and Staging Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Relevance of Prediction Tools for Clinical Trials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Brief History of Melanoma Prediction Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Prediction Tools Developed from AJCC Databases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Other Prediction Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

300 300 301 301 301 301 302 302 302

L. E. Haydu (*) Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] P. A. Gimotty Department of Biostatistics, Epidemiology and Informatics, Perelman School of Medicine, The University of Pennsylvania, Philadelphia, PA, USA D. G. Coit Department of Surgery, Memorial Sloan Kettering Cancer Center, New York, NY, USA J. F. Thompson Melanoma Institute Australia, Faculty of Medicine and Health, The University of Sydney, Sydney, NSW, Australia Department of Melanoma and Surgical Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia J. E. Gershenwald Departments of Surgical Oncology and Cancer Biology, Melanoma and Skin Center, The University of Texas MD Anderson Cancer Center, Houston, TX, USA © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_5

299

300

L. E. Haydu et al. Planning to Build a Prediction Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reporting Prediction Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Criteria for Building Prediction Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Selection of a Patient Population . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Selection of an Outcome to Predict . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conditional Survival . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Probability of Binary Outcome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Considering the Treatment Landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Selection of Relevant and Clearly Defined Predictors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

302 302 303 303 305 306 306 307 307

Model Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cox Proportional Hazards Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hazard Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Relative Risk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Proportional Hazards Assumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . When the Proportional Hazards Assumption Is Violated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

308 308 309 309 309 309

Model Validation and Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Internal Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . External Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Putting Contemporary Models to the Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

309 310 310 311

Challenges and Opportunities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sample Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Selection Bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Missing Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Survival from Metastatic Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

311 311 311 312 312

The Future . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312 Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313

Abstract

Clinical and pathological features that impact melanoma patient survival have been studied extensively for decades at major melanoma centers around the world. With the aid of powerful statistical techniques and computational methods, remarkable progress has been made in the identification of dominant factors that are linked to the natural history of melanoma and associated clinical outcome. A wide array of clinical prediction tools have been promulgated, primarily focused on forecasting survival outcomes across the melanoma continuum, with the exception of distant metastatic (Stage IV) melanoma. Recent changes in melanoma clinical practice resulting from the availability of new targeted and immune therapies that are effective in both metastatic and adjuvant settings, as well as level I evidence demonstrating no

survival benefit for completion lymph node dissection after a positive sentinel lymph node biopsy, have together changed the melanoma landscape and will no doubt impact on approaches to outcome prediction. Against this contemporary and ever-evolving backdrop, we present clinical applications, criteria, challenges, and opportunities for interpreting and building tools for predicting melanoma outcomes.

Introduction Prediction Tools and Statistical Models In the field of clinical research, broadly, a prediction tool is the implementation of a statistical model into a readily usable format such as a nomogram, classification tree, or electronic calculator that forecasts patient

Models for Predicting Melanoma Outcome

outcomes. The underlying statistical model for a given prediction tool is typically built using multiple regression methods that calculate the probability of patient survival (or hazard of death). Statistical models, in general, are built utilizing a training data set of independent variables, also known as predictors, together with known patient outcomes (representing the dependent variable, e.g., survival or sentinel lymph node positivity). A prognostic tool is a specific type of prediction tool that focuses on survival-based patient outcomes. Once a statistical model is developed, it is necessary to validate the model, which is the process by which the performance, or accuracy (of the predictions), of the model is assessed using an independent data set of predictors and patient outcomes.

301

Clinical Applications Prediction of the likely clinical course of a disease and the expected treatment outcome is an essential part of medical practice. Clinicians face daily decisions that include selection and recommendation of an “optimal” treatment for an individual patient. Follow-up evaluation strategies can vary according to a patient’s prognosis. A prediction tool may generate a prognostic summary analysis of survival and disease recurrence rates for an individual patient based on presenting characteristics. Clinicians may use these projections in conjunction with other factors, including likelihood of treatment response and the morbidity or potential toxicity associated with treatment, to guide treatment decisionmaking and the frequency and duration of follow-up.

Personalized Prognosis The concept of personalized medicine, defined by the National Cancer Institute (NCI) at the National Institutes of Health (NIH) on their website (2018), involves the use of specific information about a patient’s “genes, proteins, and environment to prevent, diagnose, plan treatment, find out how well treatment is working, or make a prognosis.” It is not possible to exactly predict the outcome of an individual patient without a crystal ball; however, the extent to which we approach outcome prediction for an individual patient is the concept of personalized prognosis (i.e., individualized prognosis). The spectrum of outcome prediction ranges from a model that estimates a broad range of survival for a heterogeneous group of patients with melanoma, for example, to a model that estimates a more specific range of outcomes for a homogeneous group of patients (e.g., 50-year-old female patient presenting with 1.5 mm Breslow thickness melanoma and a negative sentinel lymph node biopsy). The latter end of the spectrum described is considered a more “personalized” or “individualized” prognosis compared with the former.

The Link Between Prediction Tools and Staging Systems The importance of staging and classification of melanoma patients is described in chapter ▶ “Melanoma Prognosis and Staging.” The staging system is necessarily structured in order for it to serve as a common global language of cancer classification; however, due at least in part to the de facto constrained (i.e., TNM-based) approach, it cannot incorporate some potentially important disease-specific prognostic factors, may not aggregate groups of patients solely according to risk, nor assign a more individualized prognosis. A prediction tool that incorporates factors such as patient age and sex, as well as other factors beyond those used in the TNM classification, may better facilitate personalized clinical decision-making at important treatment and surveillance decision points.

The Relevance of Prediction Tools for Clinical Trials Tools that predict patient outcomes based on contemporary prognostic estimates derived from

302

historical, standard-of-care cohorts are useful during the planning of clinical trials to facilitate informed power calculations and to select patient eligibility criteria and stratification variables. For clinical trials not employing pre-randomization stratification or those stratifying patients according to inappropriate factors, a prediction tool can be used to determine whether any differences in outcome are the result of treatment or are caused by an imbalance in prognostic factors. In analyzing clinical trials, treatments are often compared within various subgroups. These subgroups are mostly defined using combinations of known relevant prognostic factors. A well-known limitation of subgroup analysis is the difficulty in extending analysis beyond two or three variables unless a very large sample is available. In the analysis of survival data, this is further complicated by the possibility of varied censoring patterns and durations of patient follow-up within the different subgroups. No adequate statistical inferences can be drawn from “overstretched” subgroup analyses. As in pre-randomization stratification, a prediction tool can be used to classify patients into an adequate number of subgroups for analysis after completion of the study.

Brief History of Melanoma Prediction Tools Dr. Seng-jaw Soong (1985) developed the first prognostic tool in the form of a scoring system for predicting survival outcome for patients with localized melanoma. The validity of the underlying statistical model was tested in independent data sets, and its high degree of predictability made it clinically useful. Next, Soong et al. (1992) developed generalized multivariable prognostic models for patients with localized melanoma to address survival at diagnosis, survival after a disease-free interval, and probability of recurrence. The models were developed by analyzing a combined database of 4568 patients from the Sydney Melanoma Unit (SMU) and the University of Alabama at Birmingham (UAB).

L. E. Haydu et al.

Prediction Tools Developed from AJCC Databases In 2000, the American Joint Committee on Cancer (AJCC) assembled a melanoma database that contained details of 17,600 patients, allowing the development of a new, evidence-based sixth edition melanoma staging system, as well as new prognostic models at diagnosis for both localized and regional melanoma (Balch et al. 2001a, b; Soong et al. 2003). Multivariable analyses were later performed on a substantially enhanced AJCC melanoma database developed for revision of the AJCC staging system and for development of statistical models. The results of these analyses, used to update the prognostic tools, including model validation with an independent data set and parametric modeling, and including hazard function estimation for localized melanoma, were published in the previous (5th) edition of this book (Soong et al. 2009).

Other Prediction Tools In addition to the pioneering prognostic modeling efforts of Soong and colleagues, as well as those initiated by the AJCC, multiple tools for predicting melanoma survival have been subsequently developed over the past two decades and published in the form of regression models, including coefficients or formulas, prognostic classification trees, nomograms or scoring systems, or as internetbased (online) electronic calculators (Table 1). These tools focused on endpoints of melanomaspecific survival (MSS), disease-specific survival (DSS), or overall survival (OS), and predicted outcomes for patients presenting with localized and/ or locoregionally metastatic melanoma.

Planning to Build a Prediction Model Reporting Prediction Models Standardized recommended approaches have been developed for Transparent Reporting of multivariable prediction models for Individual

Models for Predicting Melanoma Outcome

303

Table 1 Melanoma-specific (MSS) or overall survival (OS) prediction tools published in the prior two decades. (Adapted from Mahar et. al. 2016) Sample size 1042

Dates of cohort 1980–1990

Survival outcome MSS

Thin melanoma (i.e. Breslow thickness  1 mm)

26,921

1988–2001

MSS

Gimotty

Thin melanoma (i.e. Breslow thickness  1 mm)

2389

1972–2001

MSS

2010

Soong

Stage I–II

14,760

MSS

2011

Michaelson

Stage I–III

MSS

Online tool

2012

Callender

OS

Online tool

2013

Lyth

11,165

1990–2008

MSS

2014

Khosrotehrani

494

2000–2011

MSS

Prognostic subgroups Nomogram

2014

Maurichi

Clinically negative lymph nodes; Breslow thickness  1 mm Thin melanoma (i.e. Breslow thickness  1 mm) AJCC seventh edition stage IIIB and IIIC Thin melanoma (i.e. Breslow thickness  1 mm)

Not reported 2507

Not reported Not reported 1997–2003

Type of tool Individualized risk formula Prognostic classification tree Prognostic classification tree Online tool

2243

1996–2004

OS

Nomogram

Year 2000

First author Cochran

Population Stage I–III

2007

Gimotty

2007

Prognosis or Diagnosis (TRIPOD), broadly across disease types (Collins et al. 2015). In their groundbreaking work, Collins et al. engaged a multidisciplinary panel of methodologists, as well as healthcare and editorial professionals, to produce the TRIPOD statement, a checklist of 22 items (Table 2) required to be reported in order for prognostic or diagnostic tools to be transparently and properly published (Collins et al. 2015).

Criteria for Building Prediction Models In 2016, as the work product of a workshop conducted by the Precision Medicine Core of the American Joint Committee on Cancer, Kattan et al. published criteria for endorsement of cancer prognostic models by the AJCC. More specifically addressing statistical methodologies for cancer prediction models, Kattan et al. (2016) described 13 inclusion and 3 exclusion criteria (Table 3) for potential endorsement by the AJCC of an individualized prognostic calculator for use in cancer clinical decision-making. As an extension of these published criteria, this section provides additional

guidance for the development and interpretation of a robust model for predicting melanoma patient outcomes.

Selection of a Patient Population A useful prediction model provides estimates that are relevant to a specific patient population and selected outcome measure (e.g., melanoma-specific survival or overall survival) relevant to the clinical decision-making process. Multiple models, reflecting distinct patient populations and/or outcome measures, may be incorporated into electronic tools or calculators that provide options for the user to select the specific prediction model relevant to their patient. According to the TRIPOD statement, for each prediction model, it is necessary to report the source of data, participants, and predicted outcome measure that was the focus of the development and validation of the model. Such information includes whether the cohort was from a clinical trial, registry, or other source; specific study dates (i.e., start and end dates of patient accrual); the study

304

L. E. Haydu et al.

Table 2 Summary of the TRIPOD checklist items. (Adapted from Collins et al. 2015) Item number 1 2

Section Title Abstract

3a 3b 4a 4b 5a 5b 5c 6a 6b 7a 7b 8 9 10a 10b

Introduction Introduction Methods – source of data Methods – source of data Methods – participants Methods – participants Methods – participants Methods – outcome Methods – outcome Methods – predictors Methods – predictors Methods – sample size Methods – missing data Statistical methods Statistical methods

10c 10d 10e 11 12

Statistical methods Statistical methods Statistical methods Methods – risk groups Methods – development vs. validation Results – participants Results – participants Results – participants Results – model development Results – model development Results – model specification Results – model specification Results – model performance Results – model updating Discussion – limitations Discussion – interpretation Discussion – interpretation Discussion – implications Supplementary information Funding

13a 13b 13c 14a 14b 15a 15b 16 17 18 19a 19b 20 21 22

Checklist item to include Development or validation model; target population; outcome to be predicted Summary of objectives, study design, setting, participants, sample size, predictors, outcome, statistical analysis, results, and conclusions Medical context and rationale Objectives including whether development or validation model Study design and type of data source Key study dates; start/end accrual, follow-up Study setting – number and type of centers Eligibility criteria Details of treatment(s) received Clearly define outcome predicted Any blind outcome assessment Define all predictors and how they are measured Any blind predictor assessment How sample size determined How missing data are handled How predictors are handled in the model (development only) Type of model, model-building procedure, and internal validation method (development only) Describe how predictions were calculated (validation only) Methods to assess model performance and/or model comparisons Model updating (validation only) Details on creation of risk groups Differences from development model (validation only) Describe participants with and without outcome, and follow-up time Characteristics of participants including missing data Comparison with development data set (validation only) Number of participants and outcome events (development only) Unadjusted association between predictors and outcome (development only) Full prediction model (development only) How to use the prediction model (development only) Model performance and confidence intervals Model updating results (validation only) Limitations of the study Discuss results with reference to development model performance (validation only) Overall results interpretation Potential clinical use Availability of supplementary resources Source of funding and role of funders

Models for Predicting Melanoma Outcome Table 3 Summary of the AJCC prognostic model building criteria. (Adapted from Kattan et al. 2016) Inclusion criteria 1. Overall survival, disease-specific survival, and disease-specific mortality are the outcomes to be predicted 2. Prediction model should address a clinically relevant question 3. Include relevant predictors, or explain why a relevant predictor was not included 4. Specify which patients were used to develop the model, and how patients were selected to validate the model 5. Assess the model for generalizability or external validation 6. Define the time-zero from which prognosis is assessed 7. Predictors are to be known at time-zero and well defined 8. Provide details of the model such that it can be implemented, or free access 9. Report a measure of discrimination 10. Report calibration curves 11. Validation should occur over a contemporary and relevant time frame and clinical practice setting 12. Indicate which initial treatment(s) is applied, if applicable 13. Development and/or validation of the model must be published in a peer-reviewed journal Exclusion criteria 1. Substantial proportion of patients in the validation data set had no follow-up 2. No reporting of missing data in the validation data set 3. Small number of events in the validation data set

setting (e.g., general practice versus tertiary-referral center); and eligibility criteria and treatment details (Collins et al. 2015). The treatment center and patient referral pathway can impact the mix of treatment options for any cancer patient. Since the type of treatment a patient receives is directly related to their outcome, it is important to be aware of treatment differences in any data set used for model building or validation. In particular, newly established guidelines or contemporary evidence may be more rapidly implemented in a tertiary-referral, specialist-laden cancer center. If patient data have been captured at a tertiary-referral center, review of their preceding treatment and diagnosis in the community will be needed in order to

305

confirm that the case is in line with the clinical management and diagnosis was reported as a prerequisite for the prediction tool. Finally, for the use of international data sets, the issues outlined above can be exacerbated due to differences in clinical practice, with respect not only to treatment availability and decisionmaking but also, potentially, to differences in the standard conduct and availability of dermatology, pathology (e.g., specimen sectioning methods and standard immunohistochemistry staining practices), nuclear medicine, and imaging (e.g., availability of magnetic resonance imaging (MRI) or positron-emission tomography (PET). It is therefore important to have an in-depth understanding of the clinical practice of each institution that contributes patient data for model building. It may be possible to create and use a new variable to describe and account for any such differences deemed to significantly impact the model.

Selection of an Outcome to Predict As introduced above, whether the utility of the model is intended for decision-making to guide routine follow-up, specific treatment options, or clinical trial or public health planning, the type of prognostic estimates that the model generates needs to be designed with contemporary practice in mind. Across the melanoma continuum, the changing landscape of treatment options, including surgical, intralesional, and systemic therapies, has resulted in new priorities and points of decision-making, for which robust prognostic estimates are needed. Although AJCC endorsement of prediction models is currently limited to overall and disease-specific survival outcomes (Kattan et al. 2016), there is a need to create tools that predict time to melanoma relapse, either at any anatomic site or at a specific site (e.g., probability of melanoma brain metastasis within 3 years of initial diagnosis), as well as non-survival-based outcomes, such as the probability of tumor-involved non-sentinel lymph nodes following a positive sentinel lymph node biopsy (SLNB).

306

Time to Relapse After Initial Disease Management Time-to-relapse analysis has many pseudonyms in the scientific literature, including, but not limited to, relapse-free survival, time to recurrence, recurrence-free survival, disease-free survival, progression-free survival, time to progression, etc. In addition, various ways in which these analyses are defined may lead to challenges in the comparison and meta-analysis of results across multiple studies for any given cohort. An initiative to standardize definitions for clinical trial endpoints is known as the DATECAN project, or Definitions for the Assessment of Time-to-event Endpoints in CANcer trials (Gourgou-Bourgade et al. 2015). For the purpose of building prediction models, utilizing established standard definitions and transparent reporting are both critical for the patient and/or clinician end user to be confident with their interpretation of the estimates generated for the patient’s outcome. Published guidance for industry by the US Department of Health and Human Services (2007) provides a definition for disease-free survival as time to the first of either recurrence or death from any cause and notes that this endpoint may be useful for the assessment of adjuvant treatments (in the context of gaining approval for biologics) when the disease-free survival benefit outweighs any observed toxicity. It is important to note that the standard definition of an event in any relapse-type survival analysis, such as disease-free, recurrence-free, or relapse-free survival, includes death from any cause as an event in addition to the relapse event. Care must be taken in the interpretation of these endpoints, because a perceived inflated risk of recurrence of disease may result, particularly in low-risk populations, in which death from other causes would be significantly more likely to occur. For example, survival analysis may not be optimal for estimating time to a specific disease relapse, or site of relapse, in older patients that have a higher likelihood of death from another cause (Berry et al. 2010). Other methods, such as the cumulative incidence function, may be more appropriate to estimate the proportion of patients that will experience a specific type of relapse (or any relapse of disease), while

L. E. Haydu et al.

considering death from any cause as a competing risk (Pintilie 2006; Kim 2007).

Conditional Survival Melanoma survival models are useful for generating prognostic estimates from the time of initial diagnosis of the patient’s primary melanoma presentation. This information is essential for accurate patient staging and to guide decision-making for initial clinical management. However, estimates that are generated based on clinical information at the time of initial diagnosis do not indicate how a patient’s prognosis may change throughout follow-up. In this case, conditional survival estimates are more useful and are derived from modeling patient cohorts that have survived for specified periods of time (e.g., patients are entered into the model on the condition of surviving 1 year or any other predefined time point). For cohorts of patients that have survived for some specified time after diagnosis, the baseline predictors may no longer be significantly associated with these patient’s future survival. Conditional survival models may elucidate subsets of these survivors who are still at significant risk, for example, of death or relapse, while defining those at lower risk and providing them with some relative reassurance (Collett 2015; Hosmer et al. 2007; Haydu et al. 2017; Hieke et al. 2015).

Probability of Binary Outcome There is also interest and applicability of other, non-survival type multivariable regression models in melanoma, such as binary logistic regression (Hosmer and Lemeshow 2000). Probably the most investigated binary endpoint in melanoma is the histological status of sentinel lymph nodes (i.e., positive or negative) and, likewise, the histological status of non-sentinel lymph nodes (i.e., predicting whether a completion lymph node dissection (CLND) procedure will identify additional metastatic melanoma deposits in non-SLNs; Murali et al. 2010; Cadili et al. 2010). Since patients undergoing SLN biopsy, by definition, have no clinical evidence

Models for Predicting Melanoma Outcome

of regional lymph node metastasis, a prediction model is very useful to determine the indication for the procedure for all patient groups, but particularly in lower-risk patient groups, such as those with a thin primary melanoma (1 mm).

Considering the Treatment Landscape Clinical care milestones such as the introduction and routine uptake of SLNB in the 1990s (▶ “Melanoma Prognosis and Staging”), the FDA approval of BRAF inhibitors and immune checkpoint inhibitors beginning in 2011 (▶ “Cytokines (IL-2, IFN, GM-CSF, etc.) Melanoma” and ▶ “Managing Checkpoint Inhibitor Symptoms and Toxicity for Metastatic Melanoma”), and recent evidence indicating no overall survival benefit for CLND versus observation (Faries et al. 2017; Leiter et al. 2016) impact not only the decision-making process for melanoma patients and clinicians but also the relevance of underlying data sets that are relied upon for training and validation of prediction models. For example, the current clinical paradigm has shifted from completion lymph node dissection (CLND) to observation, as well as consideration of adjuvant treatment for SLNB-positive melanoma patients. In this new treatment paradigm, the probability of relapse for the SLNB-positive patient, and the factors that contribute to their elevated risk of relapse, will guide decisions about how they should be followed (e.g., frequency and extent of imaging) and whether and when to recommend potentially toxic and costly adjuvant therapy. Models incorporating the number of positive non-sentinel lymph nodes as a predictor of patient outcome are no longer relevant since these data are no longer collected. Furthermore, these milestones are not fixed time points at which clinical practice changes but generally indicate the commencement of a phase of change management when considering the uptake of new procedures (e.g., SLNB). Other examples include the effective implementation of new clinical trial evidence for surgery or radiotherapy (e.g. CLND or adjuvant radiotherapy), and the on protocol, compassionate use, and standard-of-care phases of use of new

307

intralesional or systemic agents either in the adjuvant or therapeutic settings. For example, although SLNB was formally introduced in 1992 (Morton et al. 1992), due to the fact that it is a complex procedure requiring a multidisciplinary team including surgeon, pathologist, and nuclear medicine clinician working in concert, there was significant evolution of technique throughout the mid- to late 1990s, depending on the treatment center (▶ “Melanoma Prognosis and Staging”). Furthermore, the indication for SLNB in specific patient groups, e.g., patients with thin primary melanomas (280–315 nm), and UVC (100–280 nm). All UVC and most UVB wavelengths are blocked by the stratospheric ozone layer. A fraction of UVB and all UVA reaches the Earth’s surface. In a meta-analysis summarizing 57 studies on sun exposure and melanoma, intermittent sun exposure (such as all activities related to tan-seeking behavior) and history of sunburns, a marker of intermittent exposure, double melanoma risk: summary relative risk (SRR) = 1.61 (95% confidence intervals (CI), 1.31–1.99) and SRR, 2.03 (95% CI, 1.73–2.37), for intermittent sun exposure and sunburns, respectively. However, skin type is a significant confounder in these associations as those with fair skin are more likely to

473

report sunburns, and it is possible that genetic factors associated with fairer skin increase melanoma risk, and these may have nothing to do with sun exposure. Genes in the melanocortin pathways have important roles not only in pigmentation but also in immune functions and energy regulation both of which will have an impact on cancer risk. Many new genes are currently been discovered in relation to hair and skin color based on several 100,000 individuals using the UK Biobank and 23andMe cohorts, and these genes will need to be investigated in melanoma cohorts. Based on existing epidemiological data and clinical observations, Green (1992) proposed a theory of site-dependent susceptibility of melanocytes to malignant transformation, which is close to the dual pathway theory for melanoma. According to this hypothesis, people with a low propensity for melanocyte proliferation need a continuous exposure to sunlight in order to drive the clonal expansion of initiated melanocytes. Melanomas developing in this pathway are more likely to be located on sun-exposed body sites, to be of lentigo maligna melanoma (LMM) subtype, and to occur in older patients with a history of solar damage and NMSC. On the contrary, individuals with a high propensity to melanocyte proliferation, as indicated by a large number of common nevi, tend to develop melanomas on intermittently sun-exposed body sites, belonging to the superficial spreading melanoma (SSM) or nodular melanoma (NM) histological subtypes and showing little if any association with a history of NMSC or sun damage (see sections “Phenotypical Factors” and “Common Melanocytic and Atypical Nevi”). In the same study by Green (1992), sun exposure measures and phenotypic characteristics were generally positively associated with all histological types of melanoma. NM, however, was not associated with sunburns, in contrast to LMM and SSM. LMM was not associated with freckling, light eye color, and hair color, in contrast to NM and SSM, which were significantly associated with all three. This two-pathway hypothesis for melanoma was confirmed and refined by many authors who

474

observed an inverse correlation between number of nevi and clinical signs of sun damage (Bataille et al. 1998; Elwood and Gallagher 1998; Whiteman et al. 1998, 2003), and few genes differentially mutated in LMM versus SSM and NM were identified (see sections “Phenotypic Risk Factors” and “Common and Atypical Nevi”). Specifically, mutations occurring on BRAF or NRAS genes, along with other genes coding for downstream components of the tyrosine kinase RAS-BRAF signal transduction pathway (like CDKN2A and CDK4), were suggested to be more frequent in melanomas occurring on intermittently exposed skin (Curtin et al. 2005; Maldonado et al. 2003; Thomas et al. 2007). MC1R, a pigmentation gene associated with melanoma risk (Pasquali et al. 2015; Raimondi et al. 2008; Tagliabue et al. 2018), is involved in the same signaling pathway and has been found to positively interact with BRAF and CDKN2A in the etiology of melanoma occurring on usually unexposed skin (Box et al. 2001; Landi et al. 2006). On the other hand, p53-positive melanomas were usually associated with features of chronical sun exposure (Whiteman et al. 1998), supporting the hypothesis that different molecular pathways can lead to melanoma development (de Snoo and Hayward 2005; Rivers 2004). The distribution by body site of different histological types of CM is uneven, with SSM the more frequent type on the trunk in men and legs in women and LMM more frequently found on the face and neck (Newell et al. 1988). It is very likely that melanocytes on different body sites have different characteristic in terms of differentiation. It is already quite clear for dermatologists that atypical nevi are more commonly found on the trunk while these are very rare on the face. Similarly, intradermal nevi, which are very mature melanocytic lesions, are commonly found on the face but are much rarer on limbs, for example. It is therefore possible that during embryogenesis, melanocytes have different properties according to the head and neck, trunk, and limbs because of migration to different body sites and this is likely to be influenced by key developmental genes. In a further meta-analysis, the results of 24 separate studies encompassing 16,180 cases of

V. Bataille et al.

melanoma did focus on the associations with sun exposure, pigmentary characteristics, and melanocytic nevi with CM at particular body sites and of particular histological types (Caini et al. 2009). A higher SRR for CM on usually sun-exposed sites (or the head when an anatomical definition of body site was used) than on the other body sites with each measure of sun exposure (including measures of actinic skin damage) was the most consistent pattern in these associations. Considering each measure of sun exposure (intermittent, chronic, sunburns, and actinic damage), SRRs were 1.31 (95% CI, 0.94–1.81) and 1.77 (95% CI, 1.30–2.41), respectively, for occasionally versus usually sun-exposed body sites. Chronic sun exposure was weakly, but significantly, negatively associated with CM on occasionally sun-exposed sites, most strongly so on the legs. Overall, these results suggest that sun exposure is associated with CM on all body sites (except mucosal) but more so for the head and neck in older individuals. The apparently protective effect of chronic sun exposure on CM on occasionally exposed sites and, at most, weakly causal effect on usually exposed sites is puzzling. While enhanced melanin production and melanosome delivery to keratinocytes (Lin et al. 2008) and increased thickness of the top layers of the epidermis due to continuing sun exposure may offer a partial explanation, they would not be expected to reduce incidence to a level below that present in the absence of sun exposure. It is important to note, however, that the reference category for calculating RRs in epidemiological studies of melanoma and sun exposure is invariably “low” sun exposure, not “no” sun exposure. Migrant studies provide convincing evidence of childhood and adolescence being critical periods for the development of a later melanoma. They suggest that adults are at increased melanoma risk if they spent their childhood in sunny geographical locations or generally received above average intermittent sun exposure during vacation and/or recreation. In an Australian case-control study published in 1984, age at arrival and duration of residence in Australia was studied. Earlier age at arrival was a

Clinical Genetics and Risk Assessment of Melanoma

predictor of melanoma risk with little residual effect of duration of residence. Specifically, migrants arriving before age 10 years appeared to have a risk similar to that of native born Australians, whereas the estimated incidence in those arriving after age 15 years was around one quarter of the native born rates, with arrival at later ages giving no additional advantage. Similarly, in an European case-control study, age 5 mm in diameter and may

display color variegation and an irregular shape. These features can make some Clark nevi difficult to distinguish from early melanoma

506

C. Navarrete-Dechent et al.

Fig. 4 Dermoscopic patterns of benign nevi. (Reproduced with permission from Dermoscopedia ®. www. dermoscopedia.org)

Fig. 5 Melanoma-specific structures that can be seen under the dermatoscope. (Reproduced with permission from Dermoscopedia ®. www.dermoscopedia.org)

According to Annessi et al., they are characterized by the presence of (1) disordered junctional nesting (variation of size, shape, and location of nests with common “bridging”), (2) lentiginous

melanocytic proliferation with elongation of rete ridges (“club” shaped), (3) cytological atypia (nuclear enlargement, nuclear pleomorphism, nuclear hyperchromasia, prominent nucleoli),

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

and (4) “shoulder phenomenon” – referring to nevi with a central compound component and a peripheral junctional component (Annessi et al. 2001). The significance of recognizing Clark nevi may be in that they can help identify individuals at an increased risk for developing melanoma (Halpern et al. 1993; Hofmann-Wellenhof et al. 2016). While most melanomas arise de novo on normal-appearing skin, up to 30% of melanomas arise in association with a Clark nevus or with a nevus with a congenital pattern; nevus-associated melanomas are more frequently encountered among patients with high nevus counts that include large-diameter nevi (Pampena et al. 2017). Thus, Clark nevi can be viewed as potential, non-obligate precursors to melanoma (see “Precursor” section below) (Hofmann-Wellenhof et al. 2016; Tsao et al. 2003). Numerous studies have reported that Clark nevi are markers for an increased risk for melanoma (Fig. 6). A meta-analysis showed that when analyzed in a dichotomous fashion, the presence of Clark nevi resulted in a tenfold increased risk

507

for developing melanoma compared to the absence of any Clark nevi (clinically determined; RR = 10.1 [95% CI: 5.0–20.3) (Gandini et al. 2005a). When adjusting for the total number of Clark nevi present, the risk increased from 1.6fold when there was 1 Clark nevus to 4.1-fold with 3 Clark nevi and to 10.5-fold with 5 Clark nevi. Individuals that harbor an increased nevus count together with the presence of multiple Clark nevi are at higher risk for developing melanoma, and these patients can be classified as having the “multiple Clark nevus phenotype” (Gandini et al. 2005a). This phenotype is associated with variable melanoma risk – the risk is highest among individuals with multiple Clark nevi and a family history of melanoma among relatives with the same phenotype and lowest among individuals with sporadic phenotype of multiple Clark nevi (Fig. 7).

Congenital Melanocytic Nevi Congenital melanocytic nevi (CMN) develop in utero and are usually visible at birth. A subset of CMN, known as tardive CMN, may not become

Fig. 6 Relative risk of developing a melanoma according to the number of Clark/dysplastic nevi in a given patient. (Data obtained from Gandini et al. 2005a)

508

C. Navarrete-Dechent et al.

Fig. 7 Patient with the socalled atypical mole syndrome. These individuals often have over 100 nevi displaying increased variability in size (with many larger than 6 mm), shape, and color. As seen in this patient, it is not uncommon for these individuals to have multiple scars from previous excisions

clinically apparent until months to years after birth; hence, tardive CMNs (syn. Congenital nevus-like nevus) may be clinically impossible to differentiate from acquired nevi such as Clark nevi. Most CMN are often larger than 5 mm, are sharply demarcated and lack a macular component, and often display a mammillated surface and hypertrichosis (Fig. 7). Most CMN grow in proportion to the growth of the child (Fig. 8) (Alikhan et al. 2012). The prevalence of CMN ranges from 0.7% to 3.9% (Ingordo et al. 2007; Kanada et al. 2012; Walton et al. 1976). CMN are classified according to the projected adult maximum diameter: small (60 cm) (Krengel et al. 2013). Large CMN >20 cm are infrequent with an incidence of 1:20,000–1:500,000 births; in contrast, the incidence for small CMN is 1% (Fig. 7) (Castilla et al. 1981; Rhodes 1986). While the mutation profile of small CMN has not been extensively studied, it appears that most harbor a BRAF mutation (Bauer et al. 2007; Ichii-Nakato et al. 2006). In contrast, up to 80% of large CMN harbor an NRAS mutation, which is a somatic mutation presumed to occur early during embryogenesis (Bauer et al. 2007). A very early NRAS mutation in utero may also affect a neural crest progenitor cell, leading to the formation of a large CMN, as well as to appearance of multiple, so-called “satellite” CMN, and proliferation of

melanocytes in the central nervous system (a constellation of findings termed “neurocutaneous melanocytosis”) (Kinsler et al. 2017a, 2013; Waelchli et al. 2015b). Indeed, Kinsler et al. has demonstrated that the same somatic NRAS mutation that was found in the melanocytes within the cutaneous CMN can also be found in the melanocytes deposited within the CNS, but not in unaffected skin or blood (Kinsler et al. 2013). While large and giant CMN are at increased risk for developing melanoma within the CMN (see “Precursor” section below), they appear not to heighten the individual’s lifetime risk for developing de novo cutaneous melanoma on normal skin. Regarding medium CMN (1.5–19.9 cm), there are but a few studies showing that they may be markers for an increased risk for developing melanoma with a RR of 4.1 for developing de novo melanoma (Kopf et al. 1985a, b). It is difficult to know whether the nevi in these studies were CMN that were present at birth, tardive CMN, or Clark nevi.

Spitz Nevi This subtype of nevus is more common in childhood. These benign neoplasms usually appear as pink dome-shaped papules but can also take on an irregular morphology making it impossible to differentiate from melanoma (Fig. 8). Spitz nevi are often associated with mutations in the HRAS gene (Bastian et al. 2000). They also present with

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

509

Fig. 8 Congenital melanocytic nevus. (a) Medium-size CMN, note the presence of evenness in color and hypertrichosis on a 9-year-old boy lower leg. (b) Large-size

CMN on the entire back of a newborn. (Reproduced with permission from ISIC archive. www.isic-archive.com)

kinase fusion/rearrangement such as ROS1 (17%), NTRK1 (16%), and ALK (10%) (Busam et al. 2014; Wiesner et al. 2014). Because of common morphological features between Spitz nevi and melanomas, Lallas et al. have advocated removal of Spitz nevi in patients aged 12 and older (Lallas et al. 2015). To date, only one small study has shown that a previous history of a Spitz nevus may confer an eightfold increased risk for developing a melanoma later in life (Sepehr et al. 2011).

any associated cancer risk, the occurrence of multiple bapomas has been associated with an autosomal dominant familial cancer syndrome associated with a germline-inactivating mutation in BAP1. These patients are predisposed to developing various malignancies, including mesothelioma, renal cell carcinoma, and cutaneous and uveal melanoma (Wiesner et al. 2011, 2012). Despite lack of formal guidelines, recommendations by experts have stated that it is not unreasonable to obtain BAP1 germline mutations testing in individuals harboring two or more bapomas or if there is positive family history of BAP1-associated tumors. If the patient is found to have the germline mutation, screening examination by pertinent specialties (e.g., gastroenterology and urology) should be considered. This would include at least once-a-year total body skin examination for cutaneous melanoma, as well as uveal melanoma screening by an ophthalmologist (Haugh et al. 2017).

BAP1-Inactivated Melanocytic Tumors (“Bapomas”) BRCA1-associated protein 1 (BAP1) is a tumor suppressor gene located on chromosome 3. BAP1-inactivated melanocytic tumors (BIMTs), also known as “bapomas” or “BAP1-deficient tumors,” are melanocytic nevi that clinically appear as dome-shaped, pink papules, resembling dermal nevi or fibromas (Fig. 9a–b) (Haugh et al. 2017; Moawad et al. 2018). These nevi display a distinct histopathological profile, with a predominantly dermal proliferation of melanocytes that includes a combination of larger epithelioid (“spitzoid”) melanocytes, as well as “conventional” smaller melanocytes (Wiesner et al. 2011). BIMTs may present as a sporadic nevus or as multiple nevi that are part of a “BAP1deficient syndrome.” While an isolated sporadic BAP1-deficient nevus probably does not confer

Classic Genotype/Phenotype Risk Correlates MC1R Polymorphisms (Variants), Skin Color, and Hair Color Melanocortin 1 receptor (MC1R) is a G proteincoupled, highly polymorphic receptor, with more

510

C. Navarrete-Dechent et al.

Fig. 9 Congenital melanocytic nevus (CMN), medium size. Column (a) shows the same CMN baseline at 1 year of age (A1), at 3 years old (A2), and at 9 years old (A3) showing a symmetric growth. Columns (b 1 – 3) and (c 1 – 3) show the corresponding close-up and dermoscopic

appearance, respectively. This nevus has grown in proportion to the growth of the child, is homogeneous brown in color, and has a reticular pattern on dermoscopy. (Reproduced with permission from ISIC archive. www. isic-archive.com)

than 80 genetic variants described. It is the main regulator of pigmentation, phototype, hair color, and sun sensitivity, controlling pigment production. Signal transduction through MC1R is responsible for the synthesis of the two main types of melanin, the brown-black eumelanin and the red-yellow pheomelanin (FukunagaKalabis and Herlyn 2012). Approximately 80% of patients with Fitzpatrick skin type 1 and red hair carry a dysfunctional MC1R polymorphism in both alleles. About 60% of persons with skin type 2 carry a single loss-of-function MC1R allele. Fewer than 20% of individuals with skin type 3 or higher carry one dysfunctional allele (Roider and Fisher 2016). The eumelanin-topheomelanin ratio explains different phototypes (Wendt et al. 2018). Research has shown that eumelanin is responsible for darker pigment and provides protection from UV damage, which helps in preventing the development of

melanoma. In contrast, pheomelanin – which is found in lighter skin – provides little to no protection from UV light and may also act as a free radical, increasing the risk for developing melanoma (Mitra et al. 2012; Morgan et al. 2013; Sturm et al. 2003; Valverde et al. 1995). A paradox in this group of patients is that administration of antioxidants may increase their risk for developing distant metastasis. These findings suggest that while oxidative stress can initiate the carcinogenesis cascade, it can also potentially inhibit the spread of a cancer to other sites, limiting distant melanoma metastasis (Harris and Brugge 2015; Piskounova et al. 2015). Numerous MC1R polymorphisms have been associated to melanomaprone patients with reported relative risk of between 1.4 and 2.5 (Potrony et al. 2015; Raimondi et al. 2008). Interestingly, MC1R red hair variants do not appear to confer risk among the Icelandic population suggesting that an epistatic

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

risk SNP (e.g., ASIP) may be responsible in this population (Gudbjartsson et al. 2008). A melanoma model in mice carrying an inactivated mutation of MC1R gene (showing a red hair phenotype) and BRAFV600E mutation had a high rate of developing invasive melanomas despite being kept in a UV radiation-free environment. When the production of pheomelanin was ablated (by introducing an albino gene), selective absence of pheomelanin was protective against melanoma. These data suggest that the pheomelanin pigment pathway present in the red hair phenotype produces ultraviolet-radiation-independent carcinogenesis and contributions to melanoma genesis via oxidative damage. Although protection from UV radiation remains important, additional strategies may be required for optimal melanoma prevention in this subgroup (Fukunaga-Kalabis and Herlyn 2012; Mitra et al. 2012). In contrast, eumelanin is a strong antioxidant and reduces the accumulation of DNA damage by absorbing reactive oxygen species (ROS), acting as a DNA shield. In a recent study, carrying a single MC1R red hair variant was significantly associated with melanoma risk only in females. Carrying two or more variants was associated with an even higher risk for melanoma (OR, 2.7; 95% CI [1.9–3.8]). These findings remained statistically significant after multivariate analysis (including age, freckling, lentigines, and sun exposure). For males, increased risk was found only if two or more red hair variants were present (OR, 1.7; 95% CI [1.1–2.4]); however, in multivariate analysis, these findings did not retain significance. The findings of this study suggest that perhaps other factors such as estrogens may mediate melanoma risk in females (Wendt et al. 2018). Melanoma risk has been associated with the presence of large nevi (Newton-Bishop et al. 2010). Interestingly, MC1R polymorphisms may also be associated with the development of larger nevi, and this may help explain the molecular underpinning of the increased melanoma risk seen in patients harboring these large nevi. It also highlights the potential role of MC1R in the phenotype of nevi and nevogenesis including possible growth-promoting effect on melanocytes (Vallone et al. 2018).

511

Fair-skinned color is a well-known risk factor for the development of melanoma. A meta-analysis showed that patients with skin type I or II have a threefold higher risk for melanoma as compared to patients with skin type V and VI (Gandini et al. 2005b). Similarly, a dichotomous risk analysis based on “light” versus “dark/medium” skin colors also showed a 1.9-fold risk increase (95% CI 1.6–2.2) (Gandini et al. 2005b). Similar to skin type, there is an association between lighter hair color and melanoma risk (Roider and Fisher 2016), whereby the strongest association is seen in individuals with red hair color, which has been linked to polymorphisms in the MC1R gene (Raimondi et al. 2008; Roider and Fisher 2016). MC1R dysfunctional variants in both alleles are present in >80% of individuals with fair skin and/or red hair color (also known as the “red hair color phenotype”) (Hawkes et al. 2016; Roider and Fisher 2016). A meta-analysis comparing red versus dark, blonde versus dark, and light brown versus dark brown hair color showed a 3.6-, 2.0-, and 1.6-fold increased risk for melanoma, respectively.

Eye Color Light eye color is independently associated with an increased risk of melanoma (Fortes et al. 2010). Similar to skin type and hair color, individuals with lighter-colored eyes (blue or green) are at a 1.6-fold higher risk for melanoma as compared to subjects with dark-colored eyes (95% CI 1.4–1.8) (Gandini et al. 2005b). A meta-analysis showed that when comparing blue versus dark, green versus dark, and hazel versus dark, there was an associated 1.5-, 1.6-, and 1.5-fold increased risk for melanoma. Freckles and Lentigines Freckles, also known as ephelides, are small-tomedium, brown macules with evenly distributed pigment arising on sun-exposed sites (McLean and Gallagher 1995). Studies have shown an increased melanoma risk in patients with freckles (Elwood et al. 1990; Fortes et al. 2010; Gandini et al. 2005b; Osterlind et al. 1988; Wendt et al. 2018; White et al. 1994; Williams et al. 2011). One study reported that intense freckling as an adult or as a

512

child were both associated with a sixfold increase in melanoma risk (Elwood et al. 1990). A metaanalysis showed that high density of freckles was associated with a melanoma RR of 2.1 (95% CI 1.8–2.4) compared to low density of freckles. Also, “high density of freckles before age 25” conferred a RR of 2.3 (95% CI, 1.7; 3.1) (Gandini et al. 2005b). Another study showed a similar risk (OR 2.8–3.0) in females and males, respectively, when there was presence of freckles in childhood (Wendt et al. 2018). Williams et al. demonstrated that for white persons ages 35–74, a higher density of freckles on the arms before the age of 20 was associated with an OR of 1.6–2.7 for melanoma (Williams et al. 2011). Solar lentigines are acquired pigmented macules or patches on sun-damaged skin that reveal both keratinocytic and melanocytic hyperplasia, but without the presence of melanocytic atypia or nests of melanocytes (Andersen et al. 1997). The presence of lentigines on different body sites was associated with an increased risk for melanoma in multivariate analysis. The presence of lentigines on the back (OR 1.9; 95% CI [1.1–3.2]) and on the face (OR 1.6; 95% CI [1.1–2.3]) was determined as significant melanoma risk factors in females. For males, the presence of lentigines on the back (PR 2.6; 95% CI [1.5–4.5]) and the hands (OR 2.3; 95% CI [1.2–4.3]) was associated with an increased risk for melanoma (Wendt et al. 2018). Another study showed that melanoma patients had a higher number of solar lentigines compared with controls (Idorn et al. 2015). There was also a positive association between melanoma and a higher number of lentigines.

Precursor Lesions The classic cancer progression model infers that cancer arises through the accumulation of genetic alterations; this ultimately leads to uncontrolled cellular proliferation, which may eventuate in metastasis (Frank 2007; Liotta and Kohn 2003; Ramaswamy et al. 2003; Shain et al. 2015; Vogelstein et al. 1988). The classic model of carcinogenesis implies the existence of intermediary

C. Navarrete-Dechent et al.

lesions possessing genetic alterations that are inbetween those seen in completely banal lesions and cancer. These intermediary lesions, while not obligate precursors to cancer, may be at a heightened risk of progressing into cancer as compared to their completely banal counterpart. Thus, if such intermediary lesions could be identified, then it may be possible to prevent their progression into cancer via chemoprevention or surgical removal. Alternatively, these intermediary lesions could be closely monitored and promptly removed should any signs of early cancer develop. In regard to melanoma, CMN (including nevus spilus), Clark nevi, and solar lentigines have been implicated as potential melanoma precursor lesion.

Clark Nevus (Syn. Large Acquired Nevi, Atypical, Dysplastic Nevi) While melanomas may develop in association with Clark nevi (Fig. 10), most melanomas developing in individuals with Clark nevi develop de novo from normal-appearing skin (Fuller et al. 2007; Kittler et al. 2000a, b; Pampena et al. 2017; Salerni et al. 2012a). Only approximately 30% of cases of melanoma will have an associated nevus, and histologically these nevi are 77.4% acquired nevi and 22.6% CMN. Interestingly, when dissecting the “acquired” subgroup, more melanomas arose in association with non-dysplastic as compared to dysplastic subtype (56.7 vs. 43.3%, respectively); however, this difference was not statically significant (Pampena et al. 2017; Rosendahl et al. 2015). The actual percentage of nevus-associated melanoma may in fact be higher than reported since the nevus may have been obliterated by the melanoma growth. The risk of transformation of any given individual nevus into melanoma has been estimated to be 1:500,000 (0.0002%) for both men and women younger than 40 years of age and 1 in 33,000 (0.003%) for men older than 60 years. For a 20-year-old individual, the estimated lifetime risk of an individual nevus transforming into melanoma by age 80 years is

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

513

Fig. 10 Dermoscopy of a Reed nevus showing a starburst pattern (inset shows the clinical features). (Image courtesy of Pablo Uribe, MD, PhD (Department of Dermatology, Facultad de Medicina, Pontificia Universidad Catolica de Chile, Santiago, Chile))

Fig. 11 Melanoma developing in a congenital melanocytic nevus. A 55year-old male with Breslow’s 0.63 mm melanoma (white arrows) arising in association with a small CMN (white asterisk). The inset shows clinical features. (Reproduced with permission from ISIC archive (www.isic-archive. com). Courtesy of Dr. Harold Rabinovitz)

calculated to be approximately 0.03% (1/3165) for men and 0.009% (1/10800) for women (Tsao et al. 2003). This data, together with the knowledge that 70% of melanomas occur de novo, suggests that while melanoma can develop in association with nevi, the risk for any individual nevus transforming into melanoma is extremely low, making prophylactic excision of all nevi a worthless endeavor (Brod et al. 2009; Zalaudek et al. 2010). (Kittler and Tschandl 2013). While Clark nevi are non-obligate, very low-risk melanoma precursors, the presence of multiple Clark nevi

is a risk marker for developing melanoma during the patient’s lifetime.

Congenital Melanocytic Nevi Melanoma developing in association with CMN is well documented (Fig. 11). The overall absolute risk of melanoma developing in a CMN, irrespective of its size, is between 0.1% and 5%, which is lower than the lifetime overall risk of developing a melanoma in the US general white population (3.7% for males and 2.4% for females)

514

but higher than the risk of an acquired nevus transforming into melanoma (as was discussed previously) (Bett 2005; Hale et al. 2005; Kinsler et al. 2009, 2017a; Siegel et al. 2018; Zaal et al. 2005). The melanoma risk for small and medium CMN is lower than the risk for large CMN (Alikhan et al. 2012; Kinsler et al. 2009; Krengel et al. 2006). The melanomas associated with small CMN tend to develop during adulthood and tend to develop near the dermo-epidermal junction and toward the periphery of the CMN (Betti et al. 2000; Illig et al. 1985). In contrast, the risk of developing cutaneous melanoma in CMN larger than 20 cm is between 2% and 5% (Alikhan et al. 2012). These melanomas tend to develop early in life and develop below the dermo-epidermal junction. In one study, the mean age at diagnosis of melanoma was 15.5 years (median age, 7 years) (Alikhan et al. 2012; DeDavid et al. 1997; Kinsler et al. 2009; Krengel et al. 2006). Another study showed that the risk for developing melanoma in a CMN was higher in patients with MRI abnormalities of the central nervous system (CNS, OR of all-site melanoma 16.7 [95% CI 3.0–92.3]) and that this proved to be a better predictor of melanoma risk than the clinical phenotype of the CMN (Kinsler et al. 2009; 2017a; Krengel 2005; Krengel et al. 2006; Waelchli et al. 2015a).

Solar Lentigo as Precursor in Xeroderma Pigmentosum Patients Xeroderma pigmentosum (XP) confers the highest risk for developing melanoma (1000 times higher than the general population). XP patients are unable to repair UV-induced DNA lesions due to a defect in nucleotide excision repair (Hawkes et al. 2016). There is evidence to suggest solar lentigo as a possible melanoma precursor in XP patients. In one study, solar lentigo appeared contiguous with the patient’s melanoma in 88% of cases. A significantly lower number of contiguous solar lentigines were observed in the basal cell carcinoma, which served as controls (22%). The authors proposed that solar lentigo may be “the most common” melanoma precursor lesion XP patients (Stern et al. 1993).

C. Navarrete-Dechent et al.

Other Yet-to-Be-Defined Potential Intermediate Lesions It is now well acknowledged that almost every nevus has a bona fide cancer-associated oncogenic mutation (e.g., BRAF*, NRAS*, GNAQ*) (Rogers et al. 2016). It is highly likely that oncogene-induced senescence is the gatekeeper that prevents most nevi from progressing down the path of malignancy. The factors or combination of mutations that eventuates in a breakdown of senescence remains to be elucidated. In the future, testing of nevi for specific somatic mutations may help better stratify risk for developing melanoma. While most studied nevi express BRAF V600E mutations, a recent study found that atypical nevi had an increased number of additional mutations, including mutations involving the TERT gene. Early melanomas displayed an even larger number of mutations, including mutations in genes controlling the G1-S cell cycle, mainly CDKN2A; additional TP53 and PTEN mutations were also found in thick, invasive melanomas. The morphology of nevi with TERT and other mutations remain to be elucidated, and their potential for melanoma development needs to be investigated (Shain et al. 2015). In summary, the study by Shain et al. provides insights into the molecular underpinning of potential “true” precursor lesions; however, the clinical morphology of this subgroup of melanocytic lesions remains to be determined. Theoretically, if true intermediary nevi could be identified, then prophylactic removal of these lesions may prevent the occurrence of melanoma. In Fig. 12, we illustrate the current understanding of the universe of melanocytic neoplasms (Figs. 13, 14, and 15).

Management of High-Risk Patients Individuals with Specific Subtypes of Nevi Congenital Melanocytic Nevi The majority of CMN are of no concern medically and cosmetically. While melanomas may develop in association with small CMN, the risk of any

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

515

Fig. 12 Melanoma in situ arising in association with a Clark/dysplastic nevus in a 29-year-old female. The inset shows the clinical features. Note the presence of multiple nevi in the patient (the white arrowhead is showing the lesion). (Reproduced with permission from ISIC archive (www.isic-archive. com). Courtesy of Dr. Harold Rabinovitz)

Fig. 13 Melanoma in situ (white arrows) arising in association with a Clark/ dysplastic nevus (white asterisk) in a 66-year-old male. The inset is showing the clinical features (white arrowhead). (Reproduced with permission from ISIC archive (www.isic-archive. com). Courtesy of Dr. Harold Rabinovitz)

given small CMN transforming into melanoma is low. In addition, if a melanoma were to develop, it tends to develop later in life and tends to form at the dermo-epidermal junction making them easy to detect based on change (Illig et al. 1985). Thus, while CMN may at times be located in esthetically sensitive areas prompting surgical removal, small CMN in general do not need to be prophylactically excised and can be monitored instead. Monitoring for change can be facilitated by having baseline clinical and dermoscopic images of the small CMN. Small CMN manifesting concerning changes or displaying morphological criteria consistent with melanoma should be excised.

Large CMN are at risk for developing melanoma within the CMN. In a subset of patients with large CMN and neurocutaneous melanocytosis (NCM), primary central nervous system (CNS) melanoma accounts for approximately 1/3 of all melanomas (Neuhold et al. 2015). NCM is defined as the abnormal proliferation of melanocytes or the presence of melanin deposits along the CNS, within leptomeninges and/or the brain parenchyma (Kadonaga and Frieden 1991). Large CMN with multiple smaller satellite CMN are at highest risk for NCM. Marghoob et al. showed that patients with large CMN and > 20 satellites had a 5.1-fold increased risk for NCM compared with patients

516

C. Navarrete-Dechent et al.

Fig. 14 BAP1-inactivated melanocytic tumors. (a) and (b) show bapomas. Note the presence of an unspecific flesh-colored soft papule. Both patients had multiple

lesions leading to the diagnosis of a BAP-1 germline mutation. (Reproduced with permission from ISIC archive. www.isic-archive.com)

Fig. 15 Melanocytic lesions can be divided into two subgroups. The benign subgroup (highlighted in green) and the malignant subgroup (highlighted in red). Within the benign subgroup, there are recognizable nevi that display characteristic clinical, histopathological, and molecular features. Some of these nevi may have overlapping features (e.g., BAP1-deficient tumors share features seen in

intradermal nevi and Spitz nevi). Theoretically, there may be a smaller subgroup of nevi that are intermediate lesions (yellow circle) with a higher risk for malignant transformation. *The existence of such subgroup, as well as their clinical and dermoscopic morphology still needs to be elucidated. AIMP: atypical intraepithelial melanocytic proliferation

with large CMN and  20 satellites (Marghoob et al. 2004). Contrast-enhanced MRI of the brain and whole spine is the imaging modality of choice for the evaluation of suspected NCM and should ideally

be performed before 6 months of age since myelination can make it difficult to detect melanin deposits (Alikhan et al. 2012; Barkovich et al. 1994). Primary CNS melanoma in individuals with

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

CMN can present in different ways. The most common presentation is as a leptomeningeal mass and, less commonly, as a solid tumor within the brain parenchyma (Kinsler et al. 2017a). Patients usually present with symptoms of increased intracranial pressure or focal neurological symptoms or seizures; the prognosis in these cases is poor, and thus consideration for the prophylactic excision of the cutaneous large CMN should be deferred (Alikhan et al. 2012; Kinsler et al. 2017a). Patients with NCM that remain asymptomatic tend to have a good prognosis (Foster et al. 2001). Surgical interventions of the CNS melanoma plays an important palliative role, but it is usually not curative (Kinsler et al. 2017a). In addition, since the majority of melanomas arising in CMN will have an NRAS mutation, BRAF inhibitors are currently contraindicated due to paradoxical activation of NRAS pathway by these drugs (Weeraratna 2012). MEK inhibitors (e.g., trametinib) have shown some objective responses in small case series (Kinsler et al. 2017b). Large CMN are at increased risk for developing a cutaneous melanoma within the CMN, and this risk is less than 5%. Hence, prophylactic excision of the large CMN, if feasible, in patients without NCM should be discussed. Most melanomas developing in association with large CMN develop before puberty and develop in the dermis, thus making their early detection via visual inspection (including dermoscopy) virtually impossible. Palpation of these lesions for the development of new subcutaneous lumps can help detect these melanomas, but unfortunately these tumors are palpable when they are often already in an advanced stage. It is because of our inability to consistently find early-stage melanoma via physical examination and imaging that prophylactic excision remains a management option to consider and discuss. Naturally, there are many factors that need to be considered in this decisionmaking process, including the risk of melanoma, cosmetic concerns of the large CMN or resulting surgical scars, and the risk of anesthesia, among others.

Spitz Nevi Dermoscopy has greatly improved upon our ability to identify Spitz nevi. These nevi can be

517

divided into those manifesting a starburst pattern (Reed nevi; Fig. 8) and those without such a pattern (Spitz nevi) (Lallas et al. 2017). The risk of a Spitz or Reed nevus transforming into a melanoma is unknown, but it is presumed to be extremely low. The main immediate conundrum with these nevi is the fact that melanoma can occasionally masquerade itself as a Spitz nevus (Lallas et al. 2015). At times, it may be difficult to nearly impossible to differentiate Spitz nevi from melanoma, both clinically and dermoscopically, as well as histopathologically (Lallas et al. 2015). Reed nevi are commonly seen in children and, if encountered in the preadolescent years, they can be left alone (Lallas et al. 2017). The dermoscopic hallmark of Reed nevi is a symmetric starburst pattern; it is extremely uncommon for a melanocytic lesion with a starburst pattern in a child to be melanoma (Lallas et al. 2017). The approach to these Reed nevi – whether they should be monitored closely during their growth phase or simply left alone – is still being studied. A study by Lallas et al. showed that if these flat pigmented Spitz and Reed nevi in children, 21% of these lesions will change in a way that requieres an excision (Lallas et al. 2018). Interestingly, none of the Reed nevi in preadolescent children proved to be melanoma on histopathology (Lallas et al. 2015). In contrast, in adults, melanoma may masquerade as a Reed nevus both clinically (symmetric pigmented papule) and dermoscopically (presence of a symmetric, regular starburst pattern), and, hence, all such lesions should be excised in older individuals (Lallas et al. 2015, 2017). The Spitz nevi that lack a starburst pattern under dermoscopy are classified as typical or atypical Spitz nevi. Many of these lesions are pink in color and nodular in contour. The general rule is that these pink nodular lesions should be biopsied since melanoma can manifest the same morphology (Lallas et al. 2017). Similar to Reed nevi, the pretest probability of melanoma increases with age. Thus, Spitz nevi detected in older individuals should be excised (Lallas et al. 2015). In addition, all Spitz nevi that look like melanoma or present atypical features (e.g., multiple colors, asymmetry of pattern) should be

518

excised regardless of age. The dilemma is the management approach to pink Spitz nevi that develop during childhood. While most of these lesions will not be melanoma, childhood melanoma can present with a similar morphology (Carrera et al. 2018). This raises the question about the rationale for monitoring, since one of the cardinal rules of dermoscopy is never to monitor equivocal nodular lesions for the fear of missing a rapidly growing nodular melanoma (Liu et al. 2006). If we follow this rule, then all of these lesions should be excised. However, we know that (1) most of these pink nodular Spitzoid lesions in young children will prove to be benign on histopathology; (2) that some Spitz nevi may eventually involute; and (3) excision will leave the patient with a permanent scar due to the unnecessary removal of a benign nevus. Unfortunately, some of these benign Spitz nevi can also grow to become quite large without showing signs of involution. Clinicians electing to follow “banal” appearing Spitzoid lesions can use growth dynamics as a means to decide when to intervene with a biopsy. Lesions showing signs of excessive growth should promptly be excised while they are still relatively small, to limit the size of the resultant scar. In conclusion, the management of Spitz nevi remains a quagmire and needs to be managed on a case-by-case basis.

Patients with BAP1-Inactivated Melanocytic Tumors As discussed above, there are no formal guidelines regarding the management of patients with BAP1-inactivated tumors. Recommendations by experts have stated that it may be reasonable to test for BAP1 germline mutations in individuals harboring two or more bapomas or if there is positive family history of BAP1-associated tumors. If the patient is found to carry a germline mutation, screening by specific specialties and a multidisciplinary team should be discussed. This would include at least once-a-year total body skin examination for cutaneous melanoma by a dermatologist, as well as uveal melanoma screening by an ophthalmologist. Further testing (e.g. CT, MRI, etc) and referral (e.g. urology, respiratory

C. Navarrete-Dechent et al.

medicine, among others) should be discussed on a case-by-case basis (Haugh et al. 2017). In patients with isolated BAP1-inactivated tumors, reassurance and yearly follow-up are recommended.

Individuals with Many Nevi and Clark Nevi Patients with Single to a Few Clark Nevi As mentioned previously, the risk of a nevus transforming into a melanoma is extremely low, and prophylactic excision of Clark nevi is futile (Pampena et al. 2017; Tsao et al. 2003). However, when the lesion manifests suspect clinical and dermoscopic features that overlap with melanoma, a biopsy may be warranted to rule out melanoma. Alternatively, if the lesion is (1) non-palpable; (2) lacks melanomaspecific criteria on dermoscopy; (3) and it has history of being stable; then short-term digital dermoscopic monitoring (3–4 months) and sequential long-term dermoscopic follow-up is an acceptable alternative to excision. Any suspicious change noted during monitoring would warrant a biopsy, while Clark nevi that are stable do not require biopsy (Menzies et al. 2001). Patients with Many Nevi and Many Clark Nevi (So-Called Atypical Mole Syndrome) A subset of high-risk patients that is difficult to manage includes those with numerous nevi (often >100 nevi) that also display marked variability in size, shape, and color of their nevi (Fig. 7). In addition, many of these nevi will display morphologic features overlapping with those encountered in early melanomas. To complicate matters, some melanomas developing in these individuals will fail to manifest clinically recognizable ABCD features, fail to be overt outlier (“ugly duckling”) lesions, and fail to display any dermoscopic melanoma-specific features; thus, the challenge is to identify the early melanomas from among an ocean full of nevi (Puig et al. 2007; Yagerman et al. 2014). The systematic excision of all nevi in high-risk patients has been advocated by some (Brod et al. 2009; Cohen et al. 1991). The

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma

underlying rationale for this is the removal of all potential precursor lesions and the removal of any unsuspected melanoma. However, with the exception of some specific genetically prone patients, most high-risk patients will never develop a melanoma in their lifetime (cumulative 10-year risk of around 7–20%), and most melanomas (approximately 70%) in these individuals will develop de novo (Fears et al. 2006; Greene et al. 1985; Halpern et al. 2014; Marghoob et al. 1994; Moloney et al. 2014; Salerni et al. 2012a, b; Soura et al. 2016). Thus, prophylactic excision of nevi cannot eliminate the risk for developing melanoma, since only 30% of melanomas arise in association with a nevus (see previous section on “Precursor Lesions”) (Pampena et al. 2017; Tsao et al. 2003). In addition, the prophylactic removal of all nevi would continuously require the excision of all new nevi developing throughout the lifetime of the individual (Oliveria et al. 2013). So what is the alternative to whole-scale nevus excision? The answer is surveillance and monitoring (Menzies et al. 2001; Tucker et al. 2002). Since most nevi in high-risk adult patients are stable, and in light of the evidence that surveillance with the use of total body photography (TBP) and dermoscopy is associated with finding early melanomas, this remains the most practical management approach (Salerni et al. 2012b; Truong et al. 2016; Tucker et al. 2002; Zalaudek et al. 2010). The power in monitoring lies in the fact that most nevi will not change, while melanomas will display change over time and will be detectable during subsequent follow-up visits (Scope et al. 2016). This allows for maintaining a high sensitivity for melanoma detection while limiting the number of nevi excised. However, it is important to underscore that while “change” is a sensitive feature of melanoma, it is not specific and can also be seen in nevi, especially in younger patients (Bajaj et al. 2015; Banky et al. 2005; Liu et al. 2005; Yagerman et al. 2014). Fortunately, we have the ability to utilize dermoscopy and digital dermoscopic monitoring, as well as other imaging techniques, such as confocal microscopy (see also chapter ▶ “Dermoscopy/Confocal Microscopy for Melanoma Diagnosis”), to evaluate both the morphology and growth dynamics (i.e., biology)

519

of melanocytic neoplasms. Using a combined approach allows clinicians to subject to biopsy only lesions with morphologic features or growth dynamics concerning for melanoma, raising our specificity while still maintaining a high sensitivity (Kittler et al. 2006; Menzies et al. 2001; Rajadhyaksha et al. 2017; Salerni et al. 2013).

Conclusion Herein we have reviewed the factors that predispose individuals to melanoma and examined lesions that are considered to be potential melanoma precursor lesions. By identifying these high-risk patients and lesions, it is the hope that targeted interventions can be implemented to permit for the early detection of melanoma or for the prevention of this highly lethal skin cancer.

References Alikhan A, Ibrahimi OA, Eisen DB (2012) Congenital melanocytic nevi: where are we now? Part I Clinical presentation, epidemiology, pathogenesis, histology, malignant transformation, and neurocutaneous melanosis. J Am Acad Dermatol 67:495.e1–495.e17; quiz 512-494. https://doi.org/10.1016/j.jaad.2012. 06.023 Andersen WK, Labadie RR, Bhawan J (1997) Histopathology of solar lentigines of the face: a quantitative study. J Am Acad Dermatol 36:444–447 Annessi G et al (2001) Correlation between clinical atypia and histologic dysplasia in acquired melanocytic nevi. J Am Acad Dermatol 45:77–85. https://doi.org/ 10.1067/mjd.2001.114580 Autier P, Koechlin A, Boniol M (2015) The forthcoming inexorable decline of cutaneous melanoma mortality in light-skinned populations. Eur J Cancer 51:869–878. https://doi.org/10.1016/j.ejca.2015.01.056 Bajaj S et al (2015) Growth-curve modeling of nevi with a peripheral globular pattern. JAMA Dermatol 151:1338–1345. https://doi.org/10.1001/jamadermatol .2015.2231 Banky JP, Kelly JW, English DR, Yeatman JM, Dowling JP (2005) Incidence of new and changed nevi and melanomas detected using baseline images and dermoscopy in patients at high risk for melanoma. Arch Dermatol 141:998–1006. https://doi.org/10.10 01/archderm.141.8.998 Barkovich AJ, Frieden IJ, Williams ML (1994) MR of neurocutaneous melanosis. AJNR Am J Neuroradiol 15:859–867

520 Bastian BC, LeBoit PE, Pinkel D (2000) Mutations and copy number increase of HRAS in Spitz nevi with distinctive histopathological features. Am J Pathol 157:967–972. https://doi.org/10.1016/S0002-9440(10) 64609-3 Bauer J, Curtin JA, Pinkel D, Bastian BC (2007) Congenital melanocytic nevi frequently harbor NRAS mutations but no BRAF mutations. J Invest Dermatol 127:179–182. https://doi.org/10.1038/sj.jid.5700490 Bett BJ (2005) Large or multiple congenital melanocytic nevi: occurrence of cutaneous melanoma in 1008 persons. J Am Acad Dermatol 52:793–797. https://doi.org/ 10.1016/j.jaad.2005.02.024 Betti R, Inselvini E, Vergani R, Crosti C (2000) Small congenital nevi associated with melanoma: case reports and considerations. J Dermatol 27:583–590 Boniol M, Autier P, Gandini S (2015) Melanoma mortality following skin cancer screening in Germany. BMJ Open 5:e008158. https://doi.org/10.1136/bmjopen-20 15-008158 Breitbart EW, Waldmann A, Nolte S, Capellaro M, Greinert R, Volkmer B, Katalinic A (2012) Systematic skin cancer screening in Northern Germany. J Am Acad Dermatol 66:201–211. https://doi.org/10.1016/j.jaad.2 010.11.016 Brod C, Schippert W, Breuninger H (2009) Dysplastic nevus syndrome with development of multiple melanomas. A surgical concept for prophylaxis. J Dtsch Dermatol Ges 7:773–775. https://doi.org/10.1111/ j.1610-0387.2009.07095.x Burnett MEM, Ashfaq A, Scope A (2012) Nevogenesis: changing theories. In: Marghoob A (ed) Nevogenesis, mechanisms and clinical implications of nevus development. Springer, Berlin/Heidelberg, pp 1–23. https://doi.org/10.1007/978-3-642-28397-0 Busam KJ, Kutzner H, Cerroni L, Wiesner T (2014) Clinical and pathologic findings of Spitz nevi and atypical Spitz tumors with ALK fusions. Am J Surg Pathol 38:925–933. https://doi.org/10.1097/PAS.0000000000 000187 Campbell LB, Kreicher KL, Gittleman HR, Strodtbeck K, Barnholtz-Sloan J, Bordeaux JS (2015) Melanoma incidence in children and adolescents: decreasing trends in the United States. J Pediatr 166:1505–1513. https://doi.org/10.1016/j.jpeds.2015.02.050 Carrera C et al (2018) Clinical and dermoscopic characterization of pediatric and adolescent melanomas: multicenter study of 52 cases. J Am Acad Dermatol 78:278–288. https://doi.org/10.1016/j.jaad. 2017.09.065 Castilla EE, da Graca Dutra M, Orioli-Parreiras IM (1981) Epidemiology of congenital pigmented naevi: I. Incidence rates and relative frequencies. Br J Dermatol 104:307–315 Clark WH Jr, Reimer RR, Greene M, Ainsworth AM, Mastrangelo MJ (1978) Origin of familial malignant melanomas from heritable melanocytic lesions. ‘The BK mole syndrome’. Arch Dermatol 114:732–738

C. Navarrete-Dechent et al. Cohen MH, Cohen BJ, Shotkin JD, Morrison PT (1991) Surgical prophylaxis of malignant melanoma. Ann Surg 213:308–314 DeDavid M et al (1997) A study of large congenital melanocytic nevi and associated malignant melanomas: review of cases in the New York University registry and the world literature. J Am Acad Dermatol 36:409–416 Elwood JM, Whitehead SM, Davison J, Stewart M, Galt M (1990) Malignant melanoma in England: risks associated with naevi, freckles, social class, hair colour, and sunburn. Int J Epidemiol 19:801–810 Fears TR et al (2006) Identifying individuals at high risk of melanoma: a practical predictor of absolute risk. J Clin Oncol 24:3590–3596. https://doi.org/10.1200/ JCO.2005.04.1277 Fortes C et al (2010) Identifying individuals at high risk of melanoma: a simple tool. Eur J Cancer Prev 19:393–400. https://doi.org/10.1097/CEJ.0b013e 32833b492f Foster RD, Williams ML, Barkovich AJ, Hoffman WY, Mathes SJ, Frieden IJ (2001) Giant congenital melanocytic nevi: the significance of neurocutaneous melanosis in neurologically asymptomatic children. Plast Reconstr Surg 107:933–941 Frank SA (2007) Dynamics of cancer: incidence, inheritance, and evolution. Princeton University Press, Princeton Fukunaga-Kalabis M, Herlyn M (2012) Cancer: complexion matters. Nature 491:342–343. https://doi.org/10.10 38/nature11641 Fuller SR, Bowen GM, Tanner B, Florell SR, Grossman D (2007) Digital dermoscopic monitoring of atypical nevi in patients at risk for melanoma. Dermatol Surg 33:1198–1206; discussion 1205-1196. https://doi.org/ 10.1111/j.1524-4725.2007.33254.x Gandini S, Sera F, Cattaruzza MS, Pasquini P, Abeni D, Boyle P, Melchi CF (2005a) Meta-analysis of risk factors for cutaneous melanoma: I. Common and atypical naevi. Eur J Cancer 41:28–44. https://doi.org/10.1016/ j.ejca.2004.10.015 Gandini S et al (2005b) Meta-analysis of risk factors for cutaneous melanoma: III. Family history, actinic damage and phenotypic factors. Eur J Cancer (Oxford, England: 1990) 41:2040–2059. https://doi.org/10.101 6/j.ejca.2005.03.034 Glazer AM, Winkelmann RR, Farberg AS, Rigel DS (2016) Analysis of trends in US melanoma incidence and mortality. JAMA Derm. https://doi.org/10.1001/ jamadermatol.2016.4512 Greene MH, Clark WH Jr, Tucker MA, Kraemer KH, Elder DE, Fraser MC (1985) High risk of malignant melanoma in melanoma-prone families with dysplastic nevi. Ann Intern Med 102:458–465 Gudbjartsson DF et al (2008) ASIP and TYR pigmentation variants associate with cutaneous melanoma and basal cell carcinoma. Nat Genet 40:886–891. https://doi.org/ 10.1038/ng.161

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma Haenssle HA et al (2016) Association of patient risk factors and frequency of nevus-associated cutaneous melanomas. JAMA Dermatol 152:291–298. https://doi.org/ 10.1001/jamadermatol.2015.3775 Hale EK et al (2005) Association of melanoma and neurocutaneous melanocytosis with large congenital melanocytic naevi – results from the NYU-LCMN registry. Br J Dermatol 152:512–517. https://doi.org/ 10.1111/j.1365-2133.2005.06316.x Halpern AC, Guerry D, Elder DE, Trock B, Synnestvedt M, Humphreys T (1993) Natural history of dysplastic nevi. J Am Acad Dermatol 29:51–57 Halpern AC, Marchetti MA, Marghoob AA (2014) Melanoma surveillance in “high-risk” individuals. JAMA Dermatol 150:815–816. https://doi.org/10.10 01/jamadermatol.2014.513 Harris IS, Brugge JS (2015) Cancer: the enemy of my enemy is my friend. Nature 527:170–171. https://doi.org/10.1038/nature15644 Haugh AM et al (2017) Genotypic and phenotypic features of BAP1 cancer syndrome: a report of 8 new families and review of cases in the literature. JAMA Dermatol 153:999–1006. https://doi.org/10.1001/jamadermatol. 2017.2330 Hawkes JE, Truong A, Meyer LJ (2016) Genetic predisposition to melanoma. Semin Oncol 43:591–597. https://doi.org/10.1053/j.seminoncol.2016.08.003 Hofmann-Wellenhof R, Marghoob AA, Zalaudek I (2016) Large acquired nevus or dysplastic nevus: what’s in the name of a nevus? JAMA Dermatol 152:623–624. https://doi.org/10.1001/jamadermatol.2015.6186 Ichii-Nakato N et al (2006) High frequency of BRAFV600E mutation in acquired nevi and small congenital nevi, but low frequency of mutation in medium-sized congenital nevi. J Invest Dermatol 126:2111–2118. https://doi.org/10.1038/sj.jid.5700366 Idorn LW, Datta P, Heydenreich J, Philipsen PA, Wulf HC (2015) Black light visualized solar lentigines on the shoulders and upper back are associated with objectively measured UVR exposure and cutaneous malignant melanoma. Photochem Photobiol Sci 14:481–487. https://doi.org/10.1039/c4pp00332b Illig L, Weidner F, Hundeiker M, Gartmann H, Biess B, Leyh F, Paul E (1985) Congenital nevi less than or equal to 10 cm as precursors to melanoma. 52 cases, a review, and a new conception. Arch Dermatol 121:1274–1281 Ingordo V, Gentile C, Iannazzone SS, Cusano F, Naldi L (2007) Congenital melanocytic nevus: an epidemiologic study in Italy. Dermatology 214:227–230. https://doi.org/10.1159/000099587 Kadonaga JN, Frieden IJ (1991) Neurocutaneous melanosis: definition and review of the literature. J Am Acad Dermatol 24:747–755 Kanada KN, Merin MR, Munden A, Friedlander SF (2012) A prospective study of cutaneous findings in newborns in the United States: correlation with race, ethnicity, and gestational status using updated classification and

521

nomenclature. J Pediatr 161:240–245. https://doi.org/ 10.1016/j.jpeds.2012.02.052 Katalinic A et al (2012) Does skin cancer screening save lives?: an observational study comparing trends in melanoma mortality in regions with and without screening. Cancer 118:5395–5402. https://doi.org/10. 1002/cncr.27566 Kinsler VA, Birley J, Atherton DJ (2009) Great Ormond Street Hospital for Children registry for congenital melanocytic naevi: prospective study 1988–2007. Part 1-epidemiology, phenotype and outcomes. Br J Dermatol 160:143–150. https://doi.org/10.1111/ j.1365-2133.2008.08849.x Kinsler VA et al (2013) Multiple congenital melanocytic nevi and neurocutaneous melanosis are caused by postzygotic mutations in codon 61 of NRAS. J Invest Dermatol 133:2229–2236. https://doi.org/10.1038/jid.2013.70 Kinsler VA et al (2017a) Melanoma in congenital melanocytic naevi. Br J Dermatol 176:1131–1143. https://doi.org/10.1111/bjd.15301 Kinsler VA, O’Hare P, Jacques T, Hargrave D, Slater O (2017b) MEK inhibition appears to improve symptom control in primary NRAS-driven CNS melanoma in children. Br J Cancer 116:990–993. https://doi.org/10.1038/bjc.2017.49 Kittler H, Tschandl P (2013) Dysplastic nevus: why this term should be abandoned in dermatoscopy. Dermatol Clin 31:579–588, viii. https://doi.org/10.1016/j.det. 2013.06.009 Kittler H, Pehamberger H, Wolff K, Binder M (2000a) Follow-up of melanocytic skin lesions with digital epiluminescence microscopy: patterns of modifications observed in early melanoma, atypical nevi, and common nevi. J Am Acad Dermatol 43:467–476. https://doi.org/10.1067/mjd.2000.107504 Kittler H, Seltenheim M, Dawid M, Pehamberger H, Wolff K, Binder M (2000b) Frequency and characteristics of enlarging common melanocytic nevi. Arch Dermatol 136:316–320 Kittler H et al (2006) Identification of clinically featureless incipient melanoma using sequential dermoscopy imaging. Arch Dermatol 142:1113–1119. https://doi.org/10.1001/archderm.142.9.1113 Klein LJ, Barr RJ (1990) Histologic atypia in clinically benign nevi. A prospective study. J Am Acad Dermatol 22:275–282 Kopf AW, Levine LJ, Rigel DS, Friedman RJ, Levenstein M (1985a) Congenital-nevus-like nevi, nevi spili, and cafe-au-lait spots in patients with malignant melanoma. J Dermatol Surg Oncol 11:275–280 Kopf AW, Levine LJ, Rigel DS, Friedman RJ, Levenstein M (1985b) Prevalence of congenitalnevus-like nevi, nevi spili, and cafe au lait spots. Arch Dermatol 121:766–769 Krengel S (2005) Nevogenesis – new thoughts regarding a classical problem. Am J Dermatopathol 27:456–465 Krengel S, Hauschild A, Schafer T (2006) Melanoma risk in congenital melanocytic naevi: a systematic review.

522 Br J Dermatol 155:1–8. https://doi.org/10.1111/j.13652133.2006.07218.x Krengel S, Scope A, Dusza SW, Vonthein R, Marghoob AA (2013) New recommendations for the categorization of cutaneous features of congenital melanocytic nevi. J Am Acad Dermatol 68:441–451. https://doi.org/10.1016/j.jaad.2012.05.043 Lallas A et al (2015) Likelihood of finding melanoma when removing a Spitzoid-looking lesion in patients aged 12 years or older. J Am Acad Dermatol 72:47–53. https:// doi.org/10.1016/j.jaad.2014.09.037 Lallas A et al (2017) Update on dermoscopy of Spitz/reed naevi and management guidelines by the international dermoscopy society. Br J Dermatol 177:645–655. https://doi.org/10.1111/bjd.15339 Lallas A, Apalla Z, Papageorgiou C, Evangelou G, Ioannides D, Argenziano G (2018) Management of Flat Pigmented Spitz and Reed Nevi in Children. JAMA Dermatol 154(11):1353–1354 Liotta LA, Kohn EC (2003) Cancer’s deadly signature. Nat Genet 33:10–11. https://doi.org/10.1038/ng0103-10 Liu W et al (2005) What features do patients notice that help to distinguish between benign pigmented lesions and melanomas?: the ABCD(E) rule versus the sevenpoint checklist. Melanoma Res 15:549–554 Liu W, Dowling JP, Murray WK, McArthur GA, Thompson JF, Wolfe R, Kelly JW (2006) Rate of growth in melanomas: characteristics and associations of rapidly growing melanomas. Arch Dermatol 142:1551–1558. https://doi.org/10.1001/archderm.142.12.1551 Marghoob AA et al (1994) Risk of cutaneous malignant melanoma in patients with ‘classic’ atypical-mole syndrome. A case-control study. Arch Dermatol 130:993–998 Marghoob AA, Dusza S, Oliveria S, Halpern AC (2004) Number of satellite nevi as a correlate for neurocutaneous melanocytosis in patients with large congenital melanocytic nevi. Arch Dermatol 140:171–175. https://doi.org/10.1001/archd erm.140.2.171 McLean DI, Gallagher RP (1995) “Sunburn” freckles, cafe-au-lait macules, and other pigmented lesions of schoolchildren: the Vancouver Mole Study. J Am Acad Dermatol 32:565–570 Menzies SW, Gutenev A, Avramidis M, Batrac A, McCarthy WH (2001) Short-term digital surface microscopic monitoring of atypical or changing melanocytic lesions. Arch Dermatol 137:1583–1589 Mitra D et al (2012) An ultraviolet-radiation-independent pathway to melanoma carcinogenesis in the red hair/fair skin background. Nature 491:449–453. https://doi.org/10.1038/nature11624 Moawad S et al (2018) Clinical, dermoscopic, histological and molecular analysis of BAP1 inactivated melanocytic nevus/tumor in two familial cases of BAP1 syndrome. Br J Dermatol. https://doi.org/10.1 111/bjd.16740 Moloney FJ et al (2014) Detection of primary melanoma in individuals at extreme high risk: a prospective 5-year

C. Navarrete-Dechent et al. follow-up study. JAMA Dermatol 150:819–827. https://doi.org/10.1001/jamadermatol.2014.514 Morgan AM, Lo J, Fisher DE (2013) How does pheomelanin synthesis contribute to melanomagenesis?: two distinct mechanisms could explain the carcinogenicity of pheomelanin synthesis. BioEssays 35:672–676. https://doi.org/10.1002/bies. 201300020 Neuhold JC, Friesenhahn J, Gerdes N, Krengel S (2015) Case reports of fatal or metastasizing melanoma in children and adolescents: a systematic analysis of the literature. Pediatr Dermatol 32:13–22. https://doi.org/ 10.1111/pde.12400 Newton-Bishop JA et al (2010) Melanocytic nevi, nevus genes, and melanoma risk in a large case-control study in the United Kingdom. Cancer Epidemiology Biomarkers Prev 19:2043–2054. https://doi.org/10.1158/ 1055-9965.epi-10-0233 Nicholls EM (1973) Development and elimination of pigmented moles, and the anatomical distribution of primary malignant melanoma. Cancer 32:191–195 Oliveria SA, Yagerman SE, Jaimes N, Goodwin AI, Dusza SW, Halpern AC, Marghoob AA (2013) Clinical and dermoscopic characteristics of new naevi in adults: results from a cohort study. Br J Dermatol 169:848–853. https://doi.org/10.1111/bjd.12482 Osterlind A, Tucker MA, Hou-Jensen K, Stone BJ, Engholm G, Jensen OM (1988) The Danish case-control study of cutaneous malignant melanoma. I. Importance of host factors. Int J Cancer 42:200–206 Pampena R, Kyrgidis A, Lallas A, Moscarella E, Argenziano G, Longo C (2017) A meta-analysis of nevus-associated melanoma: prevalence and practical implications. J Am Acad Dermatol 77:938–945 e934. https://doi.org/10.1016/j.jaad.2017.06.149 Piskounova E et al (2015) Oxidative stress inhibits distant metastasis by human melanoma cells. Nature 527:186–191. https://doi.org/10.1038/nature15726 Potrony M, Badenas C, Aguilera P, Puig-Butille JA, Carrera C, Malvehy J, Puig S (2015) Update in genetic susceptibility in melanoma. Ann Transl Med 3:210. https://doi.org/10.3978/j.issn.2305-5839.2015.08.11 Puig S et al (2007) Melanomas that failed dermoscopic detection: a combined clinicodermoscopic approach for not missing melanoma. Dermatol Surg 33:1262–1273. https://doi.org/10.1111/j.1524-4725.2007.33264.x Raimondi S, Sera F, Gandini S, Iodice S, Caini S, Maisonneuve P, Fargnoli MC (2008) MC1R variants, melanoma and red hair color phenotype: a meta-analysis. Int J Cancer 122:2753–2760. https://doi.org/ 10.1002/ijc.23396 Rajadhyaksha M, Marghoob A, Rossi A, Halpern AC, Nehal KS (2017) Reflectance confocal microscopy of skin in vivo: from bench to bedside. Lasers Surg Med 49:7–19. https://doi.org/10.1002/lsm.22600 Ramaswamy S, Ross KN, Lander ES, Golub TR (2003) A molecular signature of metastasis in primary solid tumors. Nat Genet 33:49–54. https://doi.org/10.1038/ ng1060

Acquired Precursor Lesions and Phenotypic Markers of Increased Risk for Cutaneous Melanoma Rhodes AR (1986) Melanocytic precursors of cutaneous melanoma. Estimated risks and guidelines for management. Med Clin North Am 70:3–37 Roesch A, Burgdorf W, Stolz W, Landthaler M, Vogt T (2006) Dermatoscopy of “dysplastic nevi”: a beacon in diagnostic darkness. Eur J Dermatol 16:479–493 Rogers T, Marino ML, Raciti P, Jain M, Busam KJ, Marchetti MA, Marghoob AA (2016) Biologically distinct subsets of nevi. G Ital Dermatol Venereol 151:365–384 Roh MR, Eliades P, Gupta S, Tsao H (2015) Genetics of melanocytic nevi. Pigment Cell Melanoma Res 28:661–672. https://doi.org/10.1111/pcmr.12412 Roider EM, Fisher DE (2016) Red hair, light skin, and UVindependent risk for melanoma development in humans. JAMA Dermatol 152:751–753. https://doi. org/10.1001/jamadermatol.2016.0524 Rosendahl CO, Grant-Kels JM, Que SK (2015) Dysplastic nevus: fact and fiction. J Am Acad Dermatol 73:507–512. https://doi.org/10.1016/j.jaad. 2015.04.029 Salerni G et al (2012a) Characterization of 1152 lesions excised over 10 years using total-body photography and digital dermatoscopy in the surveillance of patients at high risk for melanoma. J Am Acad Dermatol 67:836–845. https://doi.org/10.1016/j.jaad. 2012.01.028 Salerni G et al (2012b) Benefits of total body photography and digital dermatoscopy (“two-step method of digital follow-up”) in the early diagnosis of melanoma in patients at high risk for melanoma. J Am Acad Dermatol 67:e17–e27. https://doi.org/10.1016/j.jaad. 2011.04.008 Salerni G, Terán T, Puig S, Malvehy J, Zalaudek I, Argenziano G, Kittler H (2013) Meta-analysis of digital dermoscopy follow-up of melanocytic skin lesions: a study on behalf of the international dermoscopy society. J Eur Acad Dermatol Venereol 27:805–814. https://doi.org/10.1111/jdv.12032 Schafer T, Merkl J, Klemm E, Wichmann HE, Ring J, KORA Study Group (2006) The epidemiology of nevi and signs of skin aging in the adult general population: results of the KORA-survey 2000. J Invest Dermatol 126:1490–1496. https://doi.org/10.1038/sj. jid.5700269 Scope A et al (2011) Clinical and dermoscopic stability and volatility of melanocytic nevi in a population-based cohort of children in Framingham school system. J Invest Dermatol 131:1615–1621. https://doi.org/ 10.1038/jid.2011.107 Scope A et al (2016) The study of nevi in children: principles learned and implications for melanoma diagnosis. J Am Acad Dermatol 75:813–823. https://doi. org/10.1016/j.jaad.2016.03.027 Sepehr A et al (2011) Long-term outcome of Spitz-type melanocytic tumors. Arch Dermatol 147:1173–1179. https://doi.org/10.1001/archdermatol.2011.170 Shain AH et al (2015) The genetic evolution of melanoma from precursor lesions the New England.

523

J Med 373:1926–1936. https://doi.org/10.1056/NEJM oa1502583 Siegel RL, Miller KD, Jemal A (2018) Cancer statistics, 2018. CA Cancer J Clin 68:7–30. https://doi.org/10. 3322/caac.21442 Soura E, Eliades PJ, Shannon K, Stratigos AJ, Tsao H (2016) Hereditary melanoma: update on syndromes and management: genetics of familial atypical multiple mole melanoma syndrome. J Am Acad Dermatol 74:395–407; quiz 408-310. https://doi.org/10.1016/j.jaad.2015.08.038 Stern JB, Peck GL, Haupt HM, Hollingsworth HC, Beckerman T (1993) Malignant melanoma in xeroderma pigmentosum: search for a precursor lesion. J Am Acad Dermatol 28:591–594 Sturm RA et al (2003) The role of melanocortin-1 receptor polymorphism in skin cancer risk phenotypes. Pigment Cell Res 16:266–272 Tan JM et al (2017) The BRAF and NRAS mutation prevalence in dermoscopic subtypes of acquired naevi reveals constitutive mitogen-activated protein kinase pathway activation. Br J Dermatol. https://doi.org/ 10.1111/bjd.15809 Terushkin V et al (2012) Changes observed in slow-growing melanomas during long-term dermoscopic monitoring. Br J Dermatol 166:1213–1220. https://doi.org/ 10.1111/j.1365-2133.2012.10846.x Truong A, Strazzulla L, March J, Boucher KM, Nelson KC, Kim CC, Grossman D (2016) Reduction in nevus biopsies in patients monitored by total body photography. J Am Acad Dermatol 75:135–143.e135. https://doi.org/10.1016/j.jaad.2016.02.1152 Tsao H, Bevona C, Goggins W, Quinn T (2003) The transformation rate of moles (melanocytic nevi) into cutaneous melanoma: a population-based estimate. Arch Dermatol 139:282–288 Tucker MA et al (2002) A natural history of melanomas and dysplastic nevi: an atlas of lesions in melanomaprone families. Cancer 94:3192–3209. https://doi.org/ 10.1002/cncr.10605 Uribe P, Wistuba II, Gonzalez S (2003) BRAF mutation: a frequent event in benign, atypical, and malignant melanocytic lesions of the skin. Am J Dermatopathol 25:365–370 Vallone MG et al (2018) Melanocortin 1 receptor (MC1R) polymorphisms’ influence on size and dermoscopic features of nevi. Pigment Cell Melanoma Res 31:39–50. https://doi.org/10.1111/pcmr.12646 Valverde P, Healy E, Jackson I, Rees JL, Thody AJ (1995) Variants of the melanocyte-stimulating hormone receptor gene are associated with red hair and fair skin in humans. Nat Genet 11:328–330. https://doi.org/10.1038/ng1195-328 Vogelstein B et al (1988) Genetic alterations during colorectal-tumor development. N Engl J Med 319:525–532. https://doi.org/10.1056/NEJM19880901 3190901 Waelchli R, Aylett SE, Atherton D, Thompson DJ, Chong WK, Kinsler VA (2015a) Classification of neurological abnormalities in children with congenital

524 melanocytic naevus syndrome identifies magnetic resonance imaging as the best predictor of clinical outcome. Br J Dermatol 173:739–750. https://doi.org/ 10.1111/bjd.13898 Waelchli R et al (2015b) Growth and hormone profiling in children with congenital melanocytic naevi. Br J Dermatol 173:1471–1478. https://doi.org/10.1111/ bjd.14091 Walton RG, Jacobs AH, Cox AJ (1976) Pigmented lesions in newborn infants. Br J Dermatol 95:389–396 Weeraratna AT (2012) RAF around the edges – the paradox of BRAF inhibitors. N Engl J Med 366:271–273. https://doi.org/10.1056/NEJMe1111636 Wendt J, Mueller C, Rauscher S, Fae I, Fischer G, Okamoto I (2018) Contributions by MC1R variants to melanoma risk in males and females. JAMA Dermatol. https://doi.org/10.1001/jamadermatol.2018.1252 White E, Kirkpatrick CS, Lee JA (1994) Case-control study of malignant melanoma in Washington State. I Constitutional factors and sun exposure. Am J Epidemiol 139:857–868 Wiecker TS, Luther H, Buettner P, Bauer J, Garbe C (2003) Moderate sun exposure and nevus counts in parents are associated with development of melanocytic nevi in childhood: a risk factor study in 1,812 kindergarten children. Cancer 97:628–638. https://doi.org/10.1002/cncr.11114 Wiesner T et al (2011) Germline mutations in BAP1 predispose to melanocytic tumors. Nat Genet 43:1018–1021. https://doi.org/10.1038/ng.910

C. Navarrete-Dechent et al. Wiesner T, Murali R, Fried I, Cerroni L, Busam K, Kutzner H, Bastian BC (2012) A distinct subset of atypical Spitz tumors is characterized by BRAF mutation and loss of BAP1 expression. Am J Surg Pathol 36:818–830. https://doi.org/10.10 97/PAS.0b013e3182498be5 Wiesner T et al (2014) Kinase fusions are frequent in Spitz tumours and spitzoid melanomas. Nat Commun 5:3116. https://doi.org/10.1038/ncomms4116 Williams LH, Shors AR, Barlow WE, Solomon C, White E (2011) Identifying persons at highest risk of melanoma using self-assessed risk factors. J Clin Exp Dermatol Res 2. https://doi.org/10.4172/2155-9554.1000129 Yagerman SE, Chen L, Jaimes N, Dusza SW, Halpern AC, Marghoob A (2014) ‘Do UC the melanoma?’ Recognising the importance of different lesions displaying unevenness or having a history of change for early melanoma detection. Australas J Dermatol 55:119–124. https://doi.org/10.1111/ ajd.12143 Zaal LH, Mooi WJ, Klip H, van der Horst CM (2005) Risk of malignant transformation of congenital melanocytic nevi: a retrospective nationwide study from the Netherlands. Plast Reconstr Surg 116:1902–1909 Zalaudek I et al (2010) Who benefits from prophylactic surgical removal of “dysplastic” nevi? J Dtsch Dermatol Ges 8:279–280. https://doi.org/10.1111/ j.1610-0387.2010.07377.x

Melanoma Prevention and Screening Susan M. Swetter, Alan C. Geller, Sancy A. Leachman, John M. Kirkwood, Alexander Katalinic, and Jeffrey E. Gershenwald

Contents Prevention of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526 Trends in Melanoma Incidence and Mortality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527 Primary Prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reducing Personal Exposure: Shade, Clothing, and Sunscreens . . . . . . . . . . . . . . . . . . . . . . . Behavioral Change Programs for Reducing Personal Exposure . . . . . . . . . . . . . . . . . . . . . . . Multicomponent Community-Wide Interventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Youth Education and Counseling Programs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Intervention Trials of the Prevention of Nevi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Controlling Exposure to Indoor Tanning Beds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

528 529 531 532 533 535 537

S. M. Swetter (*) Department of Dermatology, Pigmented Lesion and Melanoma Program, Stanford University Medical Center and Cancer Institute, Stanford, CA, USA Dermatology Service, Veterans Affairs Palo Alto Health Care System, Palo Alto, CA, USA e-mail: [email protected] A. C. Geller Department of Social and Behavioral Sciences, Harvard TH Chan School of Public Health, Boston, MA, USA e-mail: [email protected] S. A. Leachman Department of Dermatology, Oregon Health and Science University, Portland, OR, USA e-mail: [email protected] J. M. Kirkwood Departments of Medicine, Dermatology, and Translational Science, University of Pittsburgh and UPMC Hillman Cancer Center, Pittsburgh, PA, USA e-mail: [email protected] © This is a U.S. Government work and not under copyright protection in the U.S.; foreign copyright protection may apply 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_6

525

526

S. M. Swetter et al. Therapeutic Prevention of Melanoma and Populations to Target for Interventional Trials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539 Therapeutic Prevention of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539 Candidate Agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539 Secondary Prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Early Detection and Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Potential Benefits of Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Uncertainties and Conflicts in Melanoma Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Screening-Related Harms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Prevalence of Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Evidence Relating to the Effectiveness of Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Economic Assessments of Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Programs of Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Adjuncts to Clinical Examination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Challenges in the Detection of More Lethal Melanoma Subtypes . . . . . . . . . . . . . . . . . . . . . Advances in Screening Technologies and Community Outreach to Improve Early Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

545 545 545 546 546 547 548 550 551 556 556 559

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559

Abstract

Primary and secondary prevention of melanoma are critical to reducing incidence and mortality rates. Primary prevention is focused on reducing the key established modifiable risk factor – exposure to ultraviolet A (UVA) and ultraviolet B (UVB) radiation from the sun and indoor tanning, which is responsible for the majority of cutaneous melanoma in light-skinned populations. Therapeutic prevention, also termed “chemoprevention,” is a type of primary prevention to avert the development of melanoma at the outset, using safe and tolerable oral or systemic agents, the ideal one(s) which have yet to be defined. Secondary prevention efforts are focused on early detection strategies to detect cutaneous melanoma in its earliest stages,

A. Katalinic Department of Medicine, University Lübeck, Institute for Social Medicine and Epidemiology, Lübeck, SchleswigHolstein, Germany e-mail: [email protected] J. E. Gershenwald Departments of Surgical Oncology and Cancer Biology, Melanoma and Skin Center, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected]

enhancing the likelihood of cure. Screening for melanoma, either by the individual (i.e., skin self-examination) or by the health-care provider, is likely to detect melanoma earlier, though a reduction of melanoma mortality has not yet been observed in most screening efforts worldwide. Technological and molecular advances in bedside melanoma diagnosis may aid screening efforts in the future.

Prevention of Melanoma It has been estimated that a comprehensive skin cancer prevention program in the United States could prevent 230,000 melanoma cases and $2.7 billion in initial year treatment costs from 2020 to 2030 (Guy et al. 2015). Primary prevention focuses on reducing the key established modifiable risk factor – exposure to ultraviolet A (UVA) and ultraviolet B (UVB) radiation from the sun and indoor tanning, which is responsible for the majority of cutaneous melanoma in white populations. Therapeutic prevention, previously termed “chemoprevention,” is a type of primary prevention to avert melanomagenesis, using safe and tolerable topical or systemic agents, the ideal one(s) which have not yet been defined.

Melanoma Prevention and Screening

527

Secondary prevention efforts involve early detection strategies to detect cutaneous melanoma in its earliest stages, enhancing the likelihood of cure. Screening for melanoma, either by the individual (i.e., skin self-examination) or by the health provider, is likely to detect melanoma earlier, though a reduction of melanoma mortality has not yet been observed in most screening efforts worldwide (Fig. 1).

Trends in Melanoma Incidence and Mortality The overall burden of cutaneous melanoma continues to increase, despite worldwide efforts to enhance primary prevention and early detection over the past several decades (see also chapter ▶ “Clinical Epidemiology of Melanoma”). Thinner melanomas and melanoma in situ have shown the greatest increases in incidence, although the absolute number of thicker, less curable melanomas continues to rise, particularly among lower socioeconomic status (SES) groups (Clarke et al. 2017; Glazer et al. 2016; Linos et al. 2009). Incidence rates are highest in non-Hispanic whites (NHW), who are more

susceptible than other racial-ethnic groups to the dangers of UVR, and also differ by age and sex. From 2005 to 2014, melanoma incidence rates increased in men older than 54 years and women older than 44 years, with slight decreases observed in younger NHWs (Holman et al. 2018), suggesting the potential effectiveness of primary prevention efforts in children, adolescents, and young adults. Recent slowing of annual population-adjusted melanoma mortality rates (from 2.0% per year to 1.5% in 2016) may be linked to early detection efforts, as these data do not likely represent advances in therapy for metastatic disease (Glazer et al. 2016). Shaikh and colleagues characterized melanoma thickness and survival trends among men and women in the Surveillance, Epidemiology, and End Results (SEER) 9 registries between 1989 and 2009 (Shaikh et al. 2015). Increased incidence occurred across all thickness groups, although the median melanoma thickness decreased (from 0.73 mm to 0.58 mm), and geometric mean thickness decreased by 4.6% (from 0.77 mm to 0.65 mm) (95% CI = 4.2–5.0%) every 3 years in multivariable analysis. Thickness decreased in T1/T2 tumors (0.01–1.00 and 1.01–2.00 mm) and among all age and sex groups,

Melanoma / Skin Cancer Control Primary Prevention Chemoprevention

Public Education, Publicity, and Policy Initiatives

Secondary Prevention = Early Detection

(therapeutic prevention)

Policy changes and legislation to reduce natural and artificial ultraviolet radiation exposure

Community-wide interventions

Modification of adverse sun exposure/indoor tanning behaviors

Fig. 1 Melanoma prevention overview

Skin Self-Examination and Patient Request for Clinician Skin Exam

Routine screening by health professionals (opportunistic screening)

Medical Diagnostic Visit (goal: earlier melanoma)

528

in whites, non-Hispanics, and at all body sites. However, thickness increased among T3/T4 tumors (2.01–4.00 and >4.00 mm) and for nodular melanomas; acral lentiginous melanomas approached statistical significance. Melanomaspecific survival improved (hazard ratio [HR] = 0.89, 95% CI = 0.88–0.91) every 3 years in multivariable analysis and across all subgroups except for nonblack minorities and nodular and acral lentiginous subtypes. Recently, using a more updated analysis of SEER data (1984–2014), Krauss and colleagues found that melanoma incidence for the overall white population increased by 140%, while rates increased by 226% among white men 50 (Krauss et al. 2018, unpublished work). From 1985 to 2013, total mortality rates increased by 9.33% with an annual percent change (APC) of 0.23%; however, from 2013 to 2015 rates decreased by 9.25% with an APC of 5.31. From 1985 to 2013, mortality rates in white men 50 increased by 44.7%, but for the 3 most recent years, mortality dropped by 10.3%, at an APC of 5.19% (95% CI: 8.6 to 1.6). Analysis of incidence trends for in situ and invasive melanoma during 1995–2012 using 18 European cancer registries (encompassing over 117 million inhabitants and about 415,000 skin lesions) showed a statistically significant increase for both invasive melanoma (average annual percent change [AAPC] 4.0% for men; 3.0% for women) and in situ cases (AAPC 7.7% men; 6.2% women) (Sacchetto et al. 2018). However, the increase in invasive lesions was primarily driven by thin melanomas (AAPC 10% for men; 8.3% for women), although the incidence of thick melanomas also increased over time, more slowly in recent years. Continued increases in mortality rates were noted in Norway, Iceland (for older individuals), the Netherlands, and Slovenia. A study sought to examine why Norway has the highest rate of melanoma mortality in Europe (Robsahm et al. 2018). The CMs were equally distributed between men (49.9%) and women (50.1%), and the trunk was the most common anatomic site (48%). Compared to women, men were diagnosed at an older age, with thicker and more ulcerated tumors that were more often in advanced clinical stage. The nodular subtype

S. M. Swetter et al.

made up the dominant proportion of fatal CM cases (55.3% in women, 64.6% in men). The authors concluded that efforts to improve awareness and secondary prevention of CM, including warning signs of all melanoma subtypes, are urgently required and should be targeted toward men in particular. Australian researchers examined recent trends to assess whether earlier evidence of stabilizing melanoma incidence in young people has persisted (Aitkin et al. 2018). Anonymized incidence and mortality data for in situ and invasive melanoma from 1995 through 2014 were obtained from the Queensland Cancer Registry. Over the 20-year period, the incidence of melanoma in situ increased in all age groups. Incidence of both thin (1 mm) and thick (>1 mm) invasive melanoma was either stable or decreased in people under 60, while it increased in those aged 60 and above, particularly in men. Melanoma mortality over the period was stable or decreased in all groups except in men aged 60 or over. These findings are evidence of real advances in the prevention and early detection of invasive melanoma in a very high-risk population and make a compelling case for continued public health efforts to reduce the burden of melanoma in susceptible groups.

Primary Prevention Ultraviolet radiation (UVR) exposure from both natural and artificial sources is strongly linked to CM development, especially in light-skinned populations (see also chapter ▶ “Biology of Melanocytes and Primary Melanoma”). The American Cancer Society recently estimated that nearly 95% of all cutaneous melanoma cases and deaths in the United States are attributable to UVR (Islami et al. 2018). Sun damage is cumulative, but at least 23% of lifetime sun exposure occurs in childhood and nearly 50% by age 40 (Godar et al. 2003). The increased risk for the most common melanoma subtypes stems primarily from acute intermittent UVR exposure, though long-term regular UVR exposure contributes to melanomas on the head, neck, and arms that are linked to chronic sun damage. While

Melanoma Prevention and Screening

UVB radiation has been strongly linked to CM development, increased rates of melanoma from the use of indoor tanning beds, which primarily emit UVA radiation, have demonstrated a key role for sun protection across the entire UV spectrum (280–400 nm), emphasizing the need for broad-spectrum sunscreens that contain more effective UV filters (as yet unavailable in the United States), and strict avoidance of sunburn and tanning – via natural or artificial UVR. Melanoma can be prevented by reducing UVR exposure from sunbathing and indoor tanning and increasing the use of sun protection (Guy 2015). In 2014, the US Surgeon General issued a Call to Action to Prevent Skin Cancer, the goals of which were to increase sun protection in outdoor settings, raise public awareness regarding healthy choices about UVR exposure, reduce harms for indoor tanning, and promote research and policies to prevent skin cancer (US Department of Health and Human Services 2014). Specific prevention measures included reducing personal exposure (i.e., time spent outdoors), ambient levels of sun exposure, occupational sources, and indoor tanning use, as well as promoting the use of protection including shade, UVR-protective clothing, and sunscreens. Therapeutic prevention could reduce the effects of acute exposure by increasing protective mechanisms and reducing the biologic impact of such exposure. Since the 1980s, epidemiologic studies assessing individual sun exposure have identified the causal association of CM in light-skinned individuals following intermittent sun exposure (Elwood et al. 1985; Holman et al. 1986; Osterlind et al. 1988), although it is difficult to define exactly when intermittent UVR exposure becomes chronic exposure. Likewise, geographic studies have shown that total ambient exposure in a particular location is strongly associated with melanoma risk. Thus, it is reasonable to promote sun education programs that are designed to reduce intermittent exposure from recreational and vacation activities. These programs will tend to reduce chronic solar exposure as well, and the overall benefits are likely to outweigh any risks. There is evidence that both early UVR-related mechanisms initiate and promote development of CMs from potential precursors such as nevi and

529

that late mechanisms affect the promotion, proliferation, and metastatic potential of melanomas. It has also been suggested that individuals with lower mole counts, who have a low propensity for melanocyte proliferation, may require chronic sun exposure for CM development, whereas those with higher nevus counts, who have a higher propensity for melanocyte proliferation, may require less UVR exposure for melanomagenesis (Whiteman et al. 2003). Therefore, efforts to reduce solar exposure should be across the life span and likely beneficial at all ages, but particularly in childhood and adolescence (Green et al. 2011a).

Reducing Personal Exposure: Shade, Clothing, and Sunscreens Shade and Clothing Personal sun exposure can be modified by the use of shade, which can reduce overall exposure to UVR by up to 75% (Parsons et al. 1998). A study using dosimeters to measure the exposure of Swedish preschool children to UV radiation during playtimes found that children 5–6 years of age in a shaded playground received a 41% lower dose of UV radiation compared to those in a sunny playground (Boldeman et al. 2004). Sun protection programs in high-risk countries such as Australia place considerable emphasis on providing shade in public areas, at beaches, and in schools and occupational settings, and these efforts have increased worldwide (Hill et al. 1992). The wearing of clothing, even light summer-weight clothing and hats, particularly with an adequate brim, also provides good protection (Gies et al. 2006; Rosenthal et al. 1991). Since the 1990s, sun protection programs in Australia and New Zealand have emphasized these measures, along with the use of sunscreens (Emmons and Colditz 1999). Use of Sunscreens Ultraviolet radiation – both solar and from indoor tanning devices – has been classified as a carcinogen to humans and a strong risk factor for melanoma and other skin cancers (IARC 2007; Wehner et al. 2012). Until recently, no studies had

530

conclusively showed a reduction in melanoma risk following sunscreen use (Dennis et al. 2003; Huncharek and Kupelnick 2002; Gorham et al. 2007), though ample evidence regarding reduction in the development of actinic keratosis (AK) and cutaneous squamous cell carcinoma was evident (Green et al. 1999; Darlington et al. 2003; Thompson et al. 1993; van der Pols et al. 2006). In 2011, results from the community-based randomized controlled Nambour Skin Cancer Prevention Trial were published (Green et al. 2011b). This study conducted in Queensland, Australia, assessed the daily use of sunscreen with sun protection factor (SPF) of 16 to the head and arms in 1621 adults aged 25–75 years, one half of whom used the sunscreen, while the other half used sunscreen on an optional basis or were nonusers. After the 4.5-year intervention period and 10 years of follow-up, there were 50% fewer melanomas in intervention group (11 vs. 22 melanomas, p = 0.051), including melanomas at all body sites (not just the head and arms). An even greater reduction was observed for invasive melanomas (3 vs. 11 cases, risk reduction 73%, p = 0.045). Results from this important trial suggested that long-term use of newer, broadspectrum sunscreens in children/adolescents may have even greater benefits. A Norwegian Women and Cancer Study also showed an apparent benefit from sunscreen use in terms of CM risk (Ghiasvand et al. 2016). This prospective population-based cohort study of nearly 144,000 women aged 40–75 years, who were followed up for a mean of 11 years, showed a reduction in the risk of melanoma reduced by nearly 30% for women using SPF 15 sunscreen on at least one occasion compared with nonusers or those consistently using SPF 50 (Jansen et al. 2013). However, most people apply too little sunscreen (Holman et al. 2015) resulting in an SPF ranging from 50% to 80% less than what is specified on the product label, making the use of sunscreens with SPF of at least 30 reasonable (Diffey 2000; Stokes and Diffey 1997). While the relation of UVB radiation to sunburn and development of AKs and skin cancer is well established (Thompson et al. 1993), the impact of UVA exposure and its carcinogenic effects was only more recently recognized, in part due to increased melanoma incidence resulting from indoor tanning (IARC 2007). As noted, superior, broad-spectrum UV filters have been incorporated into sunscreens in many countries but have not been FDA-approved in the United States, despite efforts to do so since 2002. It is important that sunscreen manufacturers continue to strive to produce sunscreens that provide the broadest possible protection across the entire UVA and UVB spectrum and that these agents be readily available for consumers worldwide.

Behavioral Change Programs for Reducing Personal Exposure Specific programs aimed at modifying personal exposure to UVR from the sun range from small-scale programs intended for specific groups, such as school-aged children (Buller and Borland 1999), to population-based programs with the objective of changing the average exposure levels in the whole population (Montague et al. 2001). Changing sun exposure is a complex behavioral challenge due to influences related to fashion, social and societal norms, and peer culture that may promote sun exposure and sun tanning. In general, the results of most behavior modification programs, particularly those subjected to rigorous evaluation through a randomized trial design or a well-controlled before and after study, have been disappointing in terms of measured effects on behavior, although most

532

programs are able to show some improvement in knowledge and attitudes. A prior review conducted in the United States by the Task Force on Community Preventive Services, as part of the development of the Guide to Community Services, found that education and policy interventions in primary schools and in recreational or tourism settings were effective in increasing sun protection behaviors (Saraiya et al. 2004). Interventions in primary schools generally include a sun safety curriculum, together with activities for specific grades. Among those programs that have been shown to have positive behavioral outcomes are the Kidskin program, conducted in Western Australia over a 2-year period with children in grades 1 and 2 (Milne et al. 2000); the “Sunny Days, Healthy Ways” program in Colorado, New Mexico, and Arizona, with children in grades 2–8 (Buller et al. 1999, 2006a, b); and a program tested in five French primary schools with students in grade 4 (Bastuji-Garin et al. 1999). Policies in schools should support skin cancer prevention by allowing the use of hats and protective clothing and sunscreen application in schools, as well as providing shade structures on school grounds. Interventions in recreational or tourism settings are more varied. One successful program involving young children and their parents at recreation sites in Hawaii included staff training, on-site activities, take-home booklets, and behavior-monitoring boards and incentives, with a second intervention group also receiving sunscreen (Glanz et al. 2000). Another successful program involving beachgoers in Rhode Island used pamphlets, an assessment of sun sensitivity and feedback, sunscreen, instant sun damage imaging photographs, and feedback reports matched to individual stages of change (Weinstock et al. 2002).

Multicomponent Community-Wide Interventions A more recent systematic review of the data through May 2011 by the Community Preventive Services Task Force (CPSTF) concluded that

S. M. Swetter et al.

multicomponent community-wide interventions should be recommended to prevent skin cancer by increasing UV-protective behaviors, based on the current sufficient evidence of their effectiveness in increasing sunscreen use (CPSTF 2012). Studies were eligible for review if they included at least two distinct components that were either implemented in different types of settings (e.g., schools, recreation areas) or directed at an entire community (e.g., mass media campaigns). Some evidence also noted benefits in reducing sunburns, with results for effects on other sun protective behaviors being mixed. Seven studies (Dietrich et al. 2000; Dobbinson 2008; Miller et al. 1999; New South Wales Cancer Council 1998; Olson et al. 2007; Office of National Statistics Report 2010; Rassaby et al. 1983) that evaluated intervention effects on a variety of UV-protective behaviors showed an increase in sunscreen use attributable to the intervention, with increase in sunscreen use by 10.8% in six studies. For the other protective behaviors assessed (i.e., use of shade, hats, and other protective clothing), the results were mixed, with several small or negative effect estimates. Three studies (Dobbinson 2008; Miller et al. 1999; Office of National Statistics Report 2010) showed small positive effects of efforts to limit exposure to UV radiation by decreasing sunbathing or use of indoor tanning beds or reducing time spent in the sun during peak hours. Two of these studies (Dobbinson 2008; Miller et al. 1999) also indicated a decrease in sunburns, with one study (Miller et al. 1999) showing significant decrease in sunburn incidence from 18.6% to 3.2% among children under 6 years of age (a decrease of 15.4 percentage points; 95% CI: 21.2, 9.6). The multicomponent community-wide interventions analyzed and recommended by the CPSTF to prevent skin cancer included combinations of individual-directed strategies, mass media campaigns, and environmental and policy changes across multiple settings within a defined geographic area (city, state, province, or country), to provide an integrated effort to influence UV-protective behaviors (CPSTF 2012). Multicomponent, community-wide interventions coupled with individual-focused strategies

Melanoma Prevention and Screening

(e.g., school-based interventions, interventions targeting outdoor occupational settings) can help to reduce modifiable cancer risk factors, specifically as it relates to UVR. The University of Texas MD Anderson Cancer Center (MD Anderson)’s Be Well Communities™ is an example of a community-driven, place-based approach for cancer prevention and control that promotes wellness and stops cancer before it starts (MD Anderson News Release 2017). Be Well Communities targets its UVR efforts to schools, colleges/universities, local parks and recreation, outdoor workers, and the physical environment to reduce overexposure to the sun and increase availability of sunscreen and sun safety education for the community (Fig. 2). Fig. 2 Be Well Communities™ is the University of Texas MD Anderson Cancer Center’s place-based approach to cancer prevention and control. Be Well™ Baytown is the inaugural Be Well Community. Be Well Baytown is an initiative of MD Anderson sponsored by ExxonMobil

533

Youth Education and Counseling Programs In the United States, approximately 55 million students will attend public and private elementary and secondary schools (National Center for Education on Statistics 2018). Because UVR overexposure increases the risk of melanoma, it is important to implement skin cancer prevention initiatives early, making schools an ideal setting for such efforts. Recognizing the potential impact that early childhood education could have as a component of a larger strategy, MD Anderson developed a sun safety program, Ray and the Sunbeatables ® (https://sunbeatables.org/), for preschool, kindergarten, and first-grade students (Fig. 3).

534

S. M. Swetter et al.

Evidence-based sun safety education curricula

Available at: Sunbeatables.org

Available at: Scholastic.com/sunsafety

Fig. 3 Examples of freely available evidence-based sun safety education curricula spanning preschool through fifth grade

Prioritization and activation of investments in youth education strategies are novel for an academic medical and research institution. Significant funding has been invested in the research and validation of Ray and the Sunbeatables ®. To extend the impact and return on program investment in the Sunbeatables Program, MD Anderson recently embarked on a collaboration with Scholastic Inc. to develop Be Sunbeatable™ (http:// www.scholastic.com/sunsafety/), a sun safety program for second- to fifth-grade students. Scholastic is a nationally recognized name in the education arena with extensive years of expertise devoted to increasing knowledge for school children (http://www.scholastic.com/home/). This collaboration has provided MD Anderson with the opportunity to extend evidence-based sun safety programming in a scalable model to schools around the country targeting students, educators, parents, and school nurses.

US Preventive Services Task Force (USPSTF): Behavioral Counseling Since 2012, the US Preventive Services Task Force (USPSTF) has recommended that lightskinned children, adolescents, and young adults aged 10–24 years minimize their exposure to ultraviolet radiation to reduce risk for skin cancer

(Moyer and USPSTF 2012). A “B recommendation” was issued, indicating that behavioral counseling is of at least moderate benefit in increasing sun protection behaviors in this age group. An updated recommendation was issued by the USPTF in 2018 that extended skin cancer prevention behavioral counseling to begin at age 6 months (USPSTF et al. 2018) and recommended selective counseling to adults older than 24 years with light skin.

The SunSmart Program The best-known and well-studied populationbased melanoma prevention program is SunSmart, which was initiated in Victoria, Australia, in 1988 (Montague et al. 2001) and has served as a showcase for the positive results that can be achieved by a sustained effort, including recent reductions in the population-based incidence of melanoma in younger age groups (Aitkin et al. 2018). The SunSmart program grew out of a limited public education campaign, launched in 1980, that featured the slogan “Slip! (on a shirt) Slop! (on sunscreen) Slap! (on a hat),” in an effort to reduce individual exposure to UV radiation (Montague et al. 2001). With funding from the Victorian Health Promotion Foundation (VicHealth), in 1988, the SunSmart program was

Melanoma Prevention and Screening

launched as a broad-based, multifaceted skin cancer control program. The SunSmart program aimed to promote sun protection behaviors by influencing public knowledge, attitudes, and intentions, involving social and cultural norms and the environmental and legislative context of these domains. For example, early media campaigns not only aimed to improve knowledge of the link between sun exposure and melanoma but also began the process of changing public opinion regarding the desirability of a tan. Together with sustained advocacy from the SunSmart program, this created community demand for policies and infrastructure in a variety of settings to support sun protection behaviors. The SunSmart program activities were implemented in many settings simultaneously, with the intent of changing the social context, disseminating knowledge to specific groups and the population more broadly, and working toward legislative and environmental changes. Paid media campaigns evolved from the positively framed Slip! Slop! Slap! to harder-hitting, more graphic reminders of the effects of skin cancer that were designed to influence sections of the population that have proved more resistant to the initial messages. Work with specialized groups and organizations, including local governments and school councils, resulted in an increase in community demand for shade structures and improvements in shade availability in a number of settings. As a result, the use of shade structures in outdoor swimming pools increased as did the use of sun protective clothing use (by 3% per year) from 1992 to 2002 (Dixon et al. 2008). One of the most successful components of the SunSmart program has been in schools. Schools signing up for the program must have a policy that meets the criteria established by the SunSmart program, which cover behavior, curriculum, and the environment. By 2008, 84% of the primary schools in Victoria were members of the program; this contributed to sun protection behaviors becoming the norm rather than the exception, particularly among younger children. Likewise, self-reported rates of sunburn were significantly reduced compared to the baseline incidence of sunburn before

535

the inception of the SunSmart program, with 9% of adults being sunburned on an average summer weekend in 2001–2002 compared with 14% in 1987–1988 (Dobbinson et al. 2008). Most importantly in terms of primary prevention, the sustained nature of the SunSmart program over many decades appears to have resulted in changes in melanoma incidence in adolescents and young adults. In a study comparing younger age (0–24 years) white populations in England from 1990 to 2010 (Wallingford et al. 2015), Australian youth demonstrated twice the overall melanoma incidence compared to English youths (2.2 and 1.1 per 100,000, respectively), with similar rarity of melanoma among children 70% changed in surface area by at least 50%, 59% increased, 3% decreased, and 10% disappeared (Wu et al. 2014). Compared to reticular nevi, globular nevi were significantly more likely to increase in size (65% vs. 52%) and less likely to decrease (7% vs. 18%) ( p = 0.02). This growth characteristic difference by dermoscopic pattern was most pronounced for nevi with the greatest surface area at baseline ( p = 0.001) (Wu et al. 2014). Preliminary analysis of 945 nevi from the back and legs followed from age 14 to 17 (n = 213 participants) showed that 22 nevi (2.3%) completely disappeared; all but one disappeared without a halo or regression (unpublished data). In SONIC, it was observed that globular-patterned nevi occur more frequently on the upper part of the body, where they tend to be larger in diameter; on the lower trunk and extremities, the predominant nevus pattern becomes increasingly reticular and nevi tend to be smaller (Fonseca et al. 2015; Scope et al. 2008). At age 11, nevi were more likely to be globular on the upper than on the lower back (OR = 2, 95% CI 1.6–2.6, p < 0.001) (Scope et al. 2008). At age 14, compared to referent homogeneous nevi, globular nevi were more commonly observed on the back than the legs (OR = 29.4, 95% CI 9.5–90.7, p < 0.001), whereas reticular nevi were less likely to be

Melanoma Prevention and Screening

observed on the back than the legs (OR = 0.7, 95% CI 0.5–0.8, p = 0.001) (Fonseca et al. 2014, 2015). Four controlled trials have been based on programs designed to reduce the rate of development of nevi in children by reducing sun exposure. In an individually randomized trial in Vancouver, Canada (Gallagher et al. 2000), an educational campaign incorporating the provision of free sunscreen was assessed in children who began the study in first and fourth grades and completed it 3 years later. Parents in the intervention group were supplied with a broad-spectrum sunscreen containing SPF 30+ and encouraged to apply it to their children when the children were expected to be in the sun for 30 min or more. Parents in the control group were not given any intervention. The number of new nevi acquired over the succeeding 3 years was assessed by clinical examinations before and after the intervention. The children in the intervention group had fewer nevi, although the overall effect was a modest 9% reduction in median count, which was marginally significant. The effect was greater in children with more freckles, who exhibited an estimated 30–40% reduction in new nevi. Further analysis showed a stronger effect for intermittently sun-exposed body sites (Lee et al. 2005). In Perth, Western Australia, a controlled but nonrandomized study tested the effect of an educational intervention incorporating advice and monitoring of sun protection activities at school (Milne et al. 2002). The study showed consistent improvements in sun protection, as reported by parents, and also showed that children in the intervention group had lighter skin color at the end of the summer. Although there was only a small and marginally significant effect on nevus counts, this was maintained for 2 years after the intervention ended (English et al. 2005). Two other randomized controlled trials failed to find significant differences in the number of nevi between intervention and control groups. In Colorado, health-care providers dispensed advice and materials to parents of infants between 2 and 36 months of age (Crane et al. 2006). Although differences in the parents’ sun protection behaviors were observed and appeared to increase

537

over the 3 years of follow-up, skin examinations revealed no significant differences in the number of nevi. In Germany, a trial investigated the influence of sunscreen use and education on the incidence of nevi in preschool children (Bauer et al. 2005b). However, there were no reported differences in sun protection habits attributable to the intervention efforts after 3 years of follow-up and hence no significant differences in the number of incident melanocytic nevi. The modest results from these studies suggest that the degree of change in sun exposure that can be achieved by such programs is insufficient to produce a large effect on nevi in a short time, except perhaps in higher risk groups, even if sun exposure is the predominant cause. As yet, the quantitative relationship between sun exposure and the incidence of nevi is unclear, making it difficult to predict how much change in UVR exposure is needed to affect the incidence of nevi; this limitation applies to melanoma as well.

Controlling Exposure to Indoor Tanning Beds Apart from natural sun exposure, the major source of exposure to UVR for the general population is the use of indoor tanning beds and, previously, sunlamps. Gaining popularity in the 1960s for producing a suntan, these devices emit predominantly UVA but also small amounts of UVB (usually less than 2%). A systematic review by the International Agency for Research on Cancer (IARC) found a 75% increased risk of melanoma among those who had ever used a sunbed before the age of 35 (IARC 2007). More recent data have confirmed the strong association of tanning bed use and melanoma development (and other skin cancers) (Boniol et al. 2012; Colantonio et al. 2014; Wehner et al. 2014), with epidemiologic data supporting an increased incidence in young women as a result (Purdue et al. 2008). The risk of melanoma is also higher among those who initiate indoor tanning at a young age and those who frequently indoor tan (Boniol et al. 2012; Lazovich et al. 2016).

538

Since 2011, 19 US states and the District of Columbia have enacted legislation to restrict tanning beds by minors under 18 years of age, but there is more work to be done to match tanning restrictions in other countries, including Australia and Brazil, which have banned indoor tanning beds outright (Pawlak et al. 2012). In December 2015, the US Food and Drug Administration proposed a federal rule restricting minors’ access under 18 years of age to tanning beds and requiring that sunlamp manufacturers and tanning facilities take additional measures to improve the overall safety of such devices (US FDA 2015), although this is currently on hold (US Office of Management and Budget 2017). For the first time in the United States, there appears to be a true decline in the use of indoor tanning, with rates now below 15% for teenage white girls, a drop of more than 50% in less than 5 years. As a whole, indoor tanning among US high school students has decreased by 53% between 2009 and 2015 (Guy et al. 2017), and this is likely attributed to a greater number of states containing age-specific bans on indoor tanning. A 2017 Youth Risk Behavior Surveillance Report from the Centers for Disease Control (CDC) confirms these trends, showing that 5.6% of students nationwide had used an indoor tanning device one or more times during the 12 months before the survey, with trend analyses indicating a significant linear decrease in indoor tanning during 2009–2017 (Kann et al. 2018). However, the prevalence of having had a sunburn in the 12 months pre-survey did not change significantly from 2015 (55.8%) to 2017 (57.2%). A recent publication indicates that the reach of tanning bed legislation has now affected much of the United States with the percentage of high school girls living in states affected by age restrictions precipitously rising from only 2% living in states with age restrictions in 2009 to 57% living in states with age restrictions just 6 years later (and likely higher today) (Qin et al. 2018). Furthermore, the drop in indoor tanning was observed among former indoor tanners, adolescent girls 14–17 years of age, and

S. M. Swetter et al.

adolescents of different races (white, black, and Hispanic), across all calendar years but most notably in 2015, the year that the data were last collected. In all, adolescent girls affected by age restrictions had a 47% decline in tanning bed use compared with their counterparts in states with only parental permission laws or no laws at all. The fact that only 7% of adolescent girls living in states with age restrictions report past-year use of indoor tanning beds has to be considered a huge victory for public health advocates. Although new age restrictions are not the only explanation for less tanning bed use, the decline seems to have heightened with the passage of the first tanning bed ban for those aged younger than 18 years in California in 2012. Strikingly, within 5 years, 18 other states and the District of Columbia imposed bans for those younger than 18 years; adolescent bans are now present in every geographic region of the United States, including Texas, North Carolina, Louisiana, Kansas, and Oklahoma. Such legislation left the tanning bed industry with far fewer customers, with projected reduction in tanning beds to 9000 stand-alone units in 2015 (from indoor tanning documents) compared with more than 19,000 units a decade prior. Documentation of the results from welldesigned and rigorous studies (Cust et al. 2011; Ghiasvand et al. 2017; Lazovich et al. 2008; Mayer et al. 2011) became part of the public record and testimony that profoundly influenced legislators. In addition, advocates, including the American Cancer Society Cancer Action Network, AIM at Melanoma, American Academy of Dermatology, American Academy of Pediatrics, and the National Council on Skin Cancer Prevention, and informal networks of family members of young women who lost their lives to melanoma were armed with information on the disproportionate toll of tanning bed-related melanoma in young women. The challenge is to ensure that any legislation that is put in place truly enhances public health outcomes and does not in any way legitimize an industry that is contributing to skin cancer morbidity and mortality.

Melanoma Prevention and Screening

Therapeutic Prevention of Melanoma and Populations to Target for Interventional Trials Therapeutic Prevention of Melanoma Therapeutic prevention or precision prevention (formerly referred to as chemoprevention) is a growing field in which synthetic drugs or natural agents are given to an individual at high risk for a particular disease with the objective of reducing the risk of that disease. Ideal therapeutic prevention agents for melanoma must not only reduce melanoma risk but also be well-tolerated, nontoxic, and cost-effective. The population(s) that will benefit from a therapeutic prevention agent should also be well-defined in order to maximize the risk-benefit and cost-benefit ratio of the treatment. Jeter et al. have reviewed published evidence on therapeutic melanoma prevention agents, with an emphasis on evaluating each agent as a potential candidate for a larger phase III trial in high-risk patients (Jeter et al. 2018, in press). This manuscript, reflecting the opinion of the National Clinical Trials Network (NCTN) Melanoma Prevention Working Group, has proposed a strategy to move the field toward therapeutic prevention clinical trials that includes a rigorous evaluation of each agent, using validated biomarkers at each stage of melanoma development (i.e., initiation, promotion, and progression) in (1) the in vitro setting, (2) relevant melanoma mouse models, and (3) human model systems (e.g., ex vivo skin and early-phase clinical trials). In addition to identifying candidate agents with an appropriate safety profile, it is critical to design trials that involve a patient population of sufficiently high risk to adequately power the study and include clinical and molecular endpoints of efficacy. High-risk groups such as those with a hereditary predisposition, a personal history, or a combination of multiple risk factors such as red hair color, multiple large, atypical nevi, or extensive UVR exposure are examples of populations that might be considered. Ultimately, any clinical trial of a therapeutic prevention agent will need to be tailored to the specific agent selected in order to balance the potential toxicity with the

539

population’s risk of developing melanoma. To date, a variety of different classes of therapeutic prevention agents have been proposed, including sunscreens, pigmentation enhancers, DNA repair enzymes, vitamins and minerals, repurposed therapeutic agents, and phytochemicals.

Candidate Agents Pigmentation Enhancers Increasing the production of endogenous melanin pigment is another means of reducing UVRinduced mutations in the skin. Melanin, a photoprotective pigment produced by melanocytes, is packaged into melanosomes and transported to keratinocytes, where the pigment literally forms an umbrella over the nucleus to reduce DNA damage. The production of melanin is controlled by α-melanocyte-stimulating hormone (MSH), which binds to and activates the melanocortin 1 receptor (MC1R) (Swope and Abdel-Malek 2016). MC1R signals downstream pathways that upregulate expression of the microphthalmiaassociated transcription factor (MITF), which in turn upregulates pigmentation gene expression (Kawakami and Fisher 2017). Several agents have been developed to enhance pigmentation through this pathway, including α-MSH agonists (Abdel-Malek et al. 2014) and salt-inducible kinase (SIK) inhibitors (Horike et al. 2010), which both lead to upregulation of MITF and subsequent increased pigmentation production. Treatment with α-MSH and its analogs results in reliable tanning of human skin, but also produces side effects including increased sexual libido, nausea, flushing, and loss of appetite and has been sold illicitly with contaminants (Barnetson et al. 2006; Breindahl et al. 2015; Hadley 2005). A synthetic tridecapeptide form of α-MSH (AKA NDP-MSH, afamelanotide, and marketed as Scenesse© by Clinuvel Pharmaceuticals) is approved in Europe for treatment of erythropoietic porphyria (Langendonk et al. 2015), but does not have a topical formulation available and has not been tested as a therapeutic prevention agent for melanoma. Of some concern is the report of eruptive nevi following the use of unlicensed

540

MSH analogs (melanotan I and II) (Burian and Burian 2013; Cardones and Grichnik 2009; Cousen et al. 2009; Ferrándiz-Pulido et al. 2011; Schulze et al. 2014), which could suggest that stimulation of pigmentation by MSH could inadvertently predispose to melanocyte proliferation (and possible malignancy) and outweigh the protective effect of the tan they produce. The SIK inhibitors have a topical formulation available and have been shown to increase pigmentation in mouse and human skin explants, but have not yet been tested in humans (Mujahid et al. 2017).

DNA Repair Enzymes A therapeutic prevention alternative to protection against UVR-induced DNA damage (i.e., via sunscreens or pigmentation enhancers) is to increase the repair of any DNA damage that may occur before it results in a pathogenic mutation. DNA glycosylases are able to detect and enzymatically remove UVR-induced mutations (Yarosh et al. 2001; Johnson et al. 2011), and two different topical formulations of this enzyme have been shown to prevent squamous cell carcinoma in mice. A liposomal formulation of one of these products, a bacterial T4 endonuclease, has also been shown to reduce the number of AKs and basal cell carcinomas in xeroderma pigmentosum patients (Yarosh et al. 2001), and another study using a topical DNA repair enzyme led to reduced AKs on the face and scalp relative to a placebo control (Stoddard et al. 2017). There have not been any clinical trials to test DNA repair enzymes in non-XP melanoma patients, and no FDAapproved agents are on the market. Vitamins and Minerals Vitamins and minerals represent another class of potential therapeutic prevention agents that have a relatively long track record of safety and over-the-counter use in humans, making them attractive as preventives. Vitamin A (retinoids, retinol, carotenoids, and beta-carotene), vitamin D, vitamin E, nicotinamide, and selenium have all been investigated for their potential to serve as a melanoma prevention agent (Jeter et al. 2018).

S. M. Swetter et al.

Vitamin A: Despite investigating the role of vitamin A and its precursors (retinol and the carotenoid provitamins for vitamin A including beta-carotene) in multiple preclinical studies (Mounessa et al. 2016), case-control studies (Feskanich et al. 2003; Millen et al. 2004; Naldi et al. 2004), cohort studies (Asgari et al. 2009b, 2012), a Nurses’ Health Study (DruesnePecollo et al. 2010), a Women’s Health Study (Lee et al. 1999), and the SU.VI.MAX trial (Hercberg et al. 2007), results have been mixed and not supportive of advancement into a phase III therapeutic prevention trial. Oral isotretinoin (13-cis-retinoic acid) has also been investigated in patients with dysplastic nevi, but no clinical or histologic benefit was evident (Edwards and Jaffe 1990). Vitamin D: Although exposure to UVR is the key modifiable risk factor for melanoma, it also has beneficial effects. Principal among these is the production of vitamin D. Vitamin D has long been recognized as playing a key role in maintaining musculoskeletal health and bone density (Papadimitropoulos et al. 2002; Trivedi et al. 2003) and has been linked to possible beneficial treatment or prevention effects for a number of diseases, including breast, prostate, and colorectal cancer, non-Hodgkin’s lymphoma, diabetes, and multiple sclerosis (Giovannucci et al. 2006; Hughes et al. 2004; Mathieu and Badenhoop 2005; Zittermann 2003). Appropriate levels of sun exposure to minimize melanoma risk while maintaining adequate levels of vitamin D vary according to skin type and intensity of UVR, which is dependent on the time of day, distance from the equator, season, altitude, and cloud cover. For locations close to the equator, and for more temperate areas in the summer, a few minutes outside of peak UV times should be sufficient for light-skinned people to achieve adequate vitamin D levels (more than 50 nmol/L) by exposing the face, arms, and hands or the equivalent surface area to sunlight (Samanek et al. 2006). Some studies have shown that up to 80% of individuals with dark skin, low levels of UVR exposure, or babies of individuals with low vitamin D levels display evidence of vitamin D deficiency (Clemens et al. 1982; Grover and Morley

Melanoma Prevention and Screening

2001; Lips 2001; Nozza and Rodda 2001; Riggs 2003; Thomson et al. 2004). Managing sun exposure to achieve optimal health benefits is therefore a balancing act between too much and too little (Kimlin and Tenkate 2007). Populations at risk for vitamin D deficiency or at risk of melanoma may not be able to achieve adequate vitamin D levels through sun exposure, because too much exposure is required or because exposure is not safe. In these cases, oral supplementation may be advisable. Guidelines on how to achieve a reasonable balance of risk and benefit with respect to vitamin D and sun exposure have been developed by organizations concerned with skin cancer as well as bone and mineral metabolism (IOM 2010; NIH 2018). The significance of vitamin D as a therapeutic prevention agent is unclear, despite several published case-control studies of vitamin D intake and melanoma incidence and studies of vitamin D polymorphisms and melanoma susceptibility (Asgari et al. 2009a; Caini et al. 2014; Millen et al. 2004; Newton-Bishop et al. 2015; Orlow et al. 2016; Saiag et al. 2015; Skaaby et al. 2014). A large, prospective cohort study (n = 12,000) failed to reveal a statistically significant association between vitamin D and melanoma incidence, risk, prognosis, or predisposition. In addition, neither a large randomized interventional Women’s Health Initiative (WHI) trial of 36,828 postmenopausal women taking vitamin D and calcium supplementation nor a meta-analysis of pooled investigations revealed an impact of vitamin D on melanoma risk or prognosis (Caini et al. 2014; Skaaby et al. 2014; Tang et al. 2011). However, a weak association between melanoma incidence and vitamin D intake was found in another meta-analysis (Gandini et al. 2009). A single study evaluating the role of highdose vitamin D supplementation on cutaneous inflammation found that high serum vitamin D levels were associated with decreased inflammation and redness in skin exposed to one to three times the minimal erythema dose of solar simulated radiation (Scott et al. 2017). A recent nested casecontrol study of the WHI Observational Study showed an association between higher serum vitamin D levels and greater risk for melanoma development, while lower levels were associated with

541

increased risk of melanoma-related death (Kwon et al. 2018). These studies highlight the challenges of investigating the association of vitamin D levels with melanoma risk or outcome, since levels are directly affected by sunlight (UVB) exposure, a major risk factor for melanoma development in itself. Although adequate vitamin D levels are essential for health and may be important in UV-induced inflammation, strong evidence supporting it as a therapeutic prevention agent is lacking. Vitamin E: Preclinical models suggest that vitamin E and its analogs might be useful for preventing melanoma (Chhabra et al. 2017), presumably because of the strong antioxidant properties of vitamin E, yet a vitamin E analog (Trolox) has been shown to increase migration and invasion of human melanoma cells through glutathione-related pathways (Le Gal et al. 2015). Many studies, including case-control studies and randomized trials (Ezzedine et al. 2010; Feskanich et al. 2003; Hercberg et al. 2007; Mahabir et al. 2004), have been published on vitamin E effects on melanoma incidence, but have mixed results, and prospective data from the Nurses’ Health Study showed that total and dietary vitamin E were not associated with risk for melanoma (multivariate RRs 1.11 (0.66–1.85) and 0.88 (0.59–1.32), respectively) (Feskanich et al. 2003). However, it does appear that topical and oral vitamin E can protect the skin from UVR (Placzek et al. 2005). Conversely, the SU.VI. MAX trial (also discussed above) found an increased risk for women consuming antioxidant supplements, and the Selenium and Vitamin E Cancer Prevention Trial (SELECT) also found an increased risk for prostate cancer ([HR], 1.17; 99% CI, 1.004–1.36, P = 0.008) in men consuming oral vitamin E supplements (400 IU daily as racemic alpha-tocopheryl acetate) (Klein et al. 2011). Together, these data suggest that vitamin E may not be a promising candidate melanoma preventive agent because of these potentially detrimental side effects. Nicotinamide: In vitro studies of nicotinamide and nicotinic acid (vitamin B3 group) in melanocyte and melanoma cell culture systems are apparently contradictory, suggesting both a growth

542

inhibitory effect as well as stimulation of invasiveness (Minocha et al. 2017). Oral nicotinamide (1500 mg or 500 mg daily for 3 days) decreases UV-induced immune suppression in human skin irradiated in vivo, and use of 500 mg oral nicotinamide daily for 12 months in a double-blind, randomized phase III clinical trial resulted in a 13% reduction in AKs ( p = 0.001) and 23% reduction in nonmelanoma skin cancers (NMSCs) ( p = 0.02) (Chen et al. 2015). However, the study was not adequately powered to assess an effect in melanoma and was conducted over a relatively short period of time, making it difficult to conclude whether nicotinamide might be a good candidate for melanoma prevention, though it is an area worth further investigation given the shared UVR-related etiology of these tumors. Selenium: Selenium is a required trace element for the production of antioxidant proteins, and deficiency can lead to impaired muscular, cardiac, and immune functions, as well as elevated cancer risk (Roman et al. 2014). Selenic acid has a greater pro-apoptotic effect on human melanoma cells than primary human melanocytes (Cassidy et al. 2013), and topical treatment with L-selenomethionine results in a significant delay in the time required for UV-induced tumor development in NMSC (Burke et al. 2003), but continued treatment increases the rate of growth of melanomas once tumors appear (Cassidy et al. 2013). Case-control and cohort studies have evaluated the role of selenium in melanoma, with varying results. Although one cohort of earlystage melanoma patients was found to have low selenium levels associated with worse outcomes at 2 years (Deffuant et al. 1994), an Italian cohort (Vinceti et al. 1998), the Nurses’ Health Study (Garland et al. 1995), and the VITAL cohort (Asgari et al. 2009b) all either failed to show an apparent benefit to selenium or an increased risk of melanoma with elevated selenium levels. The SU.VI.MAX (summarized in the vitamin A section) (Hercberg et al. 2007) and the Nutritional Prevention of Cancer Trial for NMSC (Vinceti et al. 1998) also failed to find a compelling benefit of selenium for melanoma. Furthermore, SELECT (summarized in the vitamin E section) (Klein et al. 2011) was stopped early due to concerns about

S. M. Swetter et al.

the increased risk of prostate cancer in men receiving selenium and vitamin E supplementation. Overall, the use of vitamins and minerals as therapeutic prevention agents for melanoma is still in its infancy, but early data suggests the need for caution due to the observation that many of these agents appear to increase rather than decrease risk for melanoma, despite a pervasive view that these agents are generally safe.

Repurposed Therapeutic Agents Another method of identifying a possible therapeutic prevention agent is by testing agents that have a well-characterized and acceptable toxicity as demonstrated through a long-standing track record of use for another indication. Several such repurposed agents have been tested in melanoma, including aspirin and nonsteroidal anti-inflammatory drugs (NSAIDs), statins, Nacetylcysteine, and difluoromethylornithine (DFMO). Aspirin and Nonsteroidal Antiinflammatory Drugs (NSAIDs) Several mechanisms of action have been proposed by which aspirin or NSAIDs might be effective therapeutic prevention agents. Inhibition of COX-2 decreases levels of prostaglandin E2 (PGE2) and reduces IL-6 production, an immunomodulatory cytokine that plays a role in the development and progression of melanoma (Kast 2007). Alternatively, in the case of aspirin, quinone metabolites have the potential to deplete GSH, leading to increased oxidative damage and consequent DNA damage. Numerous studies have been published assessing the use aspirin, diclofenac, and sulindac for melanoma risk or prevention (Albano et al. 2013; Brasky et al. 2014, Curiel-Lewandrowski et al. 2011; Famenini and Young 2014; Gamba et al. 2013; Goodman and Grossman 2014; Hu et al. 2014; Jacobs et al. 2007; Jeter et al. 2012; Johannesdottir et al. 2012; Joosse et al. 2009; Li et al. 2013; Zhu et al. 2015). Of these studies, the most reliable indicators of a possible effect are found in large cohort studies and a few clinical trials. In the Cancer Prevention Study II Nutrition Cohort (n = 146,113) (Jacobs et al. 2007) and the Vitamins and Lifestyle cohort

Melanoma Prevention and Screening

(VITAL, n = 77,719) (Asgari et al. 2008), there was no association between aspirin or NSAID use and melanoma risk. The initial analysis of data from the Nurses’ Health Study showed an increased risk for melanoma in aspirin users (Jeter et al. 2012), although subsequent analyses from the WHI found a decreased risk, which was site specific for melanoma and colorectal and ovarian cancer but not related to total cancer risk (Brasky et al. 2014; Gamba et al. 2013). Two meta-analyses of NSAIDs (Hu et al. 2014; Li et al. 2013) and one for aspirin (Zhu et al. 2015) also failed to show dramatic effects of these agents on melanoma risk. There have been no prospective clinical trials to directly test the efficacy of aspirin or other NSAIDs on melanoma incidence, but a randomized, double-blind, placebo-controlled trial of sulindac was performed in individuals at risk for melanoma that suggested treatment might reduce the threshold for apoptosis in benign nevi (CurielLewandrowski et al. 2012).

Statins It has been proposed that statins might exert a therapeutic preventive effect through HMG-CoA reductase-mediated decreases in farnesyl and geranylgeranyl diphosphate, which is critical for activation of Ras and Rho families, Rac, Rab, Cdc42, and nuclear lamins (Palsuledesai and Distefano 2015), or via caspase-dependent apoptosis via inhibition of protein geranylgeranylation and the induction of cell cycle arrest (Saito et al. 2008; Shellman et al. 2005). Numerous case-control studies and prospective cohort analyses of the effects of statins on melanoma incidence have been performed (Bonovas et al. 2010; Browning and Martin 2007; Dellavalle et al. 2005; Freeman et al. 2006; Jacobs et al. 2011; Jagtap et al. 2012; Kuoppala et al. 2008; Linden et al. 2014). A promising analysis of the Cancer Prevention Study II Nutrition Cohort (n = 133,000) revealed that 5 or more years of statin use was associated with a lower risk of melanoma (RR 0.79, 95% CI 0.66–0.96) (Jacobs et al. 2011). However, the Women’s Health Initiative (Jagtap et al. 2012) and meta-analyses showed no association of statins with melanoma incidence (Bonovas et al. 2010; Browning and

543

Martin 2007; Dellavalle et al. 2005; Freeman et al. 2006; Li et al. 2014), and one study showed an increased melanoma risk in association with statin use (Kuoppala et al. 2008). Subgroup analysis of one of the earlier meta-analyses indicated that lovastatin might have a drug-specific effect, but no further data were found to confirm this result on subsequent meta-analyses. The most direct test of a statin as a therapeutic preventive agent in melanoma was tested in a randomized, prospective, double-blind, placebo-controlled trial using nevi as a surrogate marker of effect (Linden et al. 2014). Unfortunately, this trial was negative, failing to reveal any effect of lovastatin on nevi or markers of transformation. Taken together, the case for statins as a therapeutic prevention class of agents for melanoma is not compelling.

N-Acetylcysteine N-Acetylcysteine (NAC), a potent antioxidant, is used to treat acetaminophen toxicity and prevent contrast-induced nephropathy and lung disorders (Maxwell 1995; Pei et al. 2018). In mice, NAC delays primary tumor development of UVinduced melanoma (Cotter et al. 2007), and a phase I study suggested a possible preventive effect of NAC on nevi derived from at-risk individuals exposed to UVR ex vivo (Goodson et al. 2009). However, this was not reproduced in a placebo-controlled phase II clinical trial (Cassidy et al. 2017). In addition, mice that receive chronic administration of NAC developed more metastases without an increase in the number of tumors, suggesting that although NAC may prevent initiation of melanoma, it may have deleterious effects on progression of already existing tumors (Le Gal et al. 2015). Difluoromethylornithine Difluoromethylornithine (DFMO) inhibits the synthesis of ornithine decarboxylase and has been shown to prevent sporadic colon adenomas in combination with the NSAID sulindac (Meyskens et al. 2008). Although there have been no clinical or epidemiologic studies of the effects of DMFO on melanoma in humans, in combination with interferon gamma, DFMO treatment causes melanoma cell growth arrest in

544

culture (Bregman et al. 1987). One clinical study is in progress to evaluate topical diclofenac  topical DFMO in the prevention of keratinocyte skin cancers and precancers (see http:// clinicaltrials.gov/).

Phytochemicals (Plant-Derived Biologically Active Compounds) Several phytochemicals have been identified as candidates for therapeutic prevention of melanoma, largely because they reduce sensitivity to UVR, target pathways that appear to be involved in melanoma development, or both. The best characterized and tested of these compounds include epigallocatechin-3-gallate (EGCG), resveratrol, sulforaphane (SFN), bixin, Polypodium leucotomos extracts (PLE), and silibinin. Epigallocatechin-3-gallate (EGCG) is a flavonoid (plant pigment) found in green tea and, to a lesser extent, black tea. EGCG has been shown to promote cell cycle arrest and apoptosis and inhibit angiogenesis and has potentially preventive roles in inflammation and immunity, as well as antioxidant effects (Chhabra et al. 2017). A bioavailable nanoparticle formulation has been shown to have increased potency in a human melanoma xenograft model (Siddiqui et al. 2014). Tea consumption was not shown to have an effect on melanoma incidence in either of the two cohort studies: the Iowa Women’s Health Study (n = 35,000 postmenopausal women) (Zheng et al. 1996) and the Women’s Health Initiative (n = 66,484 postmenopausal women) (Wu et al. 2015). Although clinical trials for NMSC have been performed with EGCG, no trials for melanoma have yet been conducted. Resveratrol is an antioxidant polyphenol commonly found in berry juices and red wine. Resveratrol has low bioavailability, but nanoparticle formulations and a naturally occurring analog, pterostilbene, are being investigated as a more bioavailable alternative (reviewed in Chhabra et al.). Another antioxidant, sulforaphane (SFN), is an isothiocyanate derived from broccoli, brussels sprouts, and cabbage. It is hypothesized that SFN may act as a preventive by activating Nrf2-mediated expression of multiple antioxidant enzymes, as well as inhibition of the transcription factor AP-1

S. M. Swetter et al.

(Dickinson et al. 2009). In vitro experiments have shown that SFN can reduce the growth of melanoma cells (Fisher et al. 2015), but there are no case-control, cohort, or randomized controlled trials of SFN for treatment of melanoma. A phase I dose-finding melanoma chemoprevention study of SFN has been recently completed at the University of Pittsburgh and has shown excellent tolerability of dosages of broccoli sprout-derived SFN at 50, 100, and 200 μMol daily dosages for 1 month, with dose-dependent levels of SFN reached in the blood and nevi. A larger phase II study of the 200 μMol daily dosage for longer periods of time in patients with a history of melanoma and multiple atypical nevi is being planned in the National Clinical Trials Network (NCTN) of the Eastern Cooperative Oncology Group-American College of Radiology Imaging Network (ECOG-ACRIN) (Tahata et al. 2018). Additional, less studied melanoma prevention agents include bixin, Polypodium leucotomos extract, and silibinin. Bixin is an apocarotenoid with an excellent safety record and good systemic bioavailability when administered orally. Like sulforaphane, it may act through Nrf2-mediated antioxidant effects (Tao et al. 2015). No human clinical trials of bixin for prevention of skin cancers have been published. Standardized extract of the tropical fern Polypodium leucotomos (Polypodium leucotomos, Fernblock ®, PLE) has been reported to have antioxidant and anti-inflammatory properties and has been tested for a variety of dermatologic photosensitivity conditions (Parrado et al. 2016; Kohli et al. 2017). One study evaluated PLE in high-risk melanoma patients and found that PLE treatment significantly increased the minimal erythema dose (MED), but the effect, if any, on melanoma prevention was not directly tested (Aguilera et al. 2013). PLE contains a number of phenolic compounds, but there has not been published characterization of the extract, raising some concern about whether standardization of the extract is sufficient for melanoma clinical trials. Silibinin is derived from milk thistle and has been reported to target MEK- and RSKmediated and/or TP53-mediated signaling pathways (Kumar et al. 2015; Rigby et al. 2017). No

Melanoma Prevention and Screening

trials with melanoma in humans have been performed.

Secondary Prevention Early Detection and Screening The major goal of early detection and screening programs is to reduce deaths from melanoma. Targeting or tailoring behavioral messages and screening strategies hinges on identifying populations at greatest risk of fatal melanoma that have never or rarely engaged in screening. Two demographics are illustrative – white men ages 50 and above who make up nearly 60% of all melanoma deaths and individuals with low socioeconomic status, generally defined as high school education or less and low income. There will also be overlap across these two groups. For a cancer that depends so much on awareness of clinical warning signs for the disease in hard-tosee sites, having a partner or significant other may be key to earlier detection. Notably, a number of studies have found that marriage is modestly related to earlier-stage melanoma.

Potential Benefits of Screening The value of screening for melanoma may seem obvious. In most instances, melanoma is a slowly progressive disease. Cutaneous melanoma is visible on the surface of the skin, and so visual examination by a physician or other health provider (aka “clinical skin examination,” CSE) or by patients themselves (aka “skin self-examination,” SSE) provides a simple, noninvasive screening test that can detect the disease before clinical diagnosis would otherwise take place. In most countries with moderate and high incidences of melanoma, the great majority are diagnosed before there is apparent regional or distant metastatic spread. Melanoma-specific survival after diagnosis varies dramatically according to tumor thickness, ranging from 98% 10-year melanomaspecific survival in node-negative patients with melanomas 1.0 mm thickness to 75% in those

545

with melanomas that are ulcerated and >4 mm thickness (Gershenwald et al. 2017). These facts support that skin screening for melanoma should be beneficial in terms of melanoma mortality. Provider-based screening would theoretically benefit individuals who have an evolving melanoma that could be detected before the lesion shows clinical signs symptoms that trigger the patient to seek medical attention. The concept of “opportunistic screening” by health providers is supported by data showing that melanomas detected by clinicians as part of a routine physical examination are thinner than those detected by patients or their family members (Terushkin and Halpern 2009; Swetter et al. 2012). While the majority of melanoma deaths are associated with thicker melanomas, 13–27% of melanoma deaths occur in the setting of a primary lesion 1 mm (T1) (Claeson et al. 2017; Criscione and Weinstock 2010). Although about 70% of CM is now diagnosed 1 mm thickness in the United States, Australia, and Western Europe, an estimated 3% of these individuals will die of melanoma (Whiteman et al. 2015), and there is evidence that thinner T1 lesions fare better. Thus, while there are downsides to screening, such as the possibility of an increasing number of false-positive lesions resulting from overscreening, detecting melanoma below 50) or atypical moles could not be ascertained as they were not part of the NHIS. While the USPSTF described evidence as insufficient for population-based

548

screening in 2016, the Task Force also defined the following groups as being at high risk for skin cancer: non-Hispanic white men and women aged >65 years, respondents who reported having been sunburned, and those with a family history of skin cancer, suggesting that targeted screening of atrisk groups may be warranted. Skin self-examination is complementary to clinician skin examination. Melanoma detection occurs more frequently in patients reporting change in a lesion, and some studies have demonstrated decreased thickness and reduced mortality in those who performed skin self-examination (Berwick et al. 1996, 2005). However, the prevalence of thorough SSE is estimated to be only 10–25% among adults in the United States, though rates can be improved by public education and clinician teaching (Weinstock et al. 2007).

Evidence Relating to the Effectiveness of Screening Randomized Trials Other types of cancer screening, for example, programs for breast cancer and colorectal cancer, are justified by evidence from population-based randomized trials that were analyzed on an intent-toscreen basis. These trials showed that mortality is reduced in those people offered screening, compared to a randomized comparison group that was not offered screening. The core of the problem with screening for melanoma is that there have been no randomized trials evaluating either skin examination by a physician or other health provider or by self-screening. As such, there is no direct evidence that skin screening actually reduces the mortality or major morbidity associated with melanoma – a key reason for the current USPSTF “I” statement for skin cancer screening. A randomized trial of a communitybased program utilizing both screening by primary care physicians and self-screening was developed in the 1990s in Queensland, Australia, which has the highest melanoma rates in the world (Elwood 1994; Aitken et al. 2002). This proposal was designed with the power to determine whether the screening intervention could decrease the

S. M. Swetter et al.

mortality rate for melanoma by 20%. It led to a pilot study that succeeded in its key preliminary aim, which was to demonstrate that the frequency of screening in the intervention communities could be increased to a level that provides adequate power for the definitive outcome to be assessed. This level was set as 60% of the population older than 30 years undergoing a wholebody examination by a physician during a 3-year interval (in addition to other screening and selfscreening). This level was achieved, but it stayed close to baseline (approximately 20%) in the control communities (Aitken et al. 2006b). Beyond the Australian screening trial, there have been a number of smaller RCTs to improve skin screening by physicians and SSE practices. Rat and colleagues conducted a pilot clustered randomized controlled trial, comparing a targeted screening and education intervention with a conventional information-based campaign in Western France (Rat et al. 2014). In the intervention group, ten general practitioners identified patients at increased risk for melanoma with a validated assessment tool, the Self-Assessment Melanoma Risk Score (SAMScore), examined their skin, and counseled them using information leaflets. In the control group, ten general practitioners displayed a poster and the leaflets in their waiting room and examined patients’ skin at their own discretion. The main outcome measures were sunbathing and skin self-examinations among patients at elevated risk, assessed 5 months later with a questionnaire. Intervention patients had higher levels of prevention behaviors, e.g., less likely to sunbathe in the summer (24.7% vs. 40.8%, P = 0.048) and more likely to have performed SSE in the past year (52.6% vs. 36.8%, P = 0.029). Robinson et al. sought to evaluate the effect of a structured SSE training intervention for 494 patients with melanoma and their partners (“dyads”) on SSE performance and the detection of new melanomas by the dyad or the physician via a randomized trial that involved skills training (Robinson et al. 2016). Patients in the intervention arms showed significant increases in SSE practices with their partners at 4, 12, and 24 months (P < 0.001 for all) compared with the control group and identified more new melanomas in

Melanoma Prevention and Screening

situ and invasive melanomas – without an associated increase in physician visits. Robinson et al. also conducted a RCT to demonstrate the efficacy of online training for primary care providers (Robinson et al. 2018). A mastery learning course with visual and dermoscopic assessment, diagnosis and management, and deliberate practice with feedback to reach a minimum passing standard was tested with 89 physicians. Primary care providers in the intervention group answered more melanoma detection questions correctly on the posttest, had fewer false-positive and no false-negative melanoma diagnoses, and referred fewer benign lesions as well as detected significantly more melanomas than controls. Youl et al. tested the impact of a theory-based, SMS (text message)-delivered behavioral intervention (Healthy Text) targeting sun protection or SSE behaviors compared to attention control (i.e., time-equivalent messages about physical activity) (Youl et al. 2015). Overall, 546 participants aged 18–42 years were randomized to the SSE (N = 176), sun protection (N = 187), or attention control (N = 183) text messages group. Each group received 21 text messages about their assigned topic over 12 months (12 weekly messages for 3 months, then monthly messages for the next 9 months). One year after baseline, the sun protection and SSE groups had significantly greater improvement in their sun protection habits index compared to the attention control group (reference mean change 0.02). The increase in the proportion of participants who reported any skin self-examination from baseline to 12 months was significantly greater in the SSE intervention group (103/163; 63%; P < 0.001) than the sun protection (83/173; 48%) or attention control (65/165; 36%) groups. Several randomized trials of early detection have also been reported. Among 15 early detection studies reviewed between 2000 and 2015 (Geller et al. 2018), 9 targeted groups are known to be above population risk for melanoma (age 50 and older, NMSC patients [i.e., transplant], and those identified as high risk per brief skin cancer risk assessment tool [BRAT] score) and 6 are targeted members of the general population. The

549

distribution of population scope was as follows: medical patients who did not have a melanoma diagnosis (47%), general population (13%), outdoor workers (13%), beachgoers (7%), primary care office patients (7%), those ages 50 or older (7%), and high risk per BRAT score (7%). Intervention duration ranged from 3.5 mm) fell significantly in women, although no beneficial trend was observed in men (MacKie and Hole 1992). A similar multicenter program in Britain showed no decreases in thick melanoma (Melia 1995). A subsequent randomized trial suggested that educational efforts that used photographs of lesions were more effective than emphasizing the ABCD criteria (Girardi et al. 2006).

Opportunistic Screening in Normal Medical Practice The most widespread type of melanoma screening is opportunistic screening or case finding within normal medical care: performing clinical skin examinations on patients who come to a physician for other reasons, for a regular checkup, or with a request for a skin examination. In a single study of assessing melanoma detection on routine dermatologist FBSE (i.e., opportunistic screening) in a private dermatology practice in Florida, Kantor et al. found a significant association between thinner melanomas as a group (thickness or = 50 years) comprised only 25% of screenees but comprised 44% of those with a confirmed diagnosis of melanoma. The overall yield of melanoma (the number of confirmed diagnoses per the number of screenees) was 1.5 per 1000 screenings (363 diagnoses of 242,374 screenees) compared with a yield of 2.6 per 1000 screenings among men age > or = 50 years. The yield was improved further for men age > or = 50 years who reported either a changing mole (4.6 per 1000 screenings) or skin types I and II (3.8 per 1000 screenings). The predictive value of a screening diagnosis of melanoma was more than twice as high for men age 50 years and older with either a changing mole or skin types I and II compared with all other participants. Euromelanoma Euromelanoma is a pan-European effort to educate and screen Europeans at risk of skin cancer (https://www.euromelanoma.org/). Activities are focused on reaching three key audiences, including the general public, scientific community, and European and national policy makers. Thirtythree countries participate, and the activity culminates in public screenings during an annual “Euromelanoma Screening Day.” To date, over 450,000 people have received free skin examinations. For dermatologists and the broader healthcare community, knowledge and best practices are regularly shared through scientific publications to improve care for skin cancer patients. On the governmental level, Euromelanoma hosts special events to ensure the treatment of skin cancer is fully recognized and supported in health-care systems and legislative policies. An example of one Euromelanoma campaign that took place in Italy in 2010 involved screening of 1085 participants (64.1% females, median age 44 years). Suspicion rate, detection rate, and positive predictive values for melanoma were 1.3%, 0.28%, and 21.4%, respectively (Suppa et al. 2014). Poorly educated, 35-year-old, lighterskinned males were at higher risk for skin cancer than highly educated, 0.75 mm, an increase in the diagnosis of thin and in situ melanomas, and an estimated morality reduction, with no deaths from melanoma in the workforce in later years, compared with expected numbers based on cancer registry data (Schneider et al. 2007). Screening for Selected High-Risk Groups Among high-risk groups, such as persons with previous melanoma, a genetic predisposition, and/or a very strong family history, careful regular surveillance based on clinical observation supplemented by photography has been associated with thinner subsequent primary melanomas (Vecchiato et al. 2014) Targeted screening of high-risk groups is potentially more cost-effective than screening of the general population (Noe et al. 2008), but the validity and cost of the method used to identify the high-risk group are critical factors. Risk factors for melanoma have been reviewed elsewhere (see also chapter ▶ “Clinical Genetics and Risk Assessment of Melanoma”). In addition to fundamental skin color, place of residence and of birth, age, and sex, the presence of many nevi shows the most consistent and strongest association with melanoma, based on either whole-body counts or counts of single body sites, such as the arm. Other risk factors include fair pigmentation; skin that tends to burn easily, tan poorly, and develop freckles; and a history of NMSC (indicating a history of damaging sun exposure). Various risk classification schemes have been used. A risk assessment chart patterned after those that were used successfully for cardiovascular disease produced an estimated 10-year risk of melanoma ranging from less than 0.5% to more than 10% (Whiteman and Green 2005). The risk model included information on place of residence,

S. M. Swetter et al.

age, number of nevi on the arms, and skin color in two categories (olive or fair), also incorporating information on the MC1R variant. More contemporary systematic reviews of risk prediction models concluded that they require further validation and prospective evaluation prior to implementation into clinical practice (Usher-Smith et al. 2014; Vuong et al. 2014).

Adjuncts to Clinical Examination Dermoscopy is well established as a valuable adjunct to clinical examination in dermatology practices, as discussed elsewhere (see also chapter ▶ “Dermoscopy/Confocal Microscopy for Melanoma Diagnosis”). A meta-analysis of studies up to 2001 found that dermoscopy improved performance over clinical examination when used by trained examiners (Bafounta et al. 2001). However, its value when used by primary care providers and the type and necessary amount of training such physicians may require are important issues. In a randomized trial, primary care physicians in Spain and Italy showed improved sensitivity for malignant tumors of all types, with similar specificity, by using dermoscopy assessment with a 1-day training course (Argenziano et al. 2006). The emergence and ongoing investigation of newer imaging technologies (e.g., reflectance confocal microscopy, optical coherence tomography, multiphoton microscopy), molecular adjuncts to bedside diagnosis (e.g., adhesive “tape stripping” for genomic analysis) (Jansen et al. 2018), and artificial intelligence systems (Esteva et al. 2017) may help to reduce unnecessary biopsies and aid in skin cancer diagnosis for PCPS and dermatologists alike.

Challenges in the Detection of More Lethal Melanoma Subtypes Public health efforts to enhance early detection of melanoma emphasize the ABCDE criteria for a suspicious skin lesion (asymmetry, border irregularity, color variegation, diameter more than 6 mm, evolving). Features beyond the ABCD criteria were found to be more important,

Melanoma Prevention and Screening

particularly for desmoplastic or nodular subtypes, which are often clinically amelanotic and deeply invasive at presentation (see also chapter ▶ “Clinical Presentations of Melanoma”). Symptoms such as pain and itching within a lesion also emerged as important in some studies of associated symptoms at presentation (McPherson et al. 2006). A prior systematic review assessing the value of the ABCD system identified only one large study (Thomas et al. 1998) and two smaller studies evaluating the criteria and led to a recommendation of an additional E representing evolving, meaning that the lesion had changed over time (Abbasi et al. 2004). The alphabetical convenience of E, however, may be confusing because E has also been used to mean enlarging in horizontal diameter and elevated. A prospective study in France, involving 135 dermatologists and more than 4000 lesions, assessed clinical decision-making and showed that the most important clinical clues overall were lesion irregularity, recent changes (dependent on the patient’s recall), and the “ugly duckling sign,” meaning that the lesion looked different from other lesions on the patient (Gachon et al. 2005). Emphasis on new or changing nevi and on the outlier lesions that simply look “different” from the others may be a simpler and more effective public health message. Most data support that public health and professional educational campaigns over the past several decades have resulted in earlier recognition of thin superficial spreading melanoma (SSM) (see also chapter ▶ “Clinical Presentations of Melanoma”). Since these lesions grow more slowly, the benefits of earlier diagnosis may be relatively small. More biologically aggressive melanomas are the ones in which early diagnosis could favorably impact survival, since they tend to progress in a shorter time. However, these melanomas often elude early detection. To deal with severe potentially fatal melanoma, greater emphasis should be placed on understanding the detection of deeply invasive melanomas. In a large population-based study in Australia that looked at events and time intervals from first recognition to final diagnosis, a relationship between greater time interval following initial recognition and the lesion being more deeply

557

invasive at diagnosis was only evident for nodular melanomas. Otherwise, a longer time interval did not affect depth and diagnosis (Baade et al. 2006; Berwick 2006). A related issue is the high risk of deeply invasive or nodular melanoma (NM) and increased mortality rates in men older than 50, who warrant specific attention in early diagnosis programs (Chamberlain and Kelly 2004; Geller et al. 2006b; Swetter et al. 2009). Nodular melanoma accounts for approximately half or more of melanomas 2 mm deep in many countries (Demierre et al. 2005). The presenting signs and symptoms and clinical and biologic behavior of NM differ from the more common SSM, which is usually diagnosed when it is thin. Nodular melanoma tends to be less easily detected, highlighting the need for educational campaigns for health professionals and the public that do not rely on the ABCD clinical features, which may not apply to nodular melanoma. Nodular melanomas are often symmetrical, have regular borders, are uniform in color or nonpigmented (clinically amelanotic), and less frequently show changes in color (Chamberlain et al. 2003; Bergenmar et al. 2002). They are also associated with more rapid growth (Liu et al. 2006). Amelanotic melanoma constitutes 2–8% of all melanomas and is more difficult to assess both clinically and by means of dermoscopy (De Giorgi et al. 2006). To investigate the association of SSE, PSE, and patient attitudes with the detection of thinner SSM and NM, researchers from the United States, Greece, and Hungary identified patients with newly diagnosed cutaneous melanoma at four referral hospital centers in the United States, Greece, and Hungary (Dessinioti et al. 2018). Among 920 patients with a primary invasive melanoma, 685 patients with SSM or NM subtype were included. Patients who routinely performed SSE were more likely to be diagnosed with thinner SSM (odds ratio [OR], 2.61; 95% CI, 1.14–5.40) but not thinner NM (OR, 2.39; 95% CI, 0.84–6.80). Self-detected clinical warning signs (e.g., elevation and onset of pain) were markers of thicker SSM and NM. Whole-body PSE was associated with a twofold increase in detection of thinner SSM (OR, 2.25; 95%

558

S. M. Swetter et al.

CI, 1.16–4.35) and thinner NM (OR, 2.67; 95% CI, 1.05–6.82). Patient attitudes and perceptions focusing on increased interest in skin cancer were associated with the detection of thinner NM. In a recent study, researchers tested the hypothesis that primary histologic subtype is an independent predictor of survival and could impact response to treatment in the metastatic setting using the most recent Surveillance, Epidemiology, and End Results (SEER) cohort (n = 118,508) and the New York University (NYU) cohort (n = 1621) with available protocol-driven follow-up (Lattanzi et al. 2018). Nodular melanoma was an independent risk factor for death in both the SEER (hazard ratio [HR] = 1.55, 95% confidence interval [CI] = 1.41–1.70, P < 0.001) and NYU (HR = 1.47; 95% CI = 1.05, 2.07; P = 0.03) cohorts, controlling for thickness, ulceration, stage, and other variables. In the metastatic setting, NM remained an independent risk factor for death upon treatment

with BRAF-targeted therapy (HR = 3.33, 95% CI = 1.06–10.47, P = 0.04) but showed no statistically significant difference with immune checkpoint inhibition. As previously noted, for screening programs to be most effective, they need to capture those people who, in the absence of screening, would present with thicker melanoma and be more likely to die of disease. In an Australian study, patients with thick (>3 mm) primary melanomas were more likely to be men, more than 50 years of age, with nodular melanomas of the head and neck (Chamberlain et al. 2002). Although based on clinical records only, these patients were more likely to have a history of AKs and NMSC (an association that disappeared after controlling for age) but did not have a higher prevalence of freckles or nevi. Recognition of these risk factors by patients and providers may help to promote SSE and appropriate CSE.

War on Melanoma Early Detection Campaign Public Health Education and Outreach

Expand successful elements of the campaign locally and nationally

War on Melanoma Early Detection Campaign

Study the impact of public education on knowledge, cost, and mortality Fig. 4 War on Melanoma™ was initiated in 2014 by Oregon Health & Science University as a public health campaign to promote early detection and reduce mortality from melanoma. The project provides melanoma education

Enlist and train warriors (research participants, medical providers, and community educators)

Implement technology to facilitate screening and research

and traditional and technological solutions to screening and measures impact of the interventional campaign on knowledge, cost, and mortality. Successful elements of the program will be expanded to other states

Melanoma Prevention and Screening

Advances in Screening Technologies and Community Outreach to Improve Early Detection Melanoma early detection is poised to benefit enormously from technological advances in imaging, social and mobile media, and the computational capacity to harness these technologies. Already, large electronic registries, mobile devices, and high-resolution imaging technologies are being deployed, and machine learning and artificial intelligence solutions are being tested and developed (Abbott and Smith 2018; Gareau et al. 2017; Han et al. 2018; Marchetti et al. 2018; Ray et al. 2017; Webster et al. 2017; Yu et al. 2017). Widespread use of Internet communication tools, mobile devices, and social media is permitting crowdsourcing of potential study participants (see www. WarOnMelanoma.org) (Fig. 4) as well as sharing of large nevus and melanoma image collections for research purposes (e.g., International Society for Digital Imaging of the Skin, International Skin Imaging Collaboration, https:// isic-archive.com). These tools are also being used to feed larger, better data sets into ever-improving neural networks, which is empowering the development of artificial intelligence platforms (Esteva et al. 2017; Haenssle et al. 2018; Nasr-Esfahani et al. 2016). However, most of these advances are in previously uncharted territory and will require that we begin to think in broader terms about how they can be rigorously validated, applied, and integrated into existing health and research systems that may be ill-prepared to transform as quickly as the technology does (Leachman and Merlino 2017; Mar and Soyer 2018; Navarrete-Dechent et al. 2018; Safran et al. 2018).

Conclusion For the first time in decades, mortality rates from melanoma are plateauing or even decreasing in several countries, mostly related to earlier detection. Recent progress in therapy for advanced

559

melanoma should also favorably impact mortality rates over time. Although the total incidence continues to increase, this trend is slowing or reversing in younger age groups in some countries. Melanoma prevention depends on reducing solar and artificial UVR exposure beginning in early childhood and adolescence, although the effects of UVR on the development of melanoma are evident throughout the lifespan. Primary prevention strategies to change behaviors with regard to sun exposure may result in reduced incidence, as evidenced by recent trends in Australia after broad implementation of sun protection programs. Screening for melanoma remains controversial, though recent observational studies in Germany and the United States suggest that primary care-based screening is feasible, following education of health providers to perform clinical skin examinations, and increased awareness by the public regarding melanoma clinical warning signs and skin self-exam practices. Burdens on primary care clinician time may prevent widespread implementation of skin cancer screening, though further investigation and use of novel technologies may help to promote earlier detection and appropriate triage. Coupled with a better understanding of therapeutic prevention and the likelihood of large-scale testing of candidate chemopreventive agents in melanoma, the primary and secondary prevention outlook is more promising than ever.

References Abbasi NR, Shaw HM, Rigel DS et al (2004) Early diagnosis of cutaneous melanoma: revisiting the ABCD criteria. JAMA 292:2771 Abbott LM, Smith SD (2018) Smartphone apps for skin cancer diagnosis: implications for patients and practitioners. Australas J Dermatol. https://doi.org/10.1111/ ajd.12758 Abdel-Malek ZA, Swope VB, Starner RJ et al (2014) Melanocortins and the melanocortin 1 receptor, moving translationally towards melanoma prevention. Arch Biochem Biophys 563:4–12 Aguilera P, Carrera C, Puig-Butille JA et al (2013) Benefits of oral Polypodium leucotomos extract in MM high-risk patients. J Eur Acad Dermatol Venereol 27:1095–1100

560 Aitken JF, Elwood JM, Lowe JB et al (2002) A randomised trial of population screening for melanoma. J Med Screen 9:33–37 Aitken JF, Janda M, Elwood M et al (2006a) Clinical outcomes from skin screening clinics within a community-based melanoma screening program. J Am Acad Dermatol 54:105–114 Aitken JF, Elwood M, Baade PD et al (2010) Clinical whole-body skin examination reduces the incidence of thick melanomas. Int J Cancer 126:450–458 Aitken JF, Youi PH, Janda M et al (2006b) Increase in skin cancer screening during a community-based randomised intervention trial. Int J Cancer 118: 1010–1016 Aitkin JF, Youlden DR, Baade PD (2018) Generational shift in melanoma incidence and mortality in Queensland, Australia, 1995–2014. Int J Cancer 142:1528–1535 Albano F, Arcucci A, Granato G et al (2013) Markers of mitochondrial dysfunction during the diclofenacinduced apoptosis in melanoma cell lines. Biochimie 95:934–945 Anderson AM, Matsumoto M, Saul MI et al (2018) Accuracy of skin cancer diagnosis by physician assistants compared with dermatologists in a large health care system. JAMA Dermatol 154:569–573 Argenziano G, Puig S, Zalaudek I et al (2006) Dermoscopy improves accuracy of primary care physicians to triage lesions suggestive of skin cancer. J Clin Oncol 24:1877–1882 Asgari MM, Maruti SS, White E (2008) A large cohort study of nonsteroidal anti-inflammatory drug use and melanoma incidence. J Natl Cancer Inst 100:967–971 Asgari MM, Maruti SS, Kushi LH et al (2009a) A cohort study of vitamin D intake and melanoma risk. J Invest Dermatol 129:1675–1680 Asgari MM, Maruti SS, Kushi LH et al (2009b) Antioxidant supplementation and risk of incident melanomas: results of a large prospective cohort study. Arch Dermatol 145:879–882 Asgari MM, Brasky TM, White E (2012) Association of vitamin A and carotenoid intake with melanoma risk in a large prospective cohort. J Invest Dermatol 132:1573–1582 Autier P, Dore JF, Negrier S et al (1999) Sunscreen use and duration of sun exposure: a double-blind, randomized trial. J Natl Cancer Inst 91:1304–1309 Autier P, Dore JF, Reis AC et al (2000) Sunscreen use and intentional exposure to ultraviolet A and B radiation: a double blind randomized trial using personal dosimeters. Br J Cancer 83:1243–1248 Autier P, Boniol M, Dore JF (2007) Sunscreen use and increased duration of intentional sun exposure: still a burning issue. Int J Cancer 121:1–5 Baade PD, English DR, Youl PH et al (2006) The relationship between melanoma thickness and time to diagnosis in a large population-based study. Arch Dermatol 142:1422–1427

S. M. Swetter et al. Bafounta ML, Beauchet A, Aegerter P et al (2001) Is dermoscopy (epiluminescence microscopy) useful for the diagnosis of melanoma? Results of a metaanalysis using techniques adapted to the evaluation of diagnostic tests. Arch Dermatol 137:1343–1350 Bajaj S, Dusza SW, Marchetti MA et al (2015) Growthcurve modeling of nevi with a peripheral globular pattern. JAMA Dermatol 151:1338–1345 Barnetson RS, Ooi TK, Zhuang L et al (2006) [Nle4-DPhe7]-alpha-melanocyte-stimulating hormone significantly increased pigmentation and decreased UV damage in fair-skinned Caucasian volunteers. J Invest Dermatol 126:1869–1878 Baron ED, Stevens SR (2002) Sunscreens and immune protection. Br J Dermatol 146:933–937 Bastuji-Garin S, Grob JJ, Grognard C et al (1999) Melanoma prevention: evaluation of a health education campaign for primary schools. Arch Dermatol 135: 936–940 Bauer J, Buttner P, Wiecker TS et al (2005a) Risk factors of incident melanocytic nevi: a longitudinal study in a cohort of 1,232 young German children. Int J Cancer 115:121–126 Bauer J, Buttner P, Wiecker TS et al (2005b) Interventional study in 1,232 young German children to prevent the development of melanocytic nevi failed to change sun exposure and sun protective behavior. Int J Cancer 116:755–761 Bergenmar M, Hansson J, Brandberg Y (2002) Detection of nodular and superficial spreading melanoma with tumour thickness < or = 2.0 mm – an interview study. Eur J Cancer Prev 11:49–55 Berwick M (2006) Evaluating early detection in the diagnosis of melanoma. Arch Dermatol 142:1485–1486 Berwick M, Begg CB, Fine JA et al (1996) Screening for cutaneous melanoma by skin self-examination. J Natl Cancer Inst 88:17–23 Berwick M, Armstrong BK, Ben-Porat L et al (2005) Sun exposure and mortality from melanoma. J Natl Cancer Inst 9:195–199 Black WC (2000) Overdiagnosis: an underrecognized cause of confusion and harm in cancer screening. J Natl Cancer Inst 92:1280–1282 Boldeman C, Dal H, Wester U (2004) Swedish pre-school children’s UVR exposure-a comparison between two outdoor environments. Photodermatol Photoimmunol Photomed 20:2–8 Boniol M, Autier P, Boyle P et al (2012) Cutaneous melanoma attributable to sunbed use: systematic review and meta-analysis. BMJ 345:e4757 Boniol M, Autier P, Gandini S (2015) Melanoma mortality following skin cancer screening in Germany. BMJ Open 5(9):e008158 Bonovas S, Nikolopoulos G, Filioussi K et al (2010) Can statin therapy reduce the risk of melanoma? A metaanalysis of randomized controlled trials. Eur J Epidemiol 25:29–35 Brash DE (2015) UV signature mutations. Photochem Photobiol 91:15–26

Melanoma Prevention and Screening Brasky TM, Liu J, White E et al (2014) Non-steroidal antiinflammatory drugs and cancer risk in women: results from the Women’s Health Initiative. Int J Cancer 135:1869–1883 Bregman MD, Buckmeier J, Funk C et al (1987) Median effect and long term recovery analysis of biological modifier interactions with difluoromethylornithine on the proliferation of human melanoma cells. Pigment Cell Res 1:124–128 Breindahl T, Evans-Brown M, Hindersson P et al (2015) Identification and characterization by LC-UV-MS/MS of melanotan II skin-tanning products sold illegally on the Internet. Drug Test Anal 7:164–172 Browning DR, Martin RM (2007) Statins and risk of cancer: a systematic review and metaanalysis. Int J Cancer 120:833–843 Buller DB, Borland R (1999) Skin cancer prevention for children: a critical review. Health Educ Behav 26:317–343 Buller DB, Hall JR, Powers PJ et al (1999) Evaluation of the “Sunny Days, Healthy Ways” sun safety CD-ROM program for children in grades 4 and 5. Cancer Prev Control 3:188–195 Buller DB, Reynolds KD, Yaroch A et al (2006a) Effects of the Sunny Days, Healthy Ways curriculum on students in grades 6 to 8. Am J Prev Med 30:13–22 Buller DB, Taylor AM, Buller MK et al (2006b) Evaluation of the Sunny Days, Healthy Ways sun safety curriculum for children in kindergarten through fifth grade. Pediatr Dermatol 23:321–329 Burian E, Burian E (2013) Eruptive nevi after injection of drugs marketed as melanotan II. The first two Swedish cases described. Lakartidningen 110:208–202 Burke KE, Clive J, Combs GF Jr et al (2003) Effects of topical L-selenomethionine with topical and oral vitamin E on pigmentation and skin cancer induced by ultraviolet irradiation in Skh:2 hairless mice. J Am Acad Dermatol 49:458–472 Burton RC, Armstrong BK (1994) Recent incidence trends imply a non-metastasizing form of invasive melanoma. Melanoma Res 4:107–113 Burton RC, Armstrong BK (1995) Current melanoma epidemic: a nonmetastasizing form of melanoma? World J Surg 19:330–333 Burton RC, Coates MS, Hersey P et al (1993) An analysis of a melanoma epidemic. Int J Cancer 55:765–770 Caini S, Boniol M, Tosti G et al (2014) Vitamin D and melanoma and non-melanoma skin cancer risk and prognosis: a comprehensive review and meta-analysis. Eur J Cancer 50:2649–2658 Cancer Genome Atlas Network (2015) Genomic classification of cutaneous melanoma. Cell 161:1681–1696 Cardones AR, Grichnik JM (2009) alpha-Melanocytestimulating hormone-induced eruptive nevi. Arch Dermatol 145:441–444 Carli P, Biggeri A, Nardini P et al (1998) Sun exposure and large numbers of common and atypical melanocytic naevi: an analytical study in a southern European population. Br J Dermatol 138:422–425

561 Carli P, de Giorgi V, Palli D et al (2003) Dermatologist detection and skin self-examination are associated with thinner melanomas: results from a survey of the Italian Multidisciplinary Group on Melanoma. Arch Dermatol 139:607–612 Carriere P, Baade P, Newman B et al (2007) Cancer screening in Queensland men. Med J Aust 186:404–407 Cassidy PB, Fain HD, Cassidy JP Jr et al (2013) Selenium for the prevention of cutaneous melanoma. Nutrients 5:725–749 Cassidy PB, Liu T, Florell SR et al (2017) A phase II randomized placebo-controlled trial of oral Nacetylcysteine for protection of melanocytic nevi against UV-induced oxidative stress in vivo. Cancer Prev Res (Phila) 10:36–44 Chamberlain AJ, Kelly JW (2004) Nodular melanomas and older men: a major challenge for community surveillance programs. Med J Aust 180:432 Chamberlain AJ, Fritschi L, Giles GG et al (2002) Nodular type and older age as the most significant associations of thick melanoma in Victoria, Australia. Arch Dermatol 138:609–614 Chamberlain AJ, Fritschi L, Kelly JW (2003) Nodular melanoma: patients’ perceptions of presenting features and implications for earlier detection. J Am Acad Dermatol 48:694–701 Chen SC, Bravata DM, Weil E et al (2001) A comparison of dermatologists’ and primary care physicians’ accuracy in diagnosing melanoma: a systematic review. Arch Dermatol 137:1627–1634 Chen AC, Martin AJ, Choy B et al (2015) A phase 3 randomized trial of nicotinamide for skin-cancer chemoprevention. N Engl J Med 373:1618–1626 Chhabra G, Ndiaye MA, Garcia-Peterson LM et al (2017) Melanoma chemoprevention: current status and future prospects. Photochem Photobiol 93:975–989 Claeson M, Gillstedt M, Whiteman DC et al (2017) Lethal melanomas: a population-based registry study in Western Sweden from 1990 to 2014. Acta Derm Venereol 97:1206–1211 Clarke CA, McKinley M, Hurley S et al (2017) Continued increase in melanoma incidence across all socioeconomic status groups in California, 1998–2012. J Invest Dermatol 137:2282–2290 Clemens TL, Adams JS, Henderson SL et al (1982) Increased skin pigment reduces the capacity of skin to synthesize vitamin D3. Lancet 1:74–76 Colantonio S, Bracken MB, Beecker J (2014) The association of indoor tanning and melanoma in adults: systematic review and meta-analysis. J Am Acad Dermatol 70:847–857 Community Preventive Services Task Force (2012) Preventing skin cancer: multicomponent communitywide interventions. https://www.thecommunityguide. org/sites/default/files/assets/Skin-Cancer-Multicompo nent-Communitywide.pdf. Accessed 29 June 2018 Cotter MA, Thomas J, Cassidy P et al (2007) N-acetylcysteine protects melanocytes against oxidative stress/damage and delays onset of ultraviolet-

562 induced melanoma in mice. Clin Cancer Res 13: 5952–5958 Cousen P, Colver G, Helbling I (2009) Eruptive melanocytic naevi following melanotan injection. Br J Dermatol 161:707–708 Crane LA, Deas A, Mokrohisky ST et al (2006) A randomized intervention study of sun protection promotion in well-child care. Prev Med 42:162–170 Criscione VD, Weinstock MA (2010) Melanoma thickness trends in the United States, 1988–2006. J Invest Dermatol 130:793–797 Cristofolini M, Bianchi R, Boi S et al (1993) Analysis of the costeffectiveness ratio of the health campaign for the early diagnosis of cutaneous melanoma in Trentino, Italy. Cancer 71:370–374 Curiel-Lewandrowski C, Nijsten T, Gomez ML et al (2011) Long-term use of nonsteroidal anti-inflammatory drugs decreases the risk of cutaneous melanoma: results of a United States case-control study. J Invest Dermatol 131:1460–1468 Curiel-Lewandrowski C, Swetter SM, Einspahr JG et al (2012) Randomized, double-blind, placebo-controlled trial of sulindac in individuals at risk for melanoma: evaluation of potential chemopreventive activity. Cancer 118:5848–5856 Cust AE, Armstrong BK, Goumas C et al (2011) Sunbed use during adolescence and early adulthood is associated with increased risk of early-onset melanoma. Int J Cancer 128:2425–2435 Darlington S, Siskind V, Green L et al (2002) Longitudinal study of melanocytic nevi in adolescents. J Am Acad Dermatol 46:715–722 Darlington S, Williams G, Neale R et al (2003) A randomized controlled trial to assess sunscreen application and beta carotene supplementation in the prevention of solar keratoses. Arch Dermatol 139:451–455 De Giorgi V, Sestini S, Massi D et al (2006) Dermoscopy for “true” amelanotic melanoma: a clinical dermoscopic-pathologic case study. J Am Acad Dermatol 54:341–344 Deffuant C, Celerier P, Boiteau HL et al (1994) Serum selenium in melanoma and epidermotropic cutaneous T-cell lymphoma. Acta Derm Venereol 74:90–92 Dellavalle RP, Drake A, Graber M et al (2005) Statins and fibrates for preventing melanoma. Cochrane Database Syst Rev 4:CD003697 Demierre MF, Chung C, Miller DR et al (2005) Early detection of thick melanomas in the United States: beware of the nodular subtype. Arch Dermatol 141:745–750 Dennis LK, Beane Freeman LE, VanBeek MJ (2003) Sunscreen use and the risk for melanoma: a quantitative review. Ann Intern Med 139:966–978 Dessinioti C, Geller AC, Stergiopoulou A et al (2018) Association of skin examination behaviors and thinner nodular vs superficial spreading melanoma at diagnosis. JAMA Dermatol 154:544–553 Dickinson SE, Melton TF, Olson ER et al (2009) Inhibition of activator protein-1 by sulforaphane involves

S. M. Swetter et al. interaction with cysteine in the cFos DNA-binding domain: implications for chemoprevention of UVBinduced skin cancer. Cancer Res 69:7103–7110 Dietrich AJ, Olson AL, Sox CH et al (2000) Persistent increase in children’s sun protection in a randomized controlled community trial. Prev Med 31:569–574 Diffey BL (1996) Sunscreens, suntans and skin cancer. People do not apply enough sunscreen for protection. BMJ 313:942 Diffey BL (2000) Has the sun protection factor had its day? BMJ 320:176–177 Dixon HG, Lagerlund M, Spittal MJ et al (2008) Use of sun-protective clothing at outdoor leisure settings from 1992 to 2002: serial cross-sectional observation survey. Cancer Epidemiol Biomarkers Prev 17:428–434 Dobbinson SJ (2008) National sun protection survey: report 2; Australians sun protective behaviors and sunburn incidence on summer weekends, 2006–07 and comparison with 2003–04 in the context of the first national mass media campaign. The Cancer Council Australia and Cancer Australia Dobbinson SJ, Wakefield MA, Jamsen KM et al (2008) Weekend sun protection and sunburn in Australia trends (1987–2002) and association with SunSmart television advertising. Am J Prev Med 34:94–101 Druesne-Pecollo N, Latino-Martel P, Norat T et al (2010) Beta-carotene supplementation and cancer risk: a systematic review and metaanalysis of randomized controlled trials. Int J Cancer 127:172–184 Dusza SW, Oliveria SA, Geller AC et al (2005) Studentparent agreement in self-reported sun behaviors. J Am Acad Dermatol 52:896–900 Dusza SW, Halpern AC, Satagopan JM et al (2012) Prospective study of sunburn and sun behavior patterns during adolescence. Pediatrics 129:309–317 Edwards L, Jaffe P (1990) The effect of topical tretinoin on dysplastic nevi. A preliminary trial. Arch Dermatol 126:494–499 Eide MJ, Asgari MM, Fletcher SW et al (2013) INFORMED (INternet course FOR Melanoma Early Detection) Group. Effects on skills and practice from a web-based skin cancer course for primary care providers. J Am Board Fam Med 26:648–657 Eisemann N, Waldmann A, Holleczek B et al (2017) Observed and expected mortality in the German skin cancer screening pilot project SCREEN. J Med Screen. https://doi.org/10.1177/0969141317734003 Elwood JM (1994) Screening for melanoma and options for its evaluation. J Med Screen 1:22–38 Elwood JM (1996) Skin self-examination and melanoma. J Natl Cancer Inst 88:3–5 Elwood JM, Gallagher RP, Hill GB et al (1985) Cutaneous melanoma in relation to intermittent and constant sun exposure: the Western Canada Melanoma Study. Int J Cancer 35:427–423 Emmons KM, Colditz GA (1999) Preventing excess sun exposure: it is time for a national policy. J Natl Cancer Inst 91:1269–1270

Melanoma Prevention and Screening English DR, Milne E, Jacoby P et al (2005) The effect of a school-based sun protection intervention on the development of melanocytic nevi in children: 6year follow-up. Cancer Epidemiol Biomarkers Prev 14:977–980 Esteva A, Kuprel B, Novoa RA et al (2017) Dermatologistlevel classification of skin cancer with deep neural networks. Nature 542:115–118 Ezzedine K, Latreille J, Kesse-Guyot E et al (2010) Incidence of skin cancers during 5-year follow-up after stopping antioxidant vitamins and mineral supplementation. Eur J Cancer 46:3316–3322 Famenini S, Young LC (2014) Aspirin use and melanoma risk: a review of the literature. J Am Acad Dermatol 70:187–191 Federman DG, Kravetz JD, Kirsner RS (2002) Skin cancer screening by dermatologists: prevalence and barriers. J Am Acad Dermatol 46:710–714 Ferrándiz-Pulido C, Fernández-Figueras MT, Quer A et al (2011) An eruptive pigmented lesion after melanotan injection. Clin Exp Dermatol 36:801–802 Ferris LK, Saul MI, Lin Y et al (2017) A large skin cancer screening quality initiative: description and first-year outcomes. JAMA Oncol 3:1112–1115 Feskanich D, Willett WC, Hunter DJ et al (2003) Dietary intake of vitamins A, C, and E and risk of melanoma in two cohorts of women. Br J Cancer 88:1381–1387 Fisher NM, Schaffer JV, Berwick M et al (2005) Breslow depth of cutaneous melanoma: impact of factors related to surveillance of the skin, including prior skin biopsies and family history of melanoma. J Am Acad Dermatol 53:393–406 Fisher ML, Adhikary G, Grun D et al (2015) The Ezh2 polycomb group protein drives an aggressive phenotype in melanoma cancer stem cells and is a target of diet derived sulforaphane. Mol Carcinog 55: 2024–2036 Fonseca M, Burnett M, Dusza S et al (2014) Dermoscopic patterns of nevi have a distinct anatomical distribution in adolescents. J Invest Dermatol 134(Suppl 1): S49–S60 Fonseca M, Marchetti MA, Chung E et al (2015) Crosssectional analysis of the dermoscopic patterns and structures of melanocytic naevi on the back and legs of adolescents. Br J Dermatol 173:1486–1493 Freedberg KA, Geller AC, Miller DR et al (1999) Screening for malignant melanoma: a cost-effectiveness analysis. J Am Acad Dermatol 41:738–745 Freeman SR, Drake AL, Heilig LF et al (2006) Statins, fibrates, and melanoma risk: a systematic review and meta-analysis. J Natl Cancer Inst 98:1538–1546 Gachon J, Beaulieu P, Sei JF et al (2005) First prospective study of the recognition process of melanoma in dermatological practice. Arch Dermatol 141:434–438 Gallagher RP, Rivers JK, Lee TK et al (2000) Broadspectrum sunscreen use and the development of new naevi in white children. A randomized controlled trial. JAMA 283:2955–2960

563 Gamba CA, Swetter SM, Stefanick ML et al (2013) Aspirin is associated with lower melanoma risk among postmenopausal Caucasian women: the Women’s Health Initiative. Cancer 119:1562–1569 Gandini S, Raimondi S, Gnagnarella P et al (2009) Vitamin D and skin cancer: a meta-analysis. Eur J Cancer 45:634–461 Garbe C, Buttner P, Weiss J et al (1994) Risk factors for developing cutaneous melanoma and criteria for identifying persons at risk: multicentre case-control study of the Central Malignant Melanoma Registry of the German Dermatological Society. J Invest Dermatol 102:695–699 Gareau DS, Correa da Rosa J, Yagerman S et al (2017) Digital imaging biomarkers feed machine learning for melanoma screening. Exp Dermatol 26:615–618 Garland M, Morris JS, Stampfer MJ et al (1995) Prospective study of toenail selenium levels and cancer among women. J Natl Cancer Inst 87:497–505 Gasparro FP (2000) Sunscreens, skin photobiology, and skin cancer: the need for UVA protection and evaluation of efficacy. Environ Health Perspect 108:71–78 Gasparro FP, Mitchnick M, Nash JF (1998) A review of sunscreen safety and efficacy. Photochem Photobiol 68:243–256 Geller AC, Sober AJ, Zhang Z et al (2002) Strategies for improving melanoma education and screening for men age >or = 50 years: findings from the American Academy of Dermatological National Skin Cancer Screening Program. Cancer 95:1554–1561 Geller AC, Emmons KM, Brooks DR et al (2006a) A randomized trial to improve early detection and prevention practices among siblings of melanoma patients. Cancer 107:806–814 Geller J, Swetter SM, Leyson J et al (2006b) Crafting a melanoma educational campaign to reach middle-aged and older men. J Cutan Med Surg 10:259–268 Geller AC, Oliveria SA, Bishop M et al (2007) Study of health outcomes in school children: key challenges and lessons learned from the Framingham Schools’ Natural History of Nevi Study. J Sch Health 77:312–318 Geller AC, Johnson TM, Miller DR et al (2009) Factors associated with physician discovery of early melanoma in middle-aged and older men. Arch Dermatol 145: 409–414 Geller AC, Dickerman BA, Taber JM et al (2018) Skin cancer interventions across the cancer control continuum: a review of experimental evidence (1/ 1/2000–6/30/2015) and future research directions. Prev Med 111:442–450 Gershenwald JE, Scolyer RA, Hess KR et al (2017) Melanoma staging: evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67:472–492 Ghiasvand R, Weiderpass E, Green AC et al (2016) Sunscreen use and subsequent melanoma risk: a population-based cohort study. J Clin Oncol 34:3976–3983 Ghiasvand R, Rueegg CS, Weiderpass E et al (2017) Indoor tanning and melanoma risk: long-term evidence

564 from a prospective population-based cohort study. Am J Epidemiol 185:147–156 Gies G, Javorniczky J, Colin Roy C et al (2006) Measurements of the UVR protection provided by hats used at school. Photochem Photobiol 82:750–754 Giles GG, Armstrong BK, Burton RC et al (1996) Has mortality from melanoma stopped rising in Australia? Analysis of trends between 1931 and 1994. BMJ 312:1121–1125 Giovannucci E, Liu Y, Rimm EB et al (2006) Prospective study of predictors of vitamin D status and cancer incidence and mortality in men. J Natl Cancer Inst 98:451–459 Girardi S, Gaudy C, Gouvernet J et al (2006) Superiority of a cognitive education with photographs over ABCD criteria in the education of the general population to the early detection of melanoma: a randomized study. Int J Cancer 118:2276–2280 Girgis A, Clarke P, Burton RC et al (1996) Screening for melanoma by primary health care physicians: a cost-effectiveness analysis. J Med Screen 3:47–53 Glanz K, Lew RA, Song V et al (2000) Skin cancer prevention in outdoor recreation settings: effects of the Hawaii Sun Smart Program. Eff Clin Pract 3:53–61 Glazer AM, Winkelmann RR, Farberg AS et al (2016) Analysis of trends in US melanoma incidence and mortality. JAMA Dermatol. https://doi.org/10.1001/ jamadermatol.2016.4512 Godar DE, Urbach F, Gasparro FP et al (2003) UV doses of young adults. Photochem Photobiol 77:453–457 Goodman JR, Grossman D (2014) Aspirin and other NSAIDs as chemoprevention agents in melanoma. Cancer Prev Res (Phila) 7:557–564 Goodson AG, Cotter MA, Cassidy P et al (2009) Use of oral N-acetylcysteine for protection of melanocytic nevi against UV-induced oxidative stress: towards a novel paradigm for melanoma chemoprevention. Clin Cancer Res 15:7434–7440 Gordon LG, Rowell D (2015) Health system costs of skin cancer and cost-effectiveness of skin cancer prevention and screening: a systematic review. Eur J Cancer Prev 24:141–149 Gordon LG, Brynes J, Baade PD et al (2017) Cost-effectiveness analysis of a skin awareness intervention for early detection of skin cancer targeting men older than 50 years. Value Health 20:593–601 Gorham ED, Mohr SB, Garland CF et al (2007) Do sunscreens increase risk of melanoma in populations residing at higher latitudes? Ann Epidemiol 17: 956–963 Goulart JM, Quigley EA, Dusza S et al (2011) Skin cancer education for primary care physicians: a systematic review of published evaluated interventions. J Gen Intern Med 26:1027–1035 Green A, Williams G, Neale R et al (1999) Daily sunscreen application and betacarotene supplementation in prevention of basal-cell and squamous-cell carcinomas of the skin: a randomised controlled trial. Lancet 354:723–729

S. M. Swetter et al. Green AC, Wallingford SC, McBride P (2011a) Childhood exposure to ultraviolet radiation and harmful skin effects: epidemiological evidence. Prog Biophys Mol Biol 107:349–355 Green AC, Williams GM, Logan V et al (2011b) Reduced melanoma after regular sunscreen use: randomized trial follow-up. J Clin Oncol 29:257–263 Grob JJ, Andrae L, Romano MH et al (1988) Dysplastic naevus in non-familial melanoma. A clinicopathological study of 101 cases. Br J Dermatol 118:745–752 Grover SR, Morley R (2001) Vitamin D deficiency in veiled or dark-skinned pregnant women. Med J Aust 175:251–252 Guy GP, Thomas CC, Thompson et al (2015) Melanoma incidence and mortality trends and projections - United States, 1982-2030. CDC Morb Mortal Wkly Rep 64:591–596 Guy GP, Berkowitz Z, Everett Jones S et al (2017) Prevalence of indoor tanning and association with sunburn among youth in the United States. JAMA Dermatol 153:387–390 Hadley ME (2005) Discovery that a melanocortin regulates sexual functions in male and female humans. Peptides 26:1687–1689 Haenssle HA, Fink C, Schneiderbauer R et al (2018) Man against machine: diagnostic performance of a deep learning convolutional neural network for dermoscopic melanoma recognition in comparison to 58 dermatologists. Ann Oncol. https://doi.org/10.1093/ annonc/mdy166 Han SS, Kim MS, Lim W et al (2018) Classification of the clinical images for benign and malignant cutaneous tumors using a deep learning algorithm. J Invest Dermatol 138:1529–1538 Hay JL, Geller AC, Schoenhammer M et al (2014) Tanning and beauty: mother and teenage daughters in discussion. J Health Psychol 58(5):579–581 Hennessey RC, Holderbaum AM, Bonilla A et al (2011) Ultraviolet radiation accelerates NRas-mutant melanomagenesis: a cooperative effect blocked by sunscreen. Pigment Cell Melanoma Res 30:477–487 Hercberg S, Ezzedine K, Guinot C et al (2007) Antioxidant supplementation increases the risk of skin cancers in women but not in men. J Nutr 137:2098–2105 Hill D (1999) Efficacy of sunscreens in protection against skin cancer. Lancet 354:699–700 Hill D, White V, Marks R et al (1992) Melanoma prevention: behavioral and nonbehavioral factors in sunburn among an Australian urban population. Prev Med 21:654–659 Hodi E, Watson IR, Kryukov GV et al (2012) A landscape of driver mutations in melanoma. Cell 150:251–263 Holman CD, Armstrong BK, Heenan PJ (1986) Relationship of cutaneous malignant melanoma to individual sunlight-exposure habits. Natl Cancer Inst 76: 403–414 Holman DM, Berkowitz Z, Guy GP Jr et al (2015) Patterns of sunscreen use on the face and other exposed skin among US adults. J Am Acad Dermatol 73:83–92.e1

Melanoma Prevention and Screening Holman DM, Freeman MB, Shoemaker ML (2018) Trends in melanoma incidence among non-Hispanic whites in the United States, 2005 to 2014. JAMA Dermatol 154:361–362 Hoorens I, Vossaert K, Pil L et al (2016) Total-body examination vs lesion-directed skin cancer screening. JAMA Dermatol 152:27–34 Horike N, Kumagai A, Shimono Y et al (2010) Downregulation of SIK2 expression promotes the melanogenic program in mice. Pigment Cell Melanoma Res 23:809–819 Hu H, Xie Y, Yang G et al (2014) Nonsteroidal anti-inflammatory drug use and the risk of melanoma: a metaanalysis. Eur J Cancer Prev 23:62–68 Hübner J, Waldmann A, Geller AC et al (2017) Interval cancers after skin cancer screening: incidence, tumour characteristics and risk factors for cutaneous melanoma. Br J Cancer 116:253–259 Hughes AM, Armstrong BK, Vajdic CM et al (2004) Sun exposure may protect against non-Hodgkin lymphoma: a case-control study. Int J Cancer 112:865–871 Huncharek M, Kupelnick B (2002) Use of topical sunscreens and the risk of malignant melanoma: a metaanalysis of 9067 patients from 11 case-control studies. Am J Public Health 92:1173–1177 Institute of Medicine, Food and Nutrition Board (2010) Dietary reference intakes for calcium and vitamin D. National Academy Press, Washington, DC International Agency for Research on Cancer, Working Group on artificial ultraviolet (UV) light and skin cancer (2007) The association of use of sunbeds with cutaneous malignant melanoma and other skin cancer: a systematic review. Int J Cancer 120:1116–2112 Islami F, Goding Sauer A et al (2018) Proportion and number of cancer cases and deaths attributable to potentially modifiable risk factors in the United States. CA Cancer J Clin 68:31–54 Jacobs EJ, Thun MJ, Bain EB et al (2007) A large cohort study of long-term daily use of adult-strength aspirin and cancer incidence. J Natl Cancer Inst 99:608–615 Jacobs EJ, Newton CC, Thun MJ et al (2011) Long-term use of cholesterol-lowering drugs and cancer incidence in a large United States cohort. Cancer Res 71: 1763–1771 Jagtap D, Rosenberg CA, Martin LW et al (2012) Prospective analysis of association between use of statins and melanoma risk in the Women’s Health Initiative. Cancer 118:5124–5131 Janda M, Lowe JB, Elwood M et al (2006a) Do centralised skin screening clinics increase participation in melanoma screening (Australia)? Cancer Causes Control 17:161–168 Janda M, Youl PH, Lowe JB et al (2006b) What motivates men age > or = 50 years to participate in a screening program for melanoma? Cancer 107:815–823 Jansen R, Osterwalder U, Wang SQ et al (2013) Photoprotection: part II. Sunscreen: development, efficacy, and controversies. J Am Acad Dermatol 69:867. e1–867.e14

565 Jansen B, Hansen D, Moy R et al (2018) Gene expression analysis differentiates melanomas from Spitz nevi. J Drugs Dermatol 17:574–576 Jeter JM, Han J, Martinez ME et al (2012) Non-steroidal anti-inflammatory drugs, acetaminophen, and risk of skin cancer in the Nurses’ Health Study. Cancer Causes Control 3:1451–1461 Jeter J, Bowles T, Curiel-Lewandrowski C et al (2018) Chemoprevention agents for melanoma: a path forward into phase III clinical trials. Cancer. https://doi.org/ 10.1002/cncr.31719 Johannesdottir SA, Chang ET, Mehnert F et al (2012) Nonsteroidal anti-inflammatory drugs and the risk of skin cancer: a population-based case-control study. Cancer 118:4768–4776 Johnson JL, Lowell BC, Ryabinina OP et al (2011) TATmediated delivery of a DNA repair enzyme to skin cells rapidly initiates repair of UV-induced DNA damage. J Invest Dermatol 131:753–761 Joosse A, Koomen ER, Casparie MK et al (2009) Nonsteroidal anti-inflammatory drugs and melanoma risk: large Dutch population-based case-control study. J Invest Dermatol 129(11):2620–2627 Kann L, McManus T, Harris WA et al (2018) Centers for Disease Control youth risk behavior surveillance – United States, 2017. MMWR Surveill Summ 67(8):1 Kantor J, Kantor DE (2009) Routine dermatologistperformed full-body skin examination and early melanoma detection. Arch Dermatol 145(8):873–876 Kast RE (2007) Melanoma inhibition by cyclooxygenase inhibitors: role of interleukin-6 suppression, a putative mechanism of action, and clinical implications. Med Oncol 24:1–6 Katalinic A, Waldmann A, Weinstock MA et al (2012) Does skin cancer screening save lives?: an observational study comparing trends in melanoma mortality in regions with and without screening. Cancer 118:5395–5402 Katalinic A, Eisemann N, Waldmann A (2015) Skin cancer screening in Germany. Documenting melanoma incidence and mortality from 2008 to 2013. Dtsch Arztebl Int 112:629–634 Kawakami A, Fisher DE (2017) The master role of microphthalmia-associated transcription factor in melanocyte and melanoma biology. Lab Invest 97:649–656 Kimlin MG, Tenkate TD (2007) Occupational exposure to ultraviolet radiation: the duality dilemma. Rev Environ Health 22:1–37 Klein EA, Thompson IM Jr, Tangen CM et al (2011) Vitamin E and the risk of prostate cancer: the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 306:1549–1556 Klug HL, Tooze JA, Graff-Cherry C et al (2010) Sunscreen prevention of melanoma in man and mouse. Pigment Cell Melanoma Res 23:835–837 Kohli I, Shafi R, Isedeh P et al (2017) The impact of oral Polypodium leucotomos extract on ultraviolet B response: a human clinical study. J Am Acad Dermatol 77:33–41

566 Kumar R, Deep G, Agarwal R (2015) An overview of ultraviolet B radiation-induced skin cancer chemoprevention by silibinin. Curr Pharmacol Rep 1:206–215 Kuoppala J, Lamminpaa A, Pukkala E (2008) Statins and cancer: a systematic review and meta-analysis. Eur J Cancer 44:2122–2132 Kwon G, Gamba CS, Stefanick ML et al (2018) Association of 25-hydroxyvitamin D levels and cutaneous melanoma: a nested case-control study of the Women’s Health Initiative Observation Study. J Am Acad Dermatol 79:145–147 Lakhani NA, Saraiya M, Thompson TD et al (2014) Total body skin examination for skin cancer screening among U.S. adults from 2000 to 2010. Prev Med 61:75–80 Langendonk JG, Balwani M, Anderson KE et al (2015) Afamelanotide for erythropoietic protoporphyria. N Engl J Med 373:48–59 Lattanzi M, Lee Y, Simpson D et al (2018) Primary melanoma histologic subtype: impact on survival and response to therapy. J Natl Cancer Inst. https://doi.org/ 10.1093/jnci/djy086 LaVigne EA, Oliveria SA, Dusza SW et al (2005) Clinical and dermoscopic changes in common melanocytic nevi in school children: the Framingham school nevus study. Dermatology 211:234–239 Lazovich D, Stryker JE, Mayer JA et al (2008) Measuring nonsolar tanning behavior: indoor and sunless tanning. Arch Dermatol 144:225–230 Lazovich D, Isaksson Vogel R et al (2016) Association between indoor tanning and melanoma in younger men and women. JAMA Dermatol 152:268–275 Le Gal K, Ibrahim MX, Wiel C et al (2015) Antioxidants can increase melanoma metastasis in mice. Sci Transl Med 7:308re8 Leachman SA, Merlino G (2017) Medicine: the final frontier in cancer diagnosis. Nature 542:36–38 Leblanc WG, Vidal L, Kirsner RS et al (2008) Reported skin cancer screening of US adult workers. J Am Acad Dermatol 59:55–63 Lee IM, Cook NR, Manson JE et al (1999) Beta-carotene supplementation and incidence of cancer and cardiovascular disease: the Women’s Health Study. J Natl Cancer Inst 91:2102–2106 Lee TK, Rivers JK, Gallagher RP (2005) Site-specific protective effect of broad-spectrum sunscreen on nevus development among white schoolchildren in a randomized trial. J Am Acad Dermatol 52:786–792 Li S, Liu Y, Zeng Z et al (2013) Association between nonsteroidal anti-inflammatory drug use and melanoma risk: a meta-analysis of 13 studies. Cancer Causes Control 24:1505–1516 Li X, Wu XB, Chen Q (2014) Statin use is not associated with reduced risk of skin cancer: a meta-analysis. Br J Cancer 110:802–807 Linden KG, Leachman SA, Zager JS et al (2014) A randomized, double-blind, placebo-controlled phase II clinical trial of lovastatin for various endpoints of melanoma pathobiology. Cancer Prev Res (Phila) 7:496–504

S. M. Swetter et al. Linos E, Swetter SM, Cockburn MG et al (2009) Increasing burden of melanoma in the United States. J Invest Dermatol 129:1666–1674 Lips P (2001) Vitamin D deficiency and secondary hyperparathyroidism in the elderly: consequences for bone loss and fractures and therapeutic implications. Endocr Rev 22:477–501 Liu W, Dowling JP, Murray WK et al (2006) Rate of growth in melanomas: characteristics and associations of rapidly growing melanomas. Arch Dermatol 142:1551–1558 Losina E, Walensky RP, Geller A et al (2007) Visual screening for malignant melanoma: a cost-effectiveness analysis. Arch Dermatol 143:21–28 Luther H, Altmeyer P, Garbe C et al (1996) Increase of melanocytic nevus counts in children during 5 years of follow-up and analysis of associated factors. Arch Dermatol 132:1473–1478 MacKie RM, Hole D (1992) Audit of public education campaign to encourage earlier detection of malignant melanoma. BMJ 304:1012–1015 Mahabir S, Frenkel K, Brady MS et al (2004) Randomized, placebo-controlled trial of dietary supplementation of alpha-tocopherol on mutagen sensitivity levels in melanoma patients: a pilot trial. Melanoma Res 14:49–56 Mancuso JB, Maruthi R, Wang SQ et al (2017) Sunscreens: an update. Am J Clin Dermatol 18:643–650 Mar VJ, Soyer HP (2018) Artificial intelligence for melanoma diagnosis: how can we deliver on the promise? Ann Oncol. https://doi.org/10.1093/annonc/mdy191 Marchetti MA, Fonseca M, Dusza SW et al (2015) Dermatoscopic imaging of skin lesions by high school students: a cross-sectional pilot study. Dermatol Pract Concept 5:11–28 Marchetti MA, Codella NCF, Dusza SW, International Skin Imaging Collaboration et al (2018) Results of the 2016 international skin imaging collaboration international symposium on biomedical imaging challenge: comparison of the accuracy of computer algorithms to dermatologists for the diagnosis of melanoma from dermoscopic images. J Am Acad Dermatol 78:270–277 Mathieu C, Badenhoop K (2005) Vitamin D and type I diabetes mellitus: state of the art. Trends Endocrinol Metab 16:261–266 Matsumoto M, Secrest A, Anderson A et al (2018) Estimating the cost of skin cancer detection by dermatology providers in a large health care system. J Am Acad Dermatol 78:701–709 Maxwell SR (1995) Prospects for the use of antioxidant therapies. Drugs 49:345–361 Mayer JA, Woodruff SI, Slymen DJ et al (2011) Adolescents’ use of indoor tanning: a large-scale evaluation of psychosocial, environmental, and policylevel correlates. Am J Public Health 101:930–938 McPherson M, Elwood M, English DR et al (2006) Presentation and detection of invasive melanoma in a high-risk population. J Am Acad Dermatol 54: 783–792

Melanoma Prevention and Screening MD Anderson News Release October 30, 2017 (2017) https://www.mdanderson.org/newsroom/2017/10/bay town-community-launches-be-well-baytown-initiativeled-by-md-anderson-exxonmobil.html. Accessed 20 July 2018 Melia J (1995) Early detection of cutaneous malignant melanoma in Britain. Int J Epidemiol 24(Suppl 1): S39–S44 Meyskens FL Jr, McLaren CE, Pelot D et al (2008) Difluoromethylornithine plus sulindac for the prevention of sporadic colorectal adenomas: a randomized placebo-controlled, double-blind trial. Cancer Prev Res (Phila) 1:32–38 Millen AE, Tucker MA, Hartge P et al (2004) Diet and melanoma in a case-control study. Cancer Epidemiol Biomarkers Prev 13:1042–1051 Miller DR, Geller AC, Wood MC et al (1999) The Falmouth Safe Skin Project: evaluation of a community program to promote sun protection in youth. Health Educ Behav 26:369–384 Milne E, English DR, Johnston R et al (2000) Improved sun protection behaviour in children after two years of the Kidskin intervention. Aust N Z J Public Health 24:481–487 Milne E, Johnston R, Cross D et al (2002) Effect of a schoolbased sun-protection intervention on the development of melanocytic nevi in children. Am J Epidemiol 155:739–745 Milne E, Simpson JA, English DR (2008) Appearance of melanocytic nevi on the backs of young Australian children: a 7-year longitudinal study. Melanoma Res 18:22–28 Minocha R, Damian DL, Halliday GM et al (2017) Melanoma and nonmelanoma skin cancer chemoprevention: a role for nicotinamide? Photodermatol Photoimmunol Photomed 34:5–12 Montague M, Borland R, Sinclair C (2001) Slip! Slop! Slap! and SunSmart, 1980–2000: skin cancer control and 20 years of population-based campaigning. Health Educ Behav 28:290–305 Moore MM, Geller AC, Zhang Z et al (2006) Skin cancer examination teaching in US medical education. Arch Dermatol 142:439–444 Mounessa J et al (2016) Primary and secondary chemoprevention of malignant melanoma. Am J Clin Dermatol 17(6):625–634 Moyer VA, US Preventive Services Task Force (2012) Skin cancer: behavioral counseling to prevent skin cancer. Ann Intern Med 157:59–65 Mujahid N, Liang Y, Murakami R et al (2017) A UVindependent topical small-molecule approach for melanin production in human skin. Cell Rep 19:2177–2184 Naldi L, Gallus S, Tavani A et al (2004) Risk of melanoma and vitamin A, coffee and alcohol: a case-control study from Italy. Eur J Cancer Prev 3:503–508 Nasr-Esfahani E, Samavi S, Karimi N et al (2016) Melanoma detection by analysis of clinical images using convolutional neural network. Conf Proc IEEE Eng Med Biol Soc 2016:1373–1376

567 National Center for Education on Statistics. Fast facts: back to school statistics. https://nces.ed.gov/fastfacts/ display.asp?id=372. Accessed 20 July 2018 National Institutes of Health, Office of Dietary Supplements Vitamin D. https://ods.od.nih.gov/factsheets/ VitaminD-HealthProfessional/. Accessed 15 June 2018 Navarrete-Dechent C, Dusza SW, Liopyris K et al (2018) Automated dermatological diagnosis: hype or reality? J Invest Dermatol 138:2277. pii: S0022-202X(18) 31991-2 New South Wales Cancer Council (1998) Report on the Seymour Snowman Sun Protection Campaign (1997–1998). New South Wales Cancer Council, North Sydney Newton-Bishop JA, Davies JR, Latheef F et al (2015) 25Hydroxyvitamin D2/D3 levels and factors associated with systemic inflammation and melanoma survival in the Leeds Melanoma Cohort. Int J Cancer 136:2890–2899 Noe M, Schroy P, Demierre MF et al (2008) Increased cancer risk for individuals with a family history of prostate cancer, colorectal cancer, and melanoma and their associated screening recommendations and practices. Cancer Causes Control 19:1–12 Nozza JM, Rodda CP (2001) Vitamin D deficiency in mothers of infants with rickets. Med J Aust 175: 253–255 Office of National Statistics Report (2010) SunSmart Trends 03-10 summary report (2010) Okhovat JP, Beaulieu D, Tsao H et al (2018) The first 30 years of the American Academy of Dermatology skin cancer screening program: 1985–2014. J Am Acad Dermatol 79:884. in press Oliveria SA, Geller AC, Dusza SW et al (2004) The Framingham school nevus study: a pilot study. Arch Dermatol 140:545–551 Oliveria SA, Satagopan JM, Geller AC et al (2009) Study of Nevi in Children (SONIC): baseline findings and predictors of nevus count. Am J Epidemiol 169:41–53 Oliveria SA, Scope A, Satagopan JM et al (2014) Factors associated with nevus volatility in early adolescence. J Invest Dermatol 134:2469–2471 Olson AL, Gaffney C, Starr P et al (2007) SunSafe in the Middle School Years: a community-wide intervention to change early-adolescent sun protection. Pediatrics 119:e247–e256 Orlow I, Satagopan JM, Berwick M et al (2014) Genetic factors associated with naevus count and dermoscopic patterns: preliminary results from the Study of Nevi in Children (SONIC). Br J Dermatol 172:1081–1089 Orlow I, Reiner AS, Thomas NE et al (2016) Vitamin D receptor polymorphisms and survival in patients with cutaneous melanoma: a population-based study. Carcinogenesis 37:30–33 Osterlind A, Tucker MA, Stone BJ et al (1988) The Danish casecontrol study of cutaneous malignant melanoma. II. Importance of UV-light exposure. Int J Cancer 42:319–324

568 Palsuledesai CC, Distefano MD (2015) Protein prenylation: enzymes, therapeutics, and biotechnology applications. ACS Chem Biol 10:51–62 Papadimitropoulos E, Wells G, Shea B et al (2002) Metaanalysis of the efficacy of vitamin D treatment in preventing osteoporosis in postmenopausal women. Endocr Rev 23:560–569 Parrado C, Mascaraque M, Gilaberte Y et al (2016) Fernblock (Polypodium leucotomos extract): molecular mechanisms and pleiotropic effects in light-related skin conditions, photoaging and skin cancers, a review. Int J Mol Sci 17(7). pii: E1026. https://doi.org/10.3390/ ijms17071026 Parsons PG, Neale R, Wolski P et al (1998) The shady side of solar protection. Med J Aust 168:327–330 Pawlak MT, Bui M, Amir M et al (2012) Legislation restricting access to indoor tanning throughout the world. Arch Dermatol 148:1006–1012 Pei Y, Liu H, Yang Y et al (2018) Biological activities and potential oral applications of N-acetylcysteine: progress and prospects. Oxid Med Cell Longev 2018:2835787 Placzek M, Gaube S, Kerkmann U et al (2005) Ultraviolet B-induced DNA damage in human epidermis is modified by the antioxidants ascorbic acid and D-alphatocopherol. J Invest Dermatol 124:304–307 Pollitt RA, Geller AC, Brooks DR et al (2009) Efficacy of skin self-examination practices for early melanoma detection. Cancer Epidemiol Biomarkers Prev 18: 3018–3023 Purdue MP, Freeman LE, Anderson WF et al (2008) Recent trends in incidence of cutaneous melanoma among US Caucasian young adults. J Invest Dermatol 128:2905–2908 Qin J, Holman DM, Jones SE et al (2018) The association between state indoor tanning laws and adolescent indoor tanning. Am J Public Health 108:951–956 Rassaby JL, Hill D, Wake R (1983) Slip Slop Slap: health education about skin cancer. Cancer Forum 7:69 Rat C, Quereux G, Riviere C et al (2014) Targeted melanoma prevention intervention: a cluster randomized controlled trial. Ann Fam Med 12:21–28 Ray PP, Dash D, De D (2017) A systematic review of wearable systems for cancer detection: current state and challenges. J Med Syst 41:180 Rigby CM, Roy S, Deep G et al (2017) Role of p53 in silibinin-mediated inhibition of ultraviolet B radiation-induced DNA damage, inflammation and skin carcinogenesis. Carcinogenesis 38:40–50 Riggs BL (2003) Role of the vitamin D-endocrine system in the pathophysiology of postmenopausal osteoporosis. J Cell Biochem 88:209–215 Risica PM, Matthews NH, Dionne L et al (2018) Psychosocial consequences of skin cancer screening. Prev Med Rep 10:310–316 Robinson JK, Wayne JD, Martini MC et al (2016) Early detection of new melanomas by patients with melanoma and their partners using a structured skin self-

S. M. Swetter et al. examination skills training intervention: a randomized clinical trial. JAMA Dermatol 152:979–985 Robinson JK, Jain N, Marghoob AA et al (2018) A randomized trial on the efficacy of mastery learning for primary care provider melanoma opportunistic screening skills and practice. J Gen Intern Med 33: 855–862 Robsahm TE, Helsing P, Nilssen Y et al (2018) High mortality due to cutaneous melanoma in Norway: a study of prognostic factors in a nationwide cancer registry. Clin Epidemiol 10:537–548 Roman MP, Jitaru P, Barbante C (2014) Selenium biochemistry and its role for human health. Metallomics 6:25–54 Rosenthal FS, West SK, Munoz B et al (1991) Ocular and facial skin exposure to ultraviolet radiation in sunlight. Health Phys 61:77–86 Sacchetto L, Zanetti R, Comber H et al (2018) Trends in incidence of thick, thin and in situ melanoma in Europe. Eur J Cancer 92:108–118 Safran T, Viezel-Mathieu A, Corban J et al (2018) Machine learning and melanoma: the future of screening. J Am Acad Dermatol 78:620–621 Saiag P, Aegerter P, Vitoux D et al (2015) Prognostic value of 25-hydroxyvitamin D3 levels at diagnosis and during follow-up in melanoma patients. J Natl Cancer Inst 107:djv264 Saito A, Saito N, Mol W et al (2008) Simvastatin inhibits growth via apoptosis and the induction of cell cycle arrest in human melanoma cells. Melanoma Res 18:85–94 Samanek AJ, Croager EJ, Gies P et al (2006) Estimates of beneficial and harmful sun exposure times during the year for major Australian population centres. Med J Aust 184:338–341 Saraiya M, Glanz K, Briss PA et al (2004) Interventions to prevent skin cancer by reducing exposure to ultraviolet radiation: a systematic review. Am J Prev Med 27: 422–466 Satagopan JM, Oliveria SA, Arora A et al (2015) Sunburn, sun exposure, and sun sensitivity in the Study of Nevi in Children. Ann Epidemiol 25:839–843 Schilling K, Bradford B, Castelli D et al (2010) Human safety review of “nano” titanium dioxide and zinc oxide. Photochem Photobiol Sci 9:495–509 Schneider JS, Moore DH, Sagebiel RW (1994) Risk factors for melanoma incidence in prospective follow-up. The importance of atypical (dysplastic) nevi. Arch Dermatol 130:1002–1007 Schneider JS, Moore DH 2nd, Mendelsohn ML (2007) Screening program reduced melanoma mortality at the Lawrence Livermore National Laboratory, 1984 to 1996. J Am Acad Dermatol 58:741–749 Schulze F, Erdmann H, Hardkop LH et al (2014) Eruptive naevi and darkening of pre-existing naevi 24 h after a single mono-dose injection of melanotan II. Eur J Dermatol 24:107–109 Scope A, Marghoob AA, Dusza SW et al (2008) Dermoscopic patterns of naevi in fifth grade children

Melanoma Prevention and Screening of the Framingham school system. Br J Dermatol 158:1041–1049 Scope A, Marghoob AA, Chen CS et al (2009) Dermoscopic patterns and subclinical melanocytic nests in normal-appearing skin. Br J Dermatol 160:1318–1321 Scope A, Dusza SW, Marghoob AA et al (2011) Clinical and dermoscopic stability and volatility of melanocytic nevi in a population-based cohort of children in Framingham school system. J Invest Dermatol 131:1615–1621 Scott JF, Das LM, Ahsanuddin S et al (2017) Oral vitamin D rapidly attenuates inflammation from sunburn: an interventional study. J Invest Dermatol 137:2078–2086 Shaikh WR, Dusza SW, Weinstock MA et al (2015) Melanoma thickness and survival trends in the United States, 1989 to 2009. J Natl Cancer Inst 108. pii: djv294 Shellman YG, Ribble D, Miller L et al (2005) Lovastatininduced apoptosis in human melanoma cell lines. Melanoma Res 15:83–89 Siddiqui IA, Bharali DJ, Nihal M et al (2014) Excellent antiproliferative and pro-apoptotic effects of (-)-epigallocatechin-3-gallate encapsulated in chitosan nanoparticles on human melanoma cell growth both in vitro and in vivo. Nanomedicine 10:1619–1626 Skaaby T, Husemoen LL, Thuesen BH et al (2014) Prospective population-based study of the association between serum 25-hydroxyvitamin-D levels and the incidence of specific types of cancer. Cancer Epidemiol Biomarkers Prev 23:1220–1229 Stoddard M, Herrmann J, Moy L et al (2017) Improvement of actinic keratoses using topical DNA repair enzymes: a randomized placebo-controlled trial. J Drugs Dermatol 16:1030–1034 Stokes RP, Diffey BL (1997) How well are sunscreen users protected? Photodermatol Photoimmunol Photomed 13:186–188 Suppa M, Altomare G, Cannavò SP, Italian investigators for the Euromelanoma prevention campaign et al (2014) The Italian Euromelanoma Day: evaluation of results and implications for future prevention campaigns. Int J Dermatol 53:699–706 Swerlick RA, Chen S (1996) The melanoma epidemic. Is increased surveillance the solution or the problem? Arch Dermatol 132:881–884 Swetter SM, Johnson TM, Miller DR et al (2009) Melanoma in middle-aged and older men: a multi-institutional survey study of factors related to tumor thickness. Arch Dermatol 145:397–404 Swetter SM, Pollitt RA, Johnson TM et al (2012) Behavioral determinants of successful early melanoma detection: role of self and physician skin examination. Cancer 118:3725–3734 Swetter SM, Chang J, Shaub AR et al (2017) Primary carebased skin cancer screening in a Veterans Affairs health care system. JAMA Dermatol 15:797–801 Swope VB, Abdel-Malek ZA (2016) Significance of the melanocortin 1 and endothelin B receptors in

569 melanocyte homeostasis and prevention of sun-induced genotoxicity. Front Genet 17(7):146 Tahata S, Singh SV, Lin Y et al (2018) Evaluation of biodistribution of sulforaphane after administration of oral broccoli sprout extract in melanoma patients with multiple atypical nevi. Cancer Prev Res 11:429–438 Tang JY, Fu T, Leblanc E et al (2011) Calcium plus vitamin D supplementation and the risk of nonmelanoma and melanoma skin cancer: post hoc analyses of the Women’s Health Initiative randomized controlled trial. J Clin Oncol 29:3078–3084 Tao S, Park SL, Rojo de la Vega M et al (2015) Systemic administration of the apocarotenoid bixin protects skin against solar UV-induced damage through activation of NRF2. Free Radic Biol Med 89:690–700 Terushkin V, Halpern AC (2009) Melanoma early detection. Hematol Oncol Clin North Am 23:481–500 Thomas L, Tranchand P, Berard F et al (1998) Semiological value of ABCDE criteria in the diagnosis of cutaneous pigmented tumors. Dermatology 197:11–17 Thompson SC, Jolley D, Marks R (1993) Reduction of solar keratoses by regular sunscreen use. N Engl J Med 329:1147–1151 Thomson K, Morley R, Grover SR et al (2004) Postnatal evaluation of vitamin D and bone health in women who were vitamin D-deficient in pregnancy, and in their infants. Med J Aust 181:486–488 Trivedi DP, Doll R, Khaw KT (2003) Effect of four monthly oral vitamin D3 (cholecalciferol) supplementation on fractures and mortality in men and women living in the community: randomised double blind controlled trial. BMJ 326:469 Tyler I, Rivers JK, Shoveller JA et al (2005) Melanoma detection in British Columbia, Canada. J Am Acad Dermatol 52:48–54 U.S. Department of Health and Human Services (2014) The Surgeon General’s call to action to prevent skin cancer. U.S. Department of Health and Human Services, Office of the Surgeon General, Washington, DC. www.surgeongeneral.gov/library/calls/pre vent-skin-cancer/call-to-action-prevent-skin-cancer. pdf. Accessed 26 June 2018 U.S. Food and Drug Administration; Department of Health and Human Services (2015) General and plastic surgery devices: restricted sale, distribution, and use of sunlamp products. A proposed rule by the Food and Drug Administration. Fed Regist 80:79493–79505 U.S. Office of Management and Budget (2017) Memorandum for the heads of executive departments and agencies; regulatory freeze pending review. Fed Regist 82:8346 U.S. Preventive Services Task Force, Bibbins-Domingo K, Grossman DC, Curry SJ et al (2016) Screening for skin cancer: US Preventive Services Task Force recommendation statement. JAMA 316:429–435 U.S. Preventive Services Task Force, Grossman DC, Curry SJ, Owens DK et al (2018) Behavioral counseling to prevent skin cancer: US Preventive Services Task

570 Force recommendation statement. JAMA 319: 1134–1142 Usher-Smith JA, Emery J, Kassianos AP (2014) Risk prediction models for melanoma: a systematic review. Cancer Epidemiol Biomarkers Prev 23:1450–1463 Vainio V, Bianchini F (eds) (2001) IARC handbooks of cancer prevention. Sunscreens, vol 5. International Agency for Research on Cancer, Lyon Vainio H, Miller AB, Bianchini F (2000) An international evaluation of the cancer-preventive potential of sunscreens. Int J Cancer 88:838–842 Valejo Coelho MM, Matos TR, Apetato M (2016) The dark side of the light: mechanisms of photocarcinogenesis. Clin Dermatol 34:563–570 van der Pols JC, Williams GM, Pandeya N et al (2006) Prolonged prevention of squamous cell carcinoma of the skin by regular sunscreen use. Cancer Epidemiol Biomarkers Prev 15:2546–2548 Vecchiato A, Pasquali S, Menin C et al (2014) Histopathological characteristics of subsequent melanomas in patients with multiple primary melanomas. J Eur Acad Dermatol Venereol 28:58–64 Vinceti M, Rothman KJ, Bergomi M et al (1998) Excess melanoma incidence in a cohort exposed to high levels of environmental selenium. Cancer Epidemiol Biomarkers Prev 7:853–856 Viros A, Sanchez-Laorden B, Pedersen M et al (2014) Ultraviolet radiation accelerates BRAF-driven melanomagenesis by targeting TP53. Nature 511:478–482 Vuong K, McGeechan K, Armstrong BK et al (2014) Risk prediction models for incident primary cutaneous melanoma: a systematic review. JAMA Dermatol 50:434–444 Waldmann A, Nolte S, Weinstock MA (2012) Skin cancer screening participation and impact on melanoma incidence in Germany – an observational study on incidence trends in regions with and without population-based screening. Br J Cancer 106:970–974 Wallingford SC, Iannacone MR, Youlden DR et al (2015) Comparison of melanoma incidence and trends among youth under 25 years in Australia and England, 1990–2010. Int J Cancer 137:2227–2233 Webster DE, Suver C, Doerr M et al (2017) The Mole Mapper Study, mobile phone skin imaging and melanoma risk data collected using ResearchKit. Sci Data 4:170005 Wehner MR, Shive ML, Chren MM et al (2012) Indoor tanning and non-melanoma skin cancer: systematic review and meta-analysis. BMJ 345:e5909 Wehner MR, Chren MM, Nameth D et al (2014) International prevalence of indoor tanning: a systematic review and meta-analysis. JAMA Dermatol 150:390–400 Weinstock MA, Rossi JS, Redding CA et al (2002) Randomized controlled community trial of the efficacy of a multicomponent stage-matched intervention to increase sun protection among beachgoers. Prev Med 35:584–592

S. M. Swetter et al. Weinstock MA, Risica PM, Martin RA et al (2007) Melanoma early detection with thorough skin selfexamination: the “Check It Out” randomized trial. Am J Prev Med 32:517–524 Weinstock MA, Ferris LK, Saul MI et al (2016) Downstream consequences of melanoma screening in a community practice setting: first results. Cancer 122:3152–3156 Welch HG, Woloshin S, Schwartz LM (2005) Skin biopsy rates and incidence of melanoma: population based ecological study. BMJ 331:481 Westerdahl J, Ingvar C, Masback A et al (2000) Sunscreen use and malignant melanoma. Int J Cancer 87:145–150 Whiteman DC, Green AC (2005) A risk prediction tool for melanoma? Cancer Epidemiol Biomarkers Prev 14:761–763 Whiteman DC, Watt P, Purdie DM et al (2003) Melanocytic nevi, solar keratoses, and divergent pathways to cutaneous melanoma. J Natl Cancer Inst 95:806–812 Whiteman DC, Baade PD, Olsen CM (2015) More people die from thin melanomas (1 mm) than from thick melanomas (>4 mm) in Queensland, Australia. J Invest Dermatol 135:1190–1193 Wolf P, Quehenberger F, Mullegger R et al (1998) Phenotypic markers, sunlight-related factors and sunscreen use in patients with cutaneous melanoma: an Austrian case-control study. Melanoma Res 8:370–378 Wu X, Fonseca M, Marchetti MA et al (2014) Longitudinally followed nevi in children and adolescents show significant size changes. J Invest Dermatol 134(Suppl 1):S49–S560 Wu H, Reeves KW, Qian J et al (2015) Coffee, tea, and melanoma risk among postmenopausal women. Eur J Cancer Prev 24:347–352 Yarosh D, Klein J, O’Connor A et al (2001) Effect of topically applied T4 endonuclease V in liposomes on skin cancer in xeroderma pigmentosum: a randomised study. Xeroderma Pigmentosum Study Group. Lancet 357:926–929 Youl PH, Soyer HP, Baade PD et al (2015) Can skin cancer prevention and early detection be improved via mobile phone text messaging? A randomised, attention control trial. Prev Med 71:50–56 Yu L, Chen H, Dou Q et al (2017) Automated melanoma recognition in dermoscopy images via very deep residual networks. IEEE Trans Med Imaging 36:994–1004 Zheng W, Doyle TJ, Kushi LH et al (1996) Tea consumption and cancer incidence in a prospective cohort study of postmenopausal women. Am J Epidemiol 144: 175–182 Zhu Y, Cheng Y, Luo RC et al (2015) Aspirin for the primary prevention of skin cancer: a meta-analysis. Oncol Lett 9:1073–1080 Zittermann A (2003) Vitamin D in preventive medicine: are we ignoring the evidence? Br J Nutr 89:552–572

Part V Management of Primary Melanoma and Locoregional Metastases

Treatment of Primary Melanomas John F. Thompson, Michael A. Henderson, Gabrielle Williams, and Merrick I. Ross

Contents Historical Perspective and the Emergence of a Contemporary Paradigm . . . . . . . . 574 Wide Excision of Primary Melanomas: Fundamental Concepts . . . . . . . . . . . . . . . . . . . T0: Melanoma In Situ/Lentigo Maligna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . T1: Invasive Melanomas 1 mm in Thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . T2: Invasive Melanomas >1–2 mm Thick . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . T3: Invasive Melanomas >2–4 mm in Thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . T4: Melanomas >4 mm in Thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

574 577 578 580 583 586

Excision Margins Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 588 Techniques for Routine Wound Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 588 Excisions for Melanomas in Unusual or Restrictive Locations . . . . . . . . . . . . . . . . . . . . 590 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 592

Abstract J. F. Thompson (*) Melanoma Institute Australia, Faculty of Medicine and Health, The University of Sydney, Sydney, NSW, Australia Department of Melanoma and Surgical Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia e-mail: [email protected] M. A. Henderson Division of Cancer Surgery, Peter MacCallum Cancer Centre, Melbourne, VIC, Australia e-mail: [email protected] G. Williams Melanoma Institute Australia, The University of Sydney, Sydney, NSW, Australia e-mail: [email protected] M. I. Ross Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_52

The great majority of patients with newly diagnosed melanomas have early-stage disease that is clinically localized to the primary site (AJCC stages I and II). Surgical strategies to treat early-stage melanomas today include two main components: wide excision of the primary melanoma (or biopsy site) and evaluation of the regional lymph node basin by sentinel lymph node biopsy. In the past, the recommended treatment of primary melanoma was aggressive surgery, involving wide resection margins of 3–5 cm around the primary melanoma and, frequently, elective radical dissection of the regional node basin. However, this strategy was not based on sound evidence, but principally on the clinical impressions of surgeons in the early twentieth century. 573

574

Although such radical management is no longer considered appropriate, surgery remains the single most effective treatment modality for clinically localized melanoma. Over time, a better understanding of the factors governing or predicting the natural history of melanoma and the results of clinical trials that were designed to study less aggressive surgical approaches have led to evidence-based recommendations for surgical excision margins. The objectives that led to this evolution in care were to provide durable local disease control and to optimize the chance of cure, while at the same time minimizing morbidity associated with the treatment. In this chapter the evidence supporting current margin recommendations for surgical excision of primary melanomas is presented and discussed, and the management of melanomas with unusual histologic findings is also considered.

Historical Perspective and the Emergence of a Contemporary Paradigm At the beginning of the twentieth century, melanoma was rarely recognized early and was usually diagnosed when it was locally advanced. Based purely on anecdotal experience, the use of 5 cm radial margins was adopted as the standard approach. The origin of this recommendation is somewhat ambiguous, although it is usually (but probably incorrectly) attributed to a report in 1907 by Handley (1907), who described a wide resection margin that included 1 inch (i.e., 2.5 cm not 5 cm) of surrounding skin, but with an additional en bloc excision of 2 inches of subcutaneous tissue and underlying fascia. Possibly supporting the use of 5 cm skin excision margins were the pathologic descriptions by Olsen and Wong who reported that atypical melanocytes may extend for up to 5 cm beyond the primary lesion (Olsen 1966; Wong 1970). Regardless of the origin of the recommendation, the 5 cm excision approach persisted until the 1970s, when a better understanding of the natural history of melanoma and the prognostic factors that influenced outcome (tumor thickness, in particular) were established (Breslow and Macht 1977; Balch et al. 1979). Surgeons then began to

J. F. Thompson et al.

use narrower margins for lower-risk (thinner) lesions and reported excellent results (Kelly et al. 1984; Cosimi 1985; Day Jr. et al. 1982). Although several published reports documented a low incidence of local recurrence when narrow margins were used for thin melanomas, several retrospective studies suggested that local recurrence was relatively frequent when such margins were used to treat thicker lesions. This indicated that local recurrence was a function of both inherent biology and the extent of the margin (Milton et al. 1985). A paradigm of wider margins being necessary for higher-risk (thicker) lesions emerged and was subsequently tested in several prospective, randomized surgical trials. If such a paradigm was true – that outcome (the frequency of local and regional recurrence and overall survival) is affected by the extent of excision margins – the following assumptions could be made: (1) microscopic satellite disease is more common and exists at a greater distance from the periphery of the primary lesion in association with thicker melanomas; (2) these microsatellites are a source of subsequent clinically-apparent local, regional, and distant relapses; and (3) wider margins remove microscopic disease that would be left behind if narrower margins were used, resulting in an improved outcome. In this chapter, evidence relating to the role of excision margins in patients with primary melanomas is summarized. All these trials and retrospective studies were designed to test the hypothesis that wider margins would be superior to narrower margins in terms of locoregional control and survival. A concise description of trial randomization schemes and the number of patients included in each is presented in Table 1.

Wide Excision of Primary Melanomas: Fundamental Concepts Wide surgical resection of the diagnostic excision biopsy site or residual primary lesion, including a surrounding margin of normal skin en bloc with the underlying subcutaneous tissue, usually down to the deep fascia, has been adopted as the surgical standard for initial management of primary melanomas. Whereas the goal of an appropriately performed diagnostic biopsy is to establish

703 (612) 112 244 148 97 9 2

2 0.5 0.51–1.0 1.1–1.5 1.51–2.0 2.1 Missing

WHO Melanoma Program (WHO)

Combined (1–4 mm) MelMarT

400 (377) 223 121 33

989 (989) 244 745

2 1–2

Swedish Melanoma Study Group I (SMSGI)

>1 1–2 2–4 >4

337 (326) 18 141 105 61 1

2 0.5 0.5–1.0 1.01–1.5 1.5 Missing

Surgical trials Thin melanoma French Cooperative Study (French)

Number of randomized patients (N analyzed)

Breslow thickness (mm)

1 versus 2 (185 vs. 192)

1 versus 3 (305 vs. 307)

2 versus 5 (476 vs. 513)

2 versus 5 (167 vs. 170)

Treatment comparisons, in cm (N patients)

Trunk, extremities, head, and neck

Trunk, limbs

Trunk, limbs

Extremities, trunk, head, and neck

Primary site

1 (>1)

4, 8, 12 (>8)

5 and 10 (11)

5 and 10 (16)

Outcome time points, years (average follow-up)

Table 1 Prospective randomized trials addressing surgical excision margins for primary cutaneous melanoma

None reported

Local: 0.26 (0.03, 2.27) Nodal: 1.21 (0.56, 2.62) Distant: 0.41 (0.13, 1.28) Death: 1.13 (0.72, 1.78) Local: 0.65 (0.16, 2.69) Nodal: 1.24 (0.90, 1.70) Distant: 1.08 (0.79, 1.46) Death: 0.94 (0.76, 1.17) Local: 2.68 (0.72, 10.02) Nodal: 0.88 (0.50, 1.55) Distant: 1.22 (0.61, 2.44) Death: 0.85 (0.57, 1.27)

Risk ratios (95%CIs) (narrow: wide)

(continued)

Moncrieff et al. (2018)

Veronesi et al. (1988) Veronesi and Cascinelli (1991) Cascinelli (1998)

Ringborg et al. (1996) Cohn-Cedermark et al. (2000)

Khayat et al. (2003) Banzet et al. (1993)

References

Treatment of Primary Melanomas 575

900 (900) 2 99 555 242 2

936 (936) 460 204 270 2

>2 3 >3–4 >4 Missing

Number of randomized patients (N analyzed) 486 (470/468) 272 141 73

>2 4 Missing

Breslow thickness (mm) 1–4 1–1.99 2–2.99 3–4.0

2 versus 4 (465 vs. 471)

1 versus 3 (453 vs. 447)

Treatment comparisons, in cm (N patients) 2 versus 4 (244 vs. 242)

Trunk, limbs

Trunk, limbs

Primary site Trunk, proximal limb

5 and 10

5 and 10

Outcome time points, years (average follow-up) 5 and 10 (10)

Local: 1.14 (0.55, 2.37) In-transit/ regional: 1.41 (0.54, 3.67) Nodal: 1.13 (0.92, 1.39) Distant: 1.25 (0.79, 1.98) Death: 1.04 (0.92, 1.17) Local: 2.25 (1.04, 4.89) In-transit/ regional: 1.28 (0.66, 2.49) Nodal: 0.89 (0.70, 1.13) Distant: 0.71 (0.48, 1.06) Death: 1.04 (0.88, 1.22)

Risk ratios (95%CIs) (narrow: wide) Local: 0.83 (0.26, 2.67) Nodal: 0.96 (0.60, 1.53) Distant: 1.21 (0.87, 1.67) Death: 1.29 (0.95, 1.76)

Gillgren et al. (2011)

Thomas et al. (2004) Hayes et al. (2016)

References Balch et al. (1993) Karakousis et al. (1996) Balch et al. (2001)

Local and regional node recurrences were combined (this was not planned in the original trial protocol) and patients did not have regional node staging by SNB or ELND, so the groups may not have been balanced

a

Swedish Melanoma Study Group II (SMSGII)

Thick melanoma UK Melanoma Study Group (UKMSG)a

Surgical trials Intergroup Melanoma Surgical trial (Intergroup)

Table 1 (continued)

576 J. F. Thompson et al.

Treatment of Primary Melanomas

pathologic microstaging (primarily tumor thickness and ulceration status), the goal of wide local excision is to achieve durable local disease control and to cure those patients who do not already harbor disseminated micrometastases in regional lymph nodes or at distant sites. The optimal width of an excision margin has been a matter of controversy for decades. The vast majority of “narrow” excision procedures (with 1 cm surgical margins) allow primary closure of the resultant surgical defect and can be accomplished using local anesthesia, at minimal cost and with little or no morbidity (Bono et al. 1997). On the other hand, “wide” excision margins (>2 cm) are often performed with the patient under general anesthesia and, depending on the anatomic site, may require skin grafting or reconstruction with sometimes complex full-thickness skin advancement or rotation flaps. However, narrow excision margins for higher-risk lesions are often thought to result in a higher rate of locoregional recurrence, which may impact on survival. Because local recurrence is generally associated with a poor prognosis and often requires complicated and potentially morbid treatment, prevention of such events has been a compelling reason to recommend wide excision margins. At the same time, minimizing treatment-related morbidity is another important management goal. Use of excessive margins with no improvement in outcome can lead to unnecessary morbidity, additional cost, more complications, and permanent disfigurement. Therefore, establishing rational standards for the extent of surgical excisions has been a focus of extensive clinical investigation. When considering excision margins, the fundamental difference between surgical margins and histologic margins must be borne in mind. All current treatment recommendations refer to surgical margins, i.e. the in vivo margins determined before wide excision and before histologic examination of the specimen. It must also be borne in mind that there is a predictable degree of tissue shrinkage after the specimen has been removed and fixed in formalin. A detailed study by Friedman et al. (2019) showed that there was usually a 14% shrinkage rate, so that the macroscopic margins measured by the pathologist at the time of specimen cut-up will

577

be less than the margins marked out by the surgeon prior to the wide excision.

T0: Melanoma In Situ/Lentigo Maligna Melanoma in situ is the earliest recognizable stage of melanoma. It is noninvasive, but may be a precursor to invasive melanoma, which has the potential for metastasis. Lentigo maligna is a subtype of melanoma in situ associated with chronic exposure to ultraviolet radiation. These lesions are insidiously extensive microscopically and, despite the removal of normal-appearing skin surrounding the pigmented lesion, histologically positive margins may be encountered. Lentigo maligna lesions can require sequential operative excisions to achieve adequate, pathologically clear margins, and occasionally the use of a splitthickness skin graft or flap closure is warranted. There have been no randomized trials and only a limited number of case series to guide excision margins for melanoma in situ (Tzellos et al. 2014; Cancer Council Australia Melanoma Guidelines Working Party 2018) and treatment therefore is based on consensus from clinical experience. Until recently, most guidelines suggested that 5 mm excision margins were adequate for melanoma in situ. However two case series (Bosbous et al. 2009; Akhtar et al. 2014) demonstrated that 5 mm may be inadequate, leading to disease recurrence in some cases. The risk of recurrence of melanoma in situ was particularly great for the lentigo maligna subtype and for head and neck sites. In many cases, adequate clearance margins for melanoma in situ can be accurately determined preoperatively by careful examination and confirmed by pathology. However, defining the extent of the lesion can be difficult in lentigo maligna because it is characterized by the subtle presence of scattered atypical melanocytes, meaning positive margins can remain after excision, leading to possible recurrence at excision sites (Akhtar et al. 2014; Bosbous et al. 2009). A recent study reported that the use of in vivo confocal microscopy may improve lesion boundary identification and allow for more effective excision (Guitera et al. 2013). However, further studies are required to validate this technique.

578

Current Excision Margin Recommendations for T0 Melanomas European guidelines suggest 5 mm excision margins for melanoma in situ (Dummer et al. 2012; Garbe et al. 2016; NICE 2015) although a British Best Practice statement (BMJ Best Practice 2016) includes the caveat that 5 mm is inadequate in up to 50% of cases of melanoma in situ, particularly lentigo maligna, and 10 mm margins, staged excision or Mohs surgery are options. Dermatology guidelines from the USA recommend 5–10 mm margins and state that wider margins may be necessary for the lentigo maligna subtype of melanoma in situ (Bichakjian et al. 2011). Table 2 details the recommendations from these and various other national guidelines. Based on all the available evidence, it would seem most appropriate to widely excise melanoma in situ with a minimum margin of 5 mm, but with a margin of 10 mm when this is readily possible.

T1: Invasive Melanomas ≤1 mm in Thickness Randomized Trials of Excision Margins for T1 Melanomas No randomized trials have exclusively included only patients with primary tumors 1 mm thick (often referred to as “thin” melanomas). Three trials included 159/326 (49%), 356/612 (58%),

J. F. Thompson et al.

and 244/989 (25%) patients with primary tumors 5 to 10 mm in Breslow thickness (Khayat et al. 2003; Veronesi and Cascinelli 1991; Ringborg et al. 1996). Two trials compared 2 cm with 5 cm margins (Khayat et al. 2003; Ringborg et al. 1996), while the WHO trial compared 1 cm with 3 cm margins (Veronesi et al. 1988). Individually, the trials reported nonsignificant differences in the risk of death, local recurrence, nodal recurrence, or distant metastasis between the narrow (1 or 2 cm) and wider (3, 4, or 5 cm) excision margins (Table 1). In combining these trials, the pooled estimate for risk of death associated with narrower versus wider margins was 0.95 (95%CI 0.76–1.17), risk of local recurrence 0.92 (95%CI 0.72–3.37), risk of nodal recurrence 1.15 (95%CI 0.88–1.49), and risk of distant metastasis 1.00 (95%CI 0.66–1.50) (Fig. 1a–c). Thus, all estimates showed no significant differences when patients were treated with 1 or 2 cm margins compared to 3, 4, or 5 cm margins. Two of these trials (Cascinelli 1998; Khayat et al. 2003) reported local recurrence within the group of patients with primary tumors 1 mm, and when combined gave a risk ratio of 1.30 (95%CI 0.18–9.70) but with only five events, the risk estimate ranged from 82% less likely to almost 10 times more likely that a local recurrence would occur in a patient treated with a narrow (1–2 cm) compared to a wide (3–5 cm) margin (Fig. 1e). These trials do not report outcomes of death,

Table 2 National guidelines for excision margins for T0 (in situ) cutaneous melanoma Guideline organization, country NICE, United Kingdom NCCN, USA

Year 2015 2017

Margin (mm) 5 5–10

Cancer Council, Australia

2018

5–10

National Melanoma Tumour Standards Working Group

2013

5–10

Cancer Care Ontario, Canada Dutch Working Group on Melanoma, Netherlands

2017 2013

5–10 5

German Dermatological Society (DDG) and the Dermatologic Cooperative Oncology Group (DeCOG), Germany

2013

5

Reference NICE (2015) NCCN Clinical Practice Guidelines in Oncology: Melanoma (2018) Cancer Council Australia Melanoma Guidelines Working Party National Melanoma Tumour Standards Working Group (2013) Wright et al. (2017) Dutch Working Group on Melanoma (2013) Pflugfelder et al. (2013)

Treatment of Primary Melanomas

579

Fig. 1 (a) Risk of death from all causes in trials of patients with primary tumors 2 mm thick randomized to narrow (1–2 cm) margins or wide (3–5 cm) margins. (b) Risk of local recurrence in trials of patients with primary tumors 2 mm thick randomized to narrow (1–2 cm) margins or wide (3–5 cm) margins. (c) Risk of nodal recurrence in trials of patients with primary tumors 2 mm thick

randomized to narrow (1–2 cm) margins or wide (3–5 cm) margins. (d) Risk of distant metastases in trials of patients with primary tumors 2 mm thick randomized to narrow (1–2 cm) margins or wide (3–5 cm) margins. (e) Risk of local recurrence in the subgroup of patients with primary tumors 1 mm thick randomized to narrow (1 or 2 cm) margins or wide (3 or 5 cm) margins

nodal metastasis or distant metastasis within the group of patients with primary tumors 1 mm thick; therefore, it is not possible to determine the applicability of these findings to this subgroup.

Nonrandomized Studies of Excision Margins for T1 Melanomas In a case control study, 174 patients with primary melanomas 1 mm thick who experienced a recurrence of disease were matched with patients who

580

J. F. Thompson et al.

did not have a recurrence; these two groups were selected from a patient cohort of 10,505 and were analyzed for associations with recurrence. (MacKenzie Ross et al. 2016). Excision margin was associated with time to recurrence with a hazard ratio of 0.95 (95%CI 0.92–0.98), meaning that for every 1 mm decrease in excision margin, the patient had a 5% reduction in time to recurrence. Categorizing the margin into 2 mm thickness and randomized to narrow (1–2 cm) or wide (3–4 cm) margins of excision. (b) Risk of in-transit or regional recurrence in trials of patients with primary melanoma tumors of >2 mm thickness and randomized to narrow (1–2 cm) or wide (3–4 cm) margins of excision. (c) Risk of nodal recurrence

587

no impact of excision margin on survival, but the authors acknowledged the lack of power for this analysis. An earlier study of 278 patients with melanomas >4 mm thick showed that neither local recurrence nor survival was associated with excision margins of 2 cm (Heaton et al. 1998). A similar study of 108 patients with primary tumors >4 mm also showed no reduction in local recurrence or difference in survival with excision margins of 2 mm (Ruskin et al. 2016). The inference, albeit limited by the retrospective nature of these two 10-year studies, is that

in trials of patients with primary melanoma tumors of >2 mm thickness and randomized to narrow (1–2 cm) or wide (3–4 cm) margins of excision. (d) Risk of distant metastases in trials of patients with primary melanoma tumors of >2 mm thickness and randomized to narrow (1–2 cm) or wide (3–4 cm) margins of excision

588

J. F. Thompson et al.

2 cm margins can safely be used for melanomas >4 mm thick, thereby avoiding the potential morbidity, cost, and cosmetic disfigurement associated with margins greater than 2 cm.

Current Excision Margin Recommendations for T4 Melanomas Since randomized trial data and retrospective analyses show no benefit for excision with wide margins of 3, 4, or 5 cm compared with 2 cm for patients with melanomas >4 mm thick, most national guidelines suggest an excision margin of 2 cm for melanomas >4 mm thick (Table 7). Table 7 National guidelines for excision margins (in cm) for T4 cutaneous melanoma (>4 mm thick) Guideline organization, country National Institute for Health and Care Excellence, United Kingdom National Comprehensive Cancer Network, USA

Year 2015

Margin (cm) 3

2017

2

Cancer Council Australia, Australia and New Zealand

2018

2

Cancer Care Ontario, Canada Dutch Working Group on Melanoma, Netherlands German Dermatological Society (DDG) and the Dermatologic Cooperative Oncology Group (DeCOG), Germany

2017

2

2013

2

2013

2

Reference NICE (2015)

NCCN Clinical Practice Guidelines in Oncology: Melanoma (2018) Cancer Council Australia Melanoma Guidelines Working Party (2018) Wright et al. (2017) Dutch Working Group on Melanoma (2013) Pflugfelder et al. (2013)

Excision Margins Summary The results of randomized and nonrandomized studies detailed above provide the basis for the current recommendations about the width of surgical margins, based on primary melanoma thickness. Simply stated, for melanomas 2 mm, 1 cm and 2 cm margins, respectively, appear to be adequate, with still some lingering uncertainty about the appropriate margin in the 1–2 mm subgroup. No data exist that margins wider than 2 cm (3 cm, 4 cm, or 5 cm) result in any superior disease-specific outcome, but these wider margins are associated with increased surgical morbidity. However, excision margins 3 cm) positive inguinofemoral node or a positive Cloquet’s

680

K. A. Delman et al.

Fig. 9 Melanoma patient with clinically positive iliac/obturator node on PET scan (large arrow). (From Lesly A. Dossett, Ann Arbor, Michigan; with permission)

node has been considered relative indications for iliac/obturator lymph node dissection, but limited data exist to support these factors as accurate predictors of pelvic nodal disease. Additionally, patients meeting these criteria generally already have an indication for iliac/obturator dissection based on the criteria above. The majority of the data on Cloquet’s node as a predictor of pelvic nodal disease was in the era prior to SLNB. In the absence of SLNB, the sensitivity and specificity of Cloquet’s node in predicting pelvic disease are highly variable. In a series from the Netherlands Cancer Institute, Cloquet’s node and the number of positive nodes were evaluated as possible factors to predict positive deep nodes. The sensitivity and negative predictive value of a positive Cloquet’s node were 55% and 78%, respectively. Use of more than three positive nodes in the inguinofemoral dissection as a predictor revealed a sensitivity of 41% and a negative predictive value of 78%. Combining the two variables resulted in a sensitivity of 56% and a negative predictive value of 82% (Strobbe et al. 2001). The authors concluded that the sensitivity of Cloquet’s node is too low to recommend routine sampling as a predictor of iliac/obturator nodal involvement. In contrast, investigators at the John Wayne Cancer Institute have concluded that it is possible to predict positive iliac/obturator nodes by assessment of Cloquet’s node (Shen et al. 2000; Essner et al.

2006). With routine histology of Cloquet’s node, the investigators were able to achieve a sensitivity of 82%, a positive predictive value of 70%, and a negative predictive value of 84%. They concluded that Cloquet’s node assumes the role of a SLN for the iliac/obturator nodes in patients with positive inguinofemoral nodes (Shen et al. 2000; Essner et al. 2006). Routine biopsy of Cloquet’s node in SLNB patients is of low value and not recommended(Chu et al. 2010). Although likely of limited value in the current era, in the setting of microscopic inguinofemoral disease, there are two generally accepted indications for iliac/obturator node dissection: (1) a positive iliac/obturator SLN and (2) an iliac/obturator SLN identified on preoperative lymphoscintigraphy, but not sampled/removed, in the setting of a positive inguinofemoral SLN. Of note, in one study assessing the true frequency of synchronous iliac/obturator nodal metastases with microscopic inguinofemoral disease, the authors found the prevalence of synchronous disease to be 11.9%. Patients with iliac/obturator disease were more likely to have a ratio of total positive inguinal nodes to total retrieved inguinal nodes greater than 0.20 or 3 total involved inguinal nodes (Chu et al. 2011). Iliac/obturator dissection is also performed in patients who have recurrent melanoma of the extremity and are offered limb perfusion after they have already undergone an inguinofemoral lymphadenectomy.

Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma

Operative Technique Skin Incision. Usually an iliac/obturator dissection is performed simultaneously with an inguinofemoral dissection, as described earlier. Under these circumstances, the incision can be made by extending the inguinofemoral incision cephalad, raising the superior flap to reveal the lower abdominal musculature. When using a transverse infrainguinal incision for inguinofemoral dissection, a separate incision may be performed for the iliac/obturator dissection in a similar fashion to the retroperitoneal incision used for kidney transplantation. Another alternative is a lower midline incision from the umbilicus to the pubic symphysis to access the iliac/obturator nodes via an extraperitoneal approach. Abdominal Wall Incision. Access to the iliac fossa is gained via a transverse incision through the external oblique aponeurosis approximately 5–6 cm above the inguinal ligament (i.e., above the inguinal canal). The internal oblique and the transversus abdominis muscle are then split in the direction of their fibers. The incision is continued to include the lateral sheath of the rectus abdominis muscle. The deep circumflex artery and vein are identified and ligated where they lie between the internal oblique and transversus abdominis muscles. The preperitoneum is entered and freed from the abdominal wall in the distal direction. As the retroperitoneum is entered, the peritoneum is bluntly lifted up out of the iliac fossa, and the inferior epigastric vessels are identified, passing upward and medially from the point where the femoral arterial pulsations can be felt. Once encountered, the retroperitoneal space is developed by bluntly dissecting and retracting the peritoneum in a superior and medial direction. A self-retaining retractor can be used to hold the peritoneal contents off the pelvic brim. The inferior epigastric vessels are identified as they come off the distal aspect of the external iliac vessels, and they are frequently ligated and divided at their origin. The vas deferens will be seen at this stage and should be preserved, as should the testicular vessels that lift off the iliac fossa with

681

the peritoneum. The round ligament can be sacrificed in women. The use of an in-continuity dissection, described by Karakousis and Driscoll (Karakousis and Driscoll 1994), provides for an en bloc resection and can offer superior exposure, especially to the most distal external iliac nodes. This approach involves the division of the external oblique aponeurosis in a nearly vertical fashion approximately 4 cm medial to the ASIS. Inferiorly, the inguinal ligament is divided just medial to the femoral artery. This line of dissection courses through the internal oblique and transversalis fascia and allows for a more straightforward fascia-to-fascia closure compared with performing the incision precisely at the ASIS. Division of the inguinal ligament is associated with remarkable morbidity, particularly postoperative pain, and should not be undertaken lightly. The utility of an en bloc resection is questionable, but, if strongly desired as an alternative to the division of the inguinal ligament, some surgeons favor a separate obliquely oriented incision in the groin crease. This allows for the inguinal ligament to be left intact, which may lower the incidence of postoperative abdominal wall weakness or hernia formation. Iliac/Obturator Node Dissection. Dissection around the external iliac vessels is usually performed with sharp instruments up to the level of the ureter, crossing the bifurcation of the common iliac artery, or higher if clinically detectable nodes are involved along the iliac vessels. Though as noted, common iliac nodes are generally considered stage IV disease. The ureter is identified and preserved as it courses over the iliac artery. The lymphatic contents are dissected medially off the bladder wall and superiorly off the posterior rectus sheath. Beginning at the inguinal ligament inferiorly, the lymph nodes are dissected off the artery and vein, working within the vessel sheath. Small vessels and lymphatics at the perimeter of the excision should be ligated, cauterized, or clipped to avoid hemorrhage or lymphoceles. The dissection then continues superiorly up to the common iliac vessels.

682

The deeper portion of the dissection is made possible by reflecting the peritoneum medially with a broad self-retaining retractor and then working downward, using fingers or sponge sticks as blunt dissectors. The advantage of this method is that the obturator nerve, which is very close to the major lymph nodes, can be felt as a taut cord that moves away from the sidewall of the pelvis. If this part of the dissection is performed with a sharp instrument, there is some risk of damaging the nerve. The medial dissection (to remove the iliac nodes) and deep dissection (to remove the obturator nodes) are completed when the obturator nerve is identified and preserved as it courses from the lateral aspect of the internal iliac artery toward the obturator foramen. The obturator nodes are carefully resected from this area, and the specimen is removed. Interestingly, most surgeons dissect the deep aspect of this procedure along the obturator nerve using digital dissection, as a finger will prevent significant damage and is sensitive enough to alert the surgeon to structures that should be avoided. As the specimen containing the lymph nodes is lifted upward in one unit with the external iliac nodes, abdominal packs are placed firmly in the pelvis and left until the operation is complete, by which time minor venous bleeding will have stopped. After the wound is irrigated and meticulous hemostasis is achieved, the transversalis and internal oblique muscles are approximated with nonabsorbable sutures, and the external oblique aponeurosis is then approximated with nonabsorbable sutures. To obliterate the enlarged femoral canal defect, the inguinal ligament is approximated to the lacunar ligament with a figure-of-eight suture of nonabsorbable material (unless this has been closed as previously described during the inguinal part of a combined procedure). A closed suction drain may be placed in the retroperitoneal space through a separate stab wound, although this is optional. Skin closure is accomplished and the dressings applied in the manner described for inguinofemoral dissection. Postoperative Care. Patients are kept on bed rest with the operated extremity elevated overnight. If placed intraoperatively, the urinary catheter is removed on the first postoperative morning,

K. A. Delman et al.

and the patient is allowed to ambulate with assistance as needed. Most patients are discharged on the first postoperative morning, similar to those undergoing inguinofemoral dissection. Patients are instructed to keep the operated extremity elevated when they are not ambulating. Drains in the pelvis should be removed when draining less than 30 mL per day for two consecutive days.

Modifications of the Classic Technique of Iliac/Obturator Lymphadenectomy As with inguinofemoral lymphadenectomy, a minimally invasive technique, referred to as the robotic-assisted transperitoneal pelvic lymphadenectomy (rPLD), has emerged as a safe and effective technique for iliac/obturator lymphadenectomy. The advantages of minimally invasive rPLD are well-described in the literature where it is routinely used for staging and treatment of urologic and gynecologic malignancies. rPLD has been shown to improve visualization of the iliac and obturator nodes and provide equivalent nodal yield and shorter length of stay when compared to the classic technique of iliac/obturator lymphadenectomy (Dossett et al. 2016).

Robotic-Assisted Transperitoneal Pelvic Lymphadenectomy When utilizing the minimally invasive approach, preparation for conversion to open and for inguinofemoral dissection if being performed in a combined procedure should be made as described above. The patient is placed in steep Trendelenburg position, and the robot is docked between the legs of the patient. A 12-mm midline port is typically placed 18–20 cm above the pubic symphysis. Two 8-mm robotic ports are placed laterally on the contralateral side of the abdomen from the site of pelvic dissection, at least 14 cm from the pubic symphysis, and at least 6 cm away from other ports. A 12-mm assistant port is placed in the midclavicular line on the affected side of the abdomen. A third 8-mm robotic port is placed on the affected side of the abdomen at least 8 cm

Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma

lateral to the assistant port and at least 2 cm medial to the anterior superior iliac spine (Fig. 10). In the retroperitoneum, the psoas muscle, ureter, and common iliac bifurcation are identified. The lymph node packet overlying and between the external iliac artery and vein is dissected from underneath the inguinal ligament to the iliac bifurcation or higher if indicated. After removal of the iliac nodal packet, resection of the obturator lymph nodes is performed. The iliac vein is retracted laterally, and the obturator nodes are carefully dissected free of their attachments, identifying and preserving the obturator vessels and nerve. Each lymph node packet is placed in an endoscopic retrieval bag and removed through the midline port. For cases where rPLD is combined with a superficial inguinal lymphadenectomy, the rPLD precedes the inguinal lymphadenectomy to avoid leakage of CO2 out of the operative field. Of note, it is possible to achieve similar outcomes with a strictly laparoscopic approach without the costs associated with robotic technology. The robotic approach, however, has the advantage of three-dimensional visualization; ergonomic, intuitive control; and wristed instruments that

Fig. 10 Port placement for a left robotic assisted transperitoneal pelvic lymphadenectomy. There are three 8 mm robotic ports seen at the lateral aspects of the abdomen, a 12 mm camera port in the center of the abdomen and a 12 mm assistant port in the left abdomen. (From Dossett et al. 2016)

683

approximate the motion of the human hand, which are all advantageous, given the extensive dissection in close proximity to iliac vessels and the obturator nerve.

Postoperative Complications: Incidence and Risk Factors Complications of Lymph Node Dissection The morbidity after an inguinofemoral lymph node dissection is greater than that associated with axillary or cervical lymphadenectomy, particularly in the older population and in patients who are obese (Beitsch and Balch 1992). The rate of postoperative complications following inguinofemoral lymph node dissection has been estimated as high as 75% (Chang et al. 2010; Stuiver et al. 2014). Early perioperative complications include infection, hemorrhage, seroma formation, skin edge necrosis, wound dehiscence, and lymphocele. Long-term complications include paresthesias and chronic lymphedema. One explanation for the short-term morbidity associated with inguinal lymphadenectomy is that the main blood supply to the skin overlying the femoral triangle comes from a series of small anterior branches of the femoral artery. These small arteries pass directly through the cluster of femoral lymph nodes. Any complete excision of these nodes divides most, if not all, of these vessels. It follows that the vitality of the skin close to the incision and for a distance of up to 10 cm below the inguinal ligament may be impaired regardless of where the incision for lymphadenectomy is placed. The reduced blood supply to the skin may result in either slow wound healing or complete breakdown of this part of the wound, particularly in older and/or obese patients, though some of these wound complications can be minimized by excising the skin over the femoral triangle. Lymphedema is the most significant long-term complication following lower extremity node dissection. Clinical rates of lower extremity swelling following inguinofemoral dissection are reported

684

to range from 20% to 64% in most series (Chang et al. 2010). Lymphedema is generally reported as being mild, although criteria for the definition of lymphedema vary among authors. It is possible that a significant percentage of patients actively complying with a regimen of long-term elastic support stockings and/or sequential compression pump therapy have lymphedema but are successfully treated as opposed to patients who are truly free of lymphedema. Importantly, the addition of an iliac/obturator lymph node dissection with an inguinofemoral dissection has historically been associated with an increased risk of lymphedema; however, more recent data suggests the addition of an iliac/obturator lymph node dissection does not significantly increase this risk (Chang et al. 2010).

Management of Postoperative Complications Meticulous attention to surgical technique and hemostasis should help prevent a significant number of postoperative complications. Avoiding the creation of nonviable skin flaps or intraoperative wound contamination will contribute significantly to the prevention of postoperative wound edge necrosis, dehiscence, and infection. When these complications do occur, they are best managed with appropriate local wound care. Debridement should be performed aggressively to remove any and all nonviable tissue. Tissues of questioned viability, however, should be observed while they are treated with topical antibiotics. Systemic antibiotics should be reserved only for evidence of invasive soft tissue infection or cellulitis. The open wound should be packed and the dressing changed two to three times per day until there is healthy granulation tissue in the defect. Lymphocele with seroma formation is a common complication, with incidence ranging from 5% to 27% (Badgwell et al. 2007). Lymphoceles large enough to cause the overlying skin flaps to become firm or tense should be treated by prompt aspiration under sterile conditions. Aspiration attempts should be

K. A. Delman et al.

made as far away from the incision as possible. A lymph collection that rapidly re-accumulates after repeated aspirations should be considered for percutaneous suction catheter drainage rather than continuing to perform repeated serial aspirations, each of which incurs the risk of infecting the lymphocele (Hoffman et al. 1995). Lymphedema clearly represents the most serious nonmalignant long-term complication resulting from lower extremity node dissection. Prevention of lymphedema begins in the operating room by taking steps to prevent perioperative wound infection. Wound infection results in increased fibrosis of the soft tissues of the femoral area and thereby likely results in the obliteration of microscopic lymphatic vessels. Lymphedema itself predisposes patients to infection of the extremity, particularly cellulitis. A vicious cycle of infection resulting in worsening of lymphedema followed by further infections can thereby be initiated. Thus, prevention of lymphedema and perioperative infection goes hand in hand. While lymphedema surveillance and prevention protocols vary on an institutional level, all patients should receive education emphasizing the importance of early detection. At many institutions, patients are measured preoperatively for custom-fitted, medium compression (20–30 mmHg) elastic garments, so that they can be worn as soon as possible and for up to 6 months postoperatively. At other institutions, a compression garment is only used in the treatment of established lymphedema. Additionally, sequential compression devices have become an important part of the treatment armamentarium in combating significant established lymphedema. Devices are custom fit for each patient with lymphedema and are typically worn for periods of up to 1–4 h daily while the patient is at home. Sequential compression devices function by mimicking the natural pumping action of the lower extremity musculature during ambulation, which propels the protein-rich edema fluid out of the soft tissues in a cephalad direction.

Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma

Popliteal Dissection Indications Because the incidence of popliteal lymph node involvement is exceedingly rare in patients with melanoma, Thompson et al. (2000) recommends popliteal lymph node dissection only for clinical evidence of metastatic disease in a popliteal node.

Operative Technique To prepare for dissection, the patient is placed in the prone position with the operative leg slightly flexed at the knee. The patient is prepped from the mid-thigh to the inferior aspect of the gastrocnemius muscle. A lazy-S incision is made over the flexor crease to (1) allow optimal exposure and (2) heal in a manner that does not cause a deforming joint contracture (Sholar et al. 2005). The incision should begin approximately 10 cm above the joint crease of the lateral thigh and then move transversely across the joint. From there, it extends longitudinally along the medial aspect of the leg for approximately 10 cm (Fig. 11). After the incision is created and carried down through the subcutaneous tissue, lateral and medial flaps are raised, while traction is maintained with skin Fig. 11 Technique of popliteal node dissection incision and exposure of superficial structures. (From Advanced Therapy in Surgical Oncology, Pollock RE, Curley SA, Ross MI, eds.; used with permission from PMPH USA, Ltd., Raleigh, NC)

685

hooks. As the fascia is exposed, the most superficial structures that come into view are the lesser saphenous vein and some small cutaneous nerve terminal branches (Fig. 11). At this point the lesser saphenous vein must be ligated and divided. Next the deep fascia is incised vertically, taking care to avoid damaging structures below the fascia, because the nerves are quite superficial (Fig. 12). If the medial sural nerve can be retracted out of the way, it should be. However, if necessary, it can be divided to gain better access to deeper structures; this will result in cutaneous anesthesia. The tibial nerve is the most superficial midline structure. This is very gently retracted laterally with a vessel loop. Similarly, the peroneal nerve courses along the biceps femoris and semimembranosus muscles. Inferiorly the two heads of the gastrocnemius muscle can be further retracted as well to enhance distal exposure. The node-bearing tissue is swept from around the nerves and moved distally to expose the popliteal artery and vein. Nodal tissue surrounding these vessels should be dissected free, making sure to include any tissue lying on the far side of the vessels as well (Fig. 12). The dissection is continued until these vessels dive behind the gastrocnemius muscle. Once the specimen is removed, a drain is placed and the wound is closed in layers.

686

K. A. Delman et al.

Fig. 12 Technique of popliteal node dissection exposure of vessels and deeper structures. (From Advanced Therapy in Surgical Oncology, Pollock RE, Curley SA, Ross MI, eds.; used with permission from PMPH USA, Ltd., Raleigh, NC)

Generally, the patient remains hospitalized at least overnight and longer if additional nodal dissections were performed. The drain is left in place until drainage is less than 30 mL/day.

Cross-References ▶ Biopsy of the Sentinel Lymph Node ▶ Hyperthermic Regional Perfusion for Melanoma of the Limbs ▶ Lymphoscintigraphy in Patients with Melanoma ▶ Molecularly Targeted Therapy for Patients with BRAF Wild-Type Melanoma ▶ Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma ▶ Targeted Therapies for BRAF-Mutant Metastatic Melanoma

References Abbas S, Seitz M (2011) Systematic review and metaanalysis of the used surgical techniques to reduce leg lymphedema following radical inguinal nodes dissection. Surg Oncol 20:88–96 Badgwell B, Xing Y, Gershenwald JE, Lee JE, Mansfield PF, Ross MI, Cormier JN (2007) Pelvic lymph node dissection is beneficial in subsets of patients with node-positive melanoma. Ann Surg Oncol 14:2867–2875

Beitsch P, Balch C (1992) Operative morbidity and risk factor assessment in melanoma patients undergoing inguinal lymph node dissection. Am J Surg 164: 462–465; discussion 465–6 Chang SB, Askew RL, Xing Y, Weaver S, Gershenwald JE, Lee JE, Royal R, Lucci A, Ross MI, Cormier JN (2010) Prospective assessment of postoperative complications and associated costs following inguinal lymph node dissection (Ilnd) in melanoma patients. Ann Surg Oncol 17:2764–2772 Chu CK, Zager JS, Marzban SS, Gimbel MI, Murray DR, Hestley AC, Messina JL, Sondak VK, Carlson GW, Delman KA (2010) Routine biopsy of Cloquet’s node is of limited value in sentinel node positive melanoma patients. J Surg Oncol 102:315–320 Chu CK, Delman KA, Carlson GW, Hestley AC, Murray DR (2011) Inguinopelvic lymphadenectomy following positive inguinal sentinel lymph node biopsy in melanoma: true frequency of synchronous pelvic metastases. Ann Surg Oncol 18:3309–3315 Delman KA, Kooby DA, Ogan K, Hsiao W, Master V (2010) Feasibility of a novel approach to inguinal lymphadenectomy: minimally invasive groin dissection for melanoma. Ann Surg Oncol 17:731–737 Delman KA, Kooby DA, Rizzo M, Ogan K, Master V (2011) Initial experience with videoscopic inguinal lymphadenectomy. Ann Surg Oncol 18:977–982 Dossett LA, Castner NB, Pow-Sang JM, Abbott AM, Sondak VK, Sarnaik AA, Zager JS (2016) Roboticassisted transperitoneal pelvic lymphadenectomy for metastatic melanoma: early outcomes compared with open pelvic lymphadenectomy. J Am Coll Surg 222:702–709 Essner R, Scheri R, Kavanagh M, Torisu-Itakura H, Wanek LA, Morton DL (2006) Surgical management

Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma of the groin lymph nodes in melanoma in the era of sentinel lymph node dissection. Arch Surg 141: 877–882; discussion 882–4 Faries MB, Thompson JF, Cochran AJ, Andtbacka RH, Mozzillo N, Zager JS, Jahkola T, Bowles TL, Testori A, Beitsch PD, Hoekstra HJ, Moncrieff M, Ingvar C, Wouters M, Sabel MS, Levine EA, Agnese D, Henderson M, Dummer R, Rossi CR, Neves RI, Trocha SD, Wright F, Byrd DR, Matter M, Hsueh E, Mackenzie-Ross A, Johnson DB, Terheyden P, Berger AC, Huston TL, Wayne JD, Smithers BM, Neuman HB, Schneebaum S, Gershenwald JE, Ariyan CE, Desai DC, Jacobs L, Mcmasters KM, Gesierich A, Hersey P, Bines SD, Kane JM, Barth RJ, Mckinnon G, Farma JM, Schultz E, Vidal-Sicart S, Hoefer RA, Lewis JM, Scheri R, Kelley MC, Nieweg OE, Noyes RD, Hoon DSB, Wang HJ, Elashoff DA, Elashoff RM (2017) Completion dissection or observation for sentinel-node metastasis in melanoma. N Engl J Med 376:2211–2222 Hoffman MS, Mark JE, Cavanagh D (1995) A management scheme for postoperative groin lymphocysts. Gynecol Oncol 56:262–265 Karakousis CP, Driscoll DL (1994) Groin dissection in malignant melanoma. Br J Surg 81:1771–1774 Leiter U, Stadler R, Mauch C, Hohenberger W, Brockmeyer N, Berking C, Sunderkotter C, Kaatz M, Schulte KW, Lehmann P, Vogt T, Ulrich J, Herbst R, Gehring W, Simon JC, Keim U, Martus P, Garbe C, German Dermatologic Cooperative Oncology Group (2016) Complete lymph node dissection versus no dissection in patients with sentinel lymph node biopsy positive melanoma (Decog-Slt): a multicentre, randomised, phase 3 trial. Lancet Oncol 17:757–767 Mann GB, Coit DG (1999) Does the extent of operation influence the prognosis in patients with melanoma metastatic to inguinal nodes? Ann Surg Oncol 6:263–271 Martin BM, Master VA, Delman KA (2013) Videoscopic inguinal lymphadenectomy for metastatic melanoma. Cancer Control 20:255–260 Oude Ophuis CM, Van Akkooi AC, Hoekstra HJ, Bonenkamp JJ, Van Wissen J, Niebling MG,

687

De Wilt JH, Van Der Hiel B, Van De Wiel B, Koljenovic S, Grunhagen DJ, Verhoef C (2015) Risk factors for positive deep pelvic nodal involvement in patients with palpable groin melanoma metastases: can the extent of surgery be safely minimized?: a retrospective, multicenter cohort study. Ann Surg Oncol 22(Suppl 3):S1172–S1180 Ozturk MB, Akan A, Ozkaya O, Egemen O, Oreroglu AR, Kayadibi T, Akan M (2014) Saphenous vein sparing superficial inguinal dissection in lower extremity melanoma. J Skin Cancer 2014:652123 Postlewait LM, Farley CR, Diller ML, Martin B, Hart Squires M 3rd, Russell MC, Rizzo M, Ogan K, Master V, Delman K (2017) A minimally invasive approach for inguinal lymphadenectomy in melanoma and genitourinary malignancy: long-term outcomes in an attempted randomized control trial. Ann Surg Oncol 24:3237–3244 Shen P, Conforti AM, Essner R, Cochran AJ, Turner RR, Morton DL (2000) Is the node of Cloquet the sentinel node for the iliac/obturator node group? Cancer J 6:93–97 Sholar A, Martin RC 2nd, Mcmasters KM (2005) Popliteal lymph node dissection. Ann Surg Oncol 12:189–193 Strobbe LJ, Jonk A, Hart AA, Nieweg OE, Kroon BB (1999) Positive iliac and obturator nodes in melanoma: survival and prognostic factors. Ann Surg Oncol 6:255–262 Strobbe LJ, Jonk A, Hart AA, Peterse JL, Wobbes T, Nieweg OE, Kroon BB (2001) The value of Cloquet’s node in predicting melanoma nodal metastases in the pelvic lymph node basin. Ann Surg Oncol 8:209–214 Stuiver MM, Westerduin E, Ter Meulen S, Vincent AD, Nieweg OE, Wouters MW (2014) Surgical wound complications after groin dissection in melanoma patients – a historical cohort study and risk factor analysis. Eur J Surg Oncol 40:1284–1290 Thompson JF, Hunt JA, Culjak G, Uren RF, HowmanGiles R, Harman CR (2000) Popliteal lymph node metastasis from primary cutaneous melanoma. Eur J Surg Oncol 26:172–176

Neck Dissection and Parotidectomy for Melanoma Brian Gastman, Rebecca Knackstedt, Ryan P. Goepfert, Baran Sumer, Ashok Shaha, and Michael E. Kupferman

Contents Head and Neck Lymphatics and Their Impact on Melanoma Outcomes . . . . . . . . . 690 Neck Dissection and Parotidectomy for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692 Technique for Neck Dissection and Parotidectomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695 Completion Lymph Node Dissection Utility in Head and Neck Melanoma . . . . . . . 700 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 701 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702

Abstract B. Gastman (*) Department of Plastic Surgery, Cleveland Clinic, Lerner Research Institute, Cleveland, OH, USA e-mail: [email protected] R. Knackstedt Department of Plastic Surgery, Cleveland Clinic, Cleveland, OH, USA e-mail: [email protected] R. P. Goepfert Department of Head and Neck Surgery, UT MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] B. Sumer Department of Otolaryngology, UT Southwestern Medical Center, Dallas, TX, USA e-mail: [email protected] A. Shaha Department of Otolaryngology, Memorial Sloan Kettering Cancer Institute, New York, NY, USA M. E. Kupferman Department of Head and Neck Surgery, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_23

Head and neck melanomas are a surgical and medical challenge. Facial anatomy and its lymphatic drainage are complex and often unpredictable. To balance oncologic clearance with morbidity, surgeons must be familiar with the anatomy of the head and neck region, possible patterns of nodal drainage, and regions of potential anatomic danger. Unless facial surgery is part of a surgeon’s common practice, facial nerve injury and soft tissue embarrassment can result. Although head and neck melanomas are treated similarly to other body regions, they are often underrepresented in clinical trials, making extrapolation difficult. Thus, the surgeon must be aware of current clinical data and tailor care to each individual patient. The goal of this chapter is to review head and neck lymphatics, discuss relevant surgical techniques, and review clinical evidence relating to sentinel node biopsy and

689

690

complete neck dissection in head and neck melanomas. Keywords

Melanoma · Lymphatics · Sentinel node · Neck dissection · Parotidectomy · Parotid

Head and Neck Lymphatics and Their Impact on Melanoma Outcomes Cephalization in animals is an ancient evolutionary adaptation in which the mouth, major sense organs, and the majority of neural tissue become concentrated at the front end of an animal, producing a distinct head. The concentration of these tissues in one anatomic location has made the head and neck region of Bilateria, an animal phylum that contains the majority of animal species including humans, the most complex in the body. This complexity is mirrored in the lymphatics of the head and neck, as approximately one third of all lymph nodes in the body are located in this region. While lymph nodes and lymphatic vessels were described in ancient times by Hippocrates and Herophilos, later anatomists such as Gaspar Asellius and Antony Nuck rediscovered the lymphatic system in the seventeenth century (Subramanian et al. 2007). In 1787, Paolo Mascagni published the first detailed anatomic description of the lymphatic system, which was updated in 1909 by Poirer and then specifically described for the head and neck by H.A. Trotter in 1930. The classification system accepted throughout the majority of the twentieth century for lymph nodes in the head and neck was established by Rouviere. This system classified cervical nodes into groups in distinct anatomic locations that are roughly oriented in horizontal (submental, facial, submandibular, parotid, mastoid, occipital, and superficial cervical) and vertical, deep planes (anterior and posterior cervical). The advantage of this system was that it established the relationship between the nodal levels in relation to adjacent anatomic structures. A simplified system more congruent with the surgical technique for neck dissection, in which numerical levels are assigned for different nodal

B. Gastman et al.

groups, was adopted by Shah et al. (1981). Rather than searching for and removing nodes in a discontinuous fashion from different zones of the neck, a neck dissection follows fascial planes in a systematic way to remove lymph node-bearing tissue. This technique minimizes the chance that a node in a given anatomic zone will be missed, allows for greater standardization, and minimizes the effect of variables such as the amount of fat in the tissue that can alter the efficiency of nodal removal. This system was updated and expanded by classification systems from the American Academy of Otolaryngology and the American Joint Committee on Cancer dividing the lymph nodebearing basins into levels that are more easily described when standardizing surgery and the delivery of radiotherapy. A neck dissection classification update by the American Head and Neck Society and the American Academy of Otolaryngology-Head and Neck Surgery in 2002 delineates these nodal levels and can be compared to the anatomic correlates in the TNM atlas for lymph nodes of the neck (Robbins et al. 2002; Hermanek et al. 1988; Gregoire et al. 2014). Lymph nodes in areas not delineated by numbered levels are described anatomically and include superior mediastinal nodes, retropharyngeal nodes, periparotid nodes, buccinator nodes, postauricular nodes, and suboccipital nodes. Any discussion on lymphatic drainage, clinically demonstrated with lymphoscintigraphy, would be remiss not to acknowledge the important work done by Dr. Roger Uren and Dr. Hayley Reynolds. Without their pioneering work, it would be impossible to preoperatively predict the location of sentinel nodes, especially in a region as complex as the head and neck (Uren 2004; Reynolds et al. 2009; de Wilt et al. 2004). In two studies investigating drainage patterns from head and neck melanoma, bilateral drainage was found to be rare – occurring approximately 10% of the time with drainage to sentinel lymph nodes following expected patterns of anatomy (Suton et al. 2012). On the contrary, Lin et al. found that the rate of discordance was high with sentinel lymph nodes identified by lymphoscintigraphy in areas that were not clinically predicted to be at risk as well as areas that are not typically

Neck Dissection and Parotidectomy for Melanoma

691

dissected during standard neck dissections such as the postauricular region (Lin et al. 2006). The discordancy is highest in the head and neck for melanoma with at least 36% of cases showing at least one sentinel lymph node in an area that would not be clinically predicted compared to discordance rates of 25% for melanomas of the trunk and 13% for melanomas of the extremities (Leong et al. 1999). Another study found drainage to unexpected nodal regions in 13 of 51 patients with 18 of 51 patients having parotid metastases (Fincher et al. 2004). One study in patients undergoing parotidectomy and neck dissection for head and neck cutaneous melanoma demonstrated that parotid disease was associated with high likelihood of occult neck disease in greater than 40% of patients (Suton et al. 2012). These authors further recommended a posterior lymphadenectomy for melanomas in the scalp and posterior neck (see also chapter ▶ “Lymphoscintigraphy in Patients with Melanoma”).

More recently, a 2013 update by Gregoire et al. divided the neck into 10 nodal groups creating levels for these areas that are at risk of regional metastasis from malignancies arising in the skin (Gregoire et al. 2014). This system is in many ways more useful for the treatment of nodal disease in melanoma. The original six levels of the neck were described and optimized to standardize neck dissections for metastatic disease from the upper aerodigestive tract. Therefore, nodal levels most likely to be involved with metastatic disease from these areas were included, and many levels such as the parotid and postauricular nodes which are often involved with melanoma were excluded in formal definition of levels with the prior neck dissection classification system. This comprehensive updated description is useful for melanoma since in addition to cutaneous sites, melanoma can very occasionally present in the orbit, nasal cavity, sinuses, and oral cavity as well. Figure 1 and Table 1 demonstrate these nodal levels as well as the areas they receive drainage from. Clearly, an understanding of the complex drainage

Fig. 1 Nodal levels of the head and neck. (Adapted from Gregoire modification of nodal levels from the American Head and Neck Society and the American Academy of

Otolaryngology-Head and Neck Surgery * illustrations by Anna Tomkies)

692

B. Gastman et al.

Table 1 Nodal levels of the head and neck. Adapted from Gregoire modification of nodal levels from the American Head and Neck Society and the American Academy of Otolaryngology-Head and Neck Surgery Level Ia Ib

Anatomic location Submental group Submandibular group

II

Upper jugular group

III

Middle jugular group

IVa

Lower jugular group

IVb

Medial supraclavicular group

V Va

Posterior triangle group Upper posterior triangle nodes Lower posterior triangle nodes Lateral supraclavicular group Anterior compartment group Anterior jugular nodes Prelaryngeal, pretracheal, and paratracheal nodes Prevertebral compartment group Retropharyngeal nodes Retro-styloid nodes Parotid group

Vb Vc VI VIa VIb VII VIIa VIIb VIII IX X Xa Xb

Bucco-facial group Posterior skull group Retroauricular and subauricular nodes Occipital nodes

Receives drainage from Chin, lower lip, tip of tongue, anterior floor of mouth Midface, level Ia, inferior nasal cavity, palate, alveolar ridge, cheek, lips, anterior tongue Face, parotid, levels Ia and Ib, retropharyngeal nodes, nasal cavity, pharynx, larynx, external auditory canal, middle ear, sublingual submandibular glands Levels II and V, retropharyngeal and pre- and paratracheal nodes, pharynx, larynx, thyroid Levels III and V, retropharyngeal and pre- and paratracheal nodes, pharynx, larynx, thyroid Levels Iva and Vc, pre- and paratracheal nodes, hypopharynx, esophagus, larynx, trachea, thyroid Occipital and parietal scalp, skin of lateral and posterior neck and shoulder, occipital and retroauricular nodes, nasopharynx, pharynx, thyroid Occipital and parietal scalp, skin of lateral and posterior neck and shoulder, occipital and retroauricular nodes, nasopharynx, pharynx, thyroid Levels Va and Vb Lower face anterior neck skin Anterior floor of mouth, tip of tongue, lower lip, thyroid, larynx, pharynx, esophagus

Nasopharynx, soft palate Nasopharynx Forehead, temporal scalp, eyelid, orbit, auricle, ear canal, tympanic membrane, nasal cavity, nasopharynx Nose, eyelids, cheek Posterior auricle, ear canal, scalp posterior to ear canal Posterior scalp

patterns from these sites in the head and neck is essential for sentinel lymph node biopsy and the prediction of metastases when treating melanoma (see also chapters ▶ “Biopsy of the Sentinel Lymph Node” and ▶ “Lymphoscintigraphy in Patients with Melanoma”).

Neck Dissection and Parotidectomy for Melanoma Regional lymph node metastasis represents the most significant prognostic factor in early stage melanoma (Morton et al. 1991; Gershenwald et al. 1999; Balch et al. 2001). Given the potential

morbidity of elective nodal dissection, sentinel lymph node biopsy was developed and now represents the standard of care for identifying patients with clinically occult metastatic melanoma (Balch et al. 2000; Morton et al. 2006). In patients with primary melanoma located in the head and neck region, surgical management may entail either sentinel lymph node biopsy in the parotid gland and/or neck to identify clinically occult disease or therapeutic parotidectomy and/or neck dissection in patients with clinically apparent regional disease. In the evolving era of effective systemic therapies, considerations regarding type and extent of surgery represent topics of increasing debate among melanoma providers. This is particularly evident among

Neck Dissection and Parotidectomy for Melanoma

693

Fig. 2 Axial and coronal SPECT-CT images demonstrating utility of differentiating sentinel nodes in the superficial parotid versus deep parotid/upper level II. (Picture courtesy of Dr. Goepfert)

Fig. 3 Resection of incision of previously narrowly excised melanoma overlying parotid gland. (Picture courtesy of Dr. Goepfert)

head and neck surgeons who may be reluctant to risk compromising locoregional control by not performing completion parotidectomy and/or cervical lymphadenectomy (Faries et al. 2017). Classically, primary cutaneous lesions anterior to a line bisecting the ear canal drain to lymphatic channels in the parotid gland and anterior neck levels I–IV. Routine use of single-photon emission computed tomography-computed tomography (SPECT-CT) has improved preoperative localization of sentinel lymph nodes in three dimensions (Van Der Ploeg et al. 2009; Zender et al. 2014; Chapman et al. 2016) (Fig. 2). The vast majority of lymph nodes in the parotid gland are located in the superficial lobe (superficial to the facial nerve), and it is very rare to have a cutaneous lesion drain to the deep lobe of parotid as a first echelon site on lymphoscintigraphy (McKean et al. 1985). Though the nerve cannot be seen as it courses through the parotid gland, use of the retromandibular vein can give an approximation of the delineation between superficial and deep lobes to aid in surgical planning (Divi et al. 2005). For sentinel nodes in the superficial lobe, gentle blunt dissection will locate the node without need for facial nerve identification (Ollila et al. 1999; Samra et al. 2012). Moreover, concomitant use of radiotracer and blue dye can further assist with rapid and accurate localization of the sentinel node within the parotid parenchyma. In cases where the sentinel node is unable to be located with blunt dissection, a complete superficial parotidectomy with formal dissection of the facial nerve may be considered for diagnostic purposes. Similarly, for patients with primary lesions

draining to or through the parotid gland and clinically apparent adenopathy in the parotid gland and/ or upper neck, a superficial parotidectomy should be performed. Moreover, for lesions overlying the parotid gland, a superficial parotidectomy is sometimes needed for diagnostic and/or therapeutic purposes (Fig. 3). Performance of a total parotidectomy may improve disease control in some patients, but this must be balanced with increased morbidity (Wertz et al. 2017). Every effort should be made to preserve the facial nerve, even for extensive adenopathy and the rare cases of macroscopic nodes deep to the facial nerve. Completion lymphadenectomy was recently demonstrated to improve regional disease control but not melanoma-specific survival in patients with positive sentinel lymph node biopsies, but these findings have unclear application to melanoma of the head and neck region (Leiter et al. 2016; Faries et al. 2017). Conversely, therapeutic neck dissection for management of metastatic melanoma with clinically apparent neck metastases remains the standard of care though the extent and timing of surgery are a topic of great debate and are evolving along with the increasing use of more effective systemic therapies (Supriya et al. 2014; Faries 2018). The days of a uniform approach with modified radical neck dissection +/ parotidectomy for clinically apparent disease have passed (Shah et al. 1991). Similar to the decision-making for parotidectomy regarding lymphatic drainage anterior or posterior to a line bisecting the ear canal, the levels of neck dissection should vary. Primary lesions on the face or

694

B. Gastman et al.

Fig. 4 Superficial parotidectomy with level I-III neck dissection for anterior face melanoma. (Picture courtesy of Dr. Goepfert)

scalp anterior to the ear with clinically apparent metastases in general should have dissection of the perifacial nodes and levels I-III/IV in addition to a superficial parotidectomy (Fig. 4). On the other hand, primary lesions on the upper neck or scalp posterior to the ear in general should have dissection of levels II–V, as well as suboccipital nodes for scalp lesions (Fig. 5). In all situations, the decision about whether to include level IIb is controversial, but our general approach is to dissect level IIb if we are performing a neck dissection, particularly for primary lesions of the scalp, ear, or periauricular regions, or if there is adjacent macroscopic disease (Creighton et al. 2016). Surgical planning in melanoma has become increasingly multidisciplinary given the introduction of neoadjuvant and adjuvant therapeutic approaches. However, several principles regarding sentinel node biopsy, parotidectomy, and neck dissection remain useful. Sentinel node incisions should be planned along incision lines for parotidectomy and/or neck dissection should these procedures eventually be required. Parotidectomy incisions are most often made in

Fig. 5 Posterolateral levels II-IV (a) and level V (b) neck dissection for scalp melanoma. (Picture courtesy of Dr. Goepfert)

a preauricular crease and curve around the lobule toward the postauricular area into a natural neck skin crease (Fig. 6). The postauricular portion of the incision should not extend too far cephalad prior to curving into a skin crease; otherwise the blood supply may be compromised causing skin necrosis (Fig. 7). Skin flap elevation over the parotid gland should split the fat (“fat up, fat down”) and avoid becoming too thin, exposing the sweat glands, and increasing the risk or severity of Frey’s syndrome (gustatory sweating)

Neck Dissection and Parotidectomy for Melanoma

695

Fig. 8 Skin flap elevation should avoid being too thin to expose the sweat glands and lessen the risk of Frey syndrome. (Picture courtesy of Dr. Goepfert)

Fig. 6 Standard parotidectomy incision. (Picture courtesy of Dr. Goepfert)

surgical training for feedback regarding the degree of nerve stimulation during dissection. In the case of patients treated with neoadjuvant therapy with radiologic evidence to support response to treatment, our burgeoning experience is in favor of a more selective neck dissection with great effort to preserve nerves and blood vessels even despite intimate association with apparent gross nodal disease (Hodi et al. 2016; Raigani et al. 2017). In such patients, the intraoperative appearance of the nodal disease may be indistinguishable from patients who have never received neoadjuvant therapy, yet final pathology often demonstrates sheets or aggregates of pigmented macrophages without evidence of viable melanoma. In either case, decisions on the extent of surgical resection for patients with regional metastases must be made with multidisciplinary support and must carefully balance locoregional control and accurate staging with surgical morbidity (Faries 2018).

Fig. 7 Necrosis of superior skin tip. (Picture courtesy of Dr. Goepfert)

(Durgut et al. 2013) (Fig. 8). Facial nerve identification should be undertaken in every parotidectomy. Standard landmarks including the tragal pointer, posterior belly of digastric, and tympanomastoid suture line should be used to locate the facial nerve, while use of a facial nerve monitoring system is dependent on surgeon preference (Fig. 9). In our experience, use of a nerve monitor can be helpful in the setting of

Technique for Neck Dissection and Parotidectomy Surgical removal of clinically palpable nodes is termed a therapeutic neck dissection, whereas removal of clinically negative nodes is elective or prophylactic neck dissection. Comprehensive or modified radical neck dissection involves complete dissection of all five neck levels, whereas selective neck dissection involves dissection of

696

B. Gastman et al.

Fig. 9 Identification of all branches of the facial nerve. (Picture courtesy of Dr. Goepfert)

only certain node levels or groups. According to this terminology, the only dissection that requires no additional description is radical neck dissection. This dissection involves removal of the lymph nodes of all five levels, along with the internal jugular vein, the spinal accessory nerve, and the sternocleidomastoid muscle. For other comprehensive dissections, it must be specified which structures should be preserved (in modified radical neck dissection) or additionally removed (in extended radical neck dissection). A number of descriptions of neck dissections can be found in standard surgical texts. Our preferred technique will be described in this section, but the final selection of an incision and overall conduct of the operation will depend on the surgeon’s personal preference, experience, and training. The patient’s presentation will also influence the selection of a technique. Previous incisions, skin grafts, and other surgical procedures must be taken into account when planning the operation. In patients with clinical evidence of cervical nodal metastasis, a therapeutic neck dissection including levels I to V is indicated. Most often a therapeutic neck dissection is a modified radical neck dissection that spares the sternocleidomastoid muscle, the internal jugular vein, and the spinal accessory nerve.

The initial steps in planning the operation are the most important in avoiding poor outcomes. Planning the skin incision is the starting point for avoiding problems. Patients who have had previous operations will have preexisting scars that must be taken into account. Failure to carefully evaluate this will lead to parallel incision lines, which can result in poor healing and wound dehiscence. The incision is marked with a skin marker. Once the incision lines are marked on the neck, it is useful to infiltrate the skin and subcutaneous tissue with a local anesthetic mixed with epinephrine, 1:100,000 or 1:200,000, to minimize bleeding when the skin is incised. Many options are available for acceptable incisions that provide adequate exposure. One such incision is the lazyS incision, which is initiated with a transverse incision beginning at the mastoid process and running across the lateral neck approximately two fingerbreadths below the inferior border of the mandible. The exact position of the incision should be adjusted so that it will fall in a natural skin crease. The incision should be designed so that dehiscence would not lead to disastrous exposure of the underlying vessels. A vertical incision is carried from the transverse incision to the midpoint of the clavicle in a lazy-S configuration. The gentle curve of the incision will help to decrease

Neck Dissection and Parotidectomy for Melanoma

the incidence of scar contracture that is often seen with straight-line incisions. The transverse incision is opened through the skin, subcutaneous tissue, and platysma muscle, and the vertical incision is incised. Care should be exercised to avoid injury to the external jugular vein; it will be resected at a later point in the operation. The skin flaps are elevated with the use of skin hooks to control the tissue and avoid the crushing or scarring that can occur with standard forceps. Another option for incision design is the hockeystick or J-incision. The J-incision employs a vertical incision extending along the posterior aspect of the sternocleidomastoid muscle from the mastoid apex to 2 cm above the clavicle, which is then curved in the horizontal plane parallel to the bottom edge of the cricoid cartilage and extending medially to the point where the sternocleidomastoid muscle is attached to the clavicular head. A host of other incisions have been described in the literature and have been safely and successfully employed for exposure in neck dissection. These can be used at the discretion and comfort of the surgeon. However, the basic techniques of gentle tissue handling apply in all cases. The operation begins in the posterior triangle (level V). There is no platysma posterior to the sternocleidomastoid muscle. This makes the plane of dissection a bit more technically sensitive as far as the thickness of the flaps is concerned. As the superior flap is raised, care is taken to stay on the sternocleidomastoid muscle to avoid injury to the parotid gland. The inferior flap is raised, and the great auricular nerve is identified. The spinal accessory nerve exits deep to the sternocleidomastoid muscle approximately 1 cm cephalad to the greater auricular nerve. The flap is now raised at the level of the superficial layer of the deep cervical fascia. Care must be taken to avoid injury to the spinal accessory nerve. The soft tissue flap is elevated in a cephalad direction to the mastoid process. A cautery is used to minimize bleeding, and the lateral edge of the trapezius muscle is identified. The muscle is quite superficial, and care must be taken to avoid “buttonholing” the flap. Once the edge of the muscle is identified, a curved clamp is placed along the edge of the muscle, and the tissue is

697

divided by means of cautery. The spinal accessory nerve traverses the lateral neck to enter the trapezius at the junction of its middle and inferior third. This dissection should stay just dorsal to the edge of the muscle. As the junction of the middle and inferior thirds is approached, careful spreading of the fat will reveal the accessory nerve. Now that this is identified, the remainder of the flap can be elevated to the clavicle. The nerve is freed of its investing fascia using a nerve hook and sharp tissue scissors. Bipolar cautery is preferred to standard unipolar cautery for this portion of the dissection because it is more controlled and less likely to result in nerve injury caused by current transmission and heat conduction. The nerve is freed to the point where it exits from the posterior border of the sternocleidomastoid muscle. All of the fibrofatty node-bearing tissue is dissected from the posterior triangle of the neck starting high, at the mastoid process. The fascia overlying the muscles, starting with the splenius capitis, is elevated with the cautery. This dissection proceeds in a caudal direction. The accessory nerve is retracted off of the tissue, and the specimen is passed deep to it. The soft tissue in the lower portion of level V is elevated from lateral to medial. Numerous veins and small vessels will be found in this supraclavicular fat pad. The transverse cervical artery and vein are identified, ligated, and divided. The posterior belly of the omohyoid muscle is identified. The muscle is freed of its investing fascia so it can be retracted. This muscle is an important landmark in the lower neck. It lies superficial to the brachial plexus and the phrenic nerve as it crosses the anterior scalene muscle. On the left side of the neck, in the area where the internal jugular vein passes deep to the omohyoid muscle, the thoracic duct enters in the vicinity of its junction with the subclavian vein. The fat contains numerous lymphatic channels and nodes. There is significant potential for lymphatic leakage. The soft tissue in this area is best managed by division between clamps and then ligation to avoid lymphatic leakage. Now that level V has been dissected, the specimen is elevated to the posterior aspect of the internal jugular vein in the lower neck. Numerous spinal roots will

698

be encountered as they pass between the paraspinous muscles into the fat of the posterior triangle. An attempt should be made to preserve the motor roots. The sensory roots will be divided as they exit the neck to provide sensation to the upper chest and shoulder area. These areas of cutaneous numbness will improve over time, with significant return of feeling. The operation is now directed to the upper portion of the sternocleidomastoid muscle at its insertion into the mastoid process. The investing fascia is elevated from the muscle using the cautery. In the upper neck, the muscle is retracted posteriorly, and spreading the soft tissue in the area of the angle of the mandible, superficial to the internal jugular vein, will reveal the posterior belly of the digastric muscle, an important landmark in the upper neck. The internal jugular vein passes just deep to the digastric muscle, and the eleventh cranial nerve is usually found lying on or immediately posterior to the vein. The nerve is freed of its investing tissue and elevated with a nerve hook. The bipolar cautery is used to mobilize the tissue of level Ilb and pass it under the spinal accessory nerve. Now, with the use of a curved clamp, the fibrofatty, node-bearing tissue is elevated and divided with the cautery. The sternocleidomastoid muscle is completely freed from its fascia so that it can be retracted away from the surgical bed. The dissection moves from cephalad to caudad. The roots of the cervical plexus are identified, and an attempt is made to preserve as many of the motor roots as possible. Many of the sensory roots will need to be divided. The sternocleidomastoid muscle is completely freed of investing fascia and elevated. The specimen from the posterior neck is passed deep to the muscle. The omohyoid muscle is retracted so that level IV can be accessed. The entire specimen is now gradually elevated off of the internal jugular vein. As the dissection comes anterior to the internal jugular vein, the common facial vein is ligated and divided. The descendens hypoglossi is identified just anterior to the vein and preserved in an effort to maintain motor function to the strap muscles. The omohyoid muscle is followed up to the hyoid bone with cautery. The submental triangle is

B. Gastman et al.

entered, and the nodal tissue is elevated off of the mylohyoid muscle. The anterior belly of the digastric muscle is identified. The superior soft tissue flap is elevated staying close to the deep surface of the platysma muscle. As the caudal edge of the submandibular salivary gland is approached, the operation slows, and careful spreading in the soft tissue will reveal the marginal mandibular branch of the facial nerve directly on the capsule of the salivary gland. The nerve is dissected with sharp scissors and/ or bipolar cautery and retracted upward with the upper skin flap. Another technique would be to identify the posterior facial vein, ligate and divide it, and reflect the stump up with the skin flap, and the nerve will be safely found between the stump of the vein and the platysma muscle. Once the nerve is out of the way, the soft tissue at the inferior border of the mandible is divided. The facial artery and vein are ligated and divided. The dissection now proceeds from anterior to posterior. The anterior belly of the digastric muscle is retracted, and cautery is used to elevate the tissue off of the mylohyoid muscle until the posterior edge is identified. A number of veins will be encountered, and they should be ligated. A small retractor is used to retract the mylohyoid anteriorly. This will allow access to the submandibular triangle. The gland is freed and carefully retracted inferiorly. The lingual nerve is identified deep to the gland. The secretory motor fibers are identified, ligated, and divided. The gland can now be further retracted exposing Wharton’s duct and the hypoglossal nerve. The duct should be clamped and divided only after the hypoglossal nerve is identified as it crosses deep to the digastric tendon. This will avoid inadvertent injury to the nerve. Once the duct is divided, the gland can be elevated with the use of a curved clamp and cautery. The proximal facial artery is divided and suture ligated. The entire specimen can be elevated and delivered in one piece. The soft tissue is carefully labeled for the pathologist. Levels I to V should be indicated with sutures or other markers. A Valsalva maneuver is used to check for hemostasis. The flaps are closed in multiple layers over a closed suction drain, taking care to reapproximate the platysma muscle.

Neck Dissection and Parotidectomy for Melanoma

Elective dissections of levels I to III or I to IV are essentially equivalent and should be combined with a superficial parotidectomy for melanomas of the anterior scalp, face, and anterior upper neck. Level IV should be included when the dissection is being carried out for clinical disease in level I. This procedure can be carried out through a single, large cervicofacial incision. Elevation of the skin flap should be carried out with care in the submandibular region because the marginal mandibular nerve may be injured at this site. The neck dissection involves preservation of the sternocleidomastoid muscle, internal jugular vein, and spinal accessory nerve and begins with division of the investing layer of fascia along the posterior half of the sternocleidomastoid muscle. This fascia is elevated, releasing the sternocleidomastoid muscle, which is retracted posteriorly, allowing dissection onto the deep surface of the muscle. The external jugular vein and great auricular nerve are usually divided. Dissection on the deep surface of the sternocleidomastoid muscle continues until the posterior border of the muscle is reached. The branches of the cervical plexus form the deep plane of the dissection as they pass laterally behind the posterior border of the sternocleidomastoid muscle. Level III is dissected first, and the juguloomohyoid node should be identified and removed. Dissection proceeds cephalad, and at level II, the upper end of the sternocleidomastoid muscle is retracted posteriorly, allowing careful dissection of the tissue around the spinal accessory nerve and the upper end of the internal jugular vein. When a parotidectomy is to be included, the facial nerve should then be identified and the parotidectomy carried out next. The lymph nodes of the submandibular triangle, which lie in relation to the facial vessels as they cross the jaw, should then be removed, but this may be left as the last step to minimize the risk of injury to the marginal mandibular branch of the facial nerve. The other common selective procedure involves dissection of levels II to V for posteriorly situated melanomas. This may be a therapeutic procedure when involved nodes are present in the posterior triangle and the primary lesion is located on the posterior scalp. The procedure

699

should include removal of the occipital nodes because these may be a site of subsequent recurrence. Selective dissections of levels III to V may be carried out through a single incision across the lower neck, but, as is the case for all dissections for melanoma, the nodes in the investing fascia at the upper third of the sternocleidomastoid muscle should be included. The supraclavicular area is rich in lymph nodes that may be involved with metastatic disease from a primary melanoma of the anterior or posterior thorax, upper extremity, or head and neck. Patients with extensive nodal disease in the axilla may also exhibit disease extending into the neck. The operation will address nodes at risk in levels III, IV, and V. The area is approached with a collar incision, and usually no other incision is needed. The spinal accessory nerve will need to be identified as soon as the skin flaps are elevated. The nerve is freed of its investing fascia and elevated so that the specimen can be safely passed under it. The dissection then proceeds to the clavicle. The lower portion of level IV, where the internal jugular vein joins the subclavian vein, is an area that requires careful attention. This area should be cleared of all nodes in an effort to avoid recurrence in a difficult location. As mentioned earlier, the confluence of the internal jugular vein and subclavian vein on the left side is the location of the thoracic duct. The tissue in this area will need to be clamped, divided, and ligated. Frequently the standard preauricular incision that is carried into the neck will need to be incorporated into a neck dissection incision to allow for an adequate cervical lymphadenectomy. This is especially true when clinical disease is present because parotidectomy should be accompanied by a neck dissection, allowing removal of all atrisk nodes. The incision should be carefully marked in a preauricular skin crease and then carried into the neck by curving it around the tail of the parotid gland. The incision line is injected with local anesthetic with epinephrine I:100,000. It is then best to wait a few minutes until the skin begins to blanch. This will greatly improve hemostasis at the skin edge. The skin flap is elevated directly on the parotid fascia, a thick white layer of tissue that is readily

700

identified. If fat is present on the deep plane, dissection is not deep enough. If parotid tissue is seen, the fascia has been violated, and the dissection plane is too deep. The tail of the parotid gland should now be elevated from the sternocleidomastoid muscle, and that muscle should be retracted posteriorly. Further dissection deep to the parotid gland will allow identification of the posterior belly of the digastric muscle. This is a useful landmark because it has its origin on the deep surface of the mastoid process, which in turn assists with identification of the facial nerve. The parotid gland is separated from the cartilaginous auditory canal. This is easily done with a clamp and cautery. This proceeds until the junction of the bony and cartilaginous canal is reached. The parotid tail is elevated from the posterior belly of the digastric muscle. Gentle anterior traction is placed on the gland. A fine curved clamp is used to spread the gland at the point of the digastric origin on the mastoid process. The glandular tissue is separated and carefully divided with bipolar electrocautery. The main trunk will be found approximately 1 cm into the gland. Once the main trunk of the facial nerve has been identified, the dissection is slowly carried out to identify the bifurcation of the upper and lower divisions. There is great variability in the anatomy of the extracranial facial nerve. Care must be taken to confirm that the main trunk of the facial nerve has truly been identified. The dissection is carried out in a plane just superficial to the nerve. A very fine curved clamp is used to gently spread the tissue, and bipolar cautery is used to divide the tissue in an absolutely bloodless field. Dissection proceeds toward the periphery, keeping the nerve in constant view between the tines of the clamp. The superficial temporal artery and vein may need to be ligated and divided. As the cephalad portion of the dissection is completed, the gland is reflected caudad, and the buccal branches are carefully identified. These branches often run parallel to Stensen’s duct. Once the duct is identified, it should be ligated and divided. The dissection of the lower division of the nerve proceeds in a similar manner until the gland is totally freed and delivered.

B. Gastman et al.

Removal of the deep lobe, which is rarely required, begins just as the superficial parotidectomy is begun. Once the superficial lobe is removed, the main trunk of the facial nerve is carefully dissected from investing tissue and is mobilized. The deep lobe is mobilized using blunt dissection and bipolar cautery to eventually be passed under the nerve. A closed suction drain is placed, and the flaps are carefully reapproximated. The wound is closed in multiple interrupted layers, and a subcuticular suture is used to close the skin. No dressings are needed, but a thin film of antibiotic ointment can be placed on the incision line.

Completion Lymph Node Dissection Utility in Head and Neck Melanoma The role of a sentinel lymph node biopsy (SLNB) and the merit of performing a completion lymph node dissection (CLND) in melanoma were investigated in the multicenter selective lymphadenectomy trials (MSLT-I and II) and the German DeCOG-SLT trial. However, head and neck melanomas, despite representing 18% of all primary melanomas, were underrepresented or not included in these studies (Lachiewicz et al. 2008). (See also chapter ▶ “Biopsy of the Sentinel Lymph Node.”) The first Multicenter Selective Lymphadenectomy Trial (MSLT-I), published in 2014, randomized patients with a primary melanoma to observation or SLNB of their nodal basin. In the observation arm, nodal dissection was only performed if and when lymph node metastases became clinically evident. In the SLNB group, SLNB was performed at the time of wide-local excision with immediate lymph node dissection if positive. Importantly, the proportion of head and neck melanomas in this study was not indicated. At 5 years, the hazard ratio (HR) for recurrence in head and neck melanoma patients assigned to SLNB was 1.08, and the HR for death from melanoma in this group was 1.18, neither of which was statistically significant (Morton et al. 2006). At 10 years, these HRs were both 1.2, still not reaching statistical significance. The 10-year analysis importantly

Neck Dissection and Parotidectomy for Melanoma

demonstrated in subset analysis that in patients with intermediate-thickness melanoma (1.2–3.5 mm), early CLND after positive SLNB improved 10-year distant disease-free and melanoma-specific survival (Morton et al. 2014). The second Multicenter Selective Lymphadenectomy Trial (MSLT-II), published in 2017, randomized patients with a positive SLNB to undergo either CLND or observation with frequent nodal ultrasound examination. This was an international, randomized, phase three trial conducted at 63 centers and involving 1939 patients with a median follow-up of 43 months. Both dissection and observation groups consisted of 13.7% head and neck melanomas, lower than the incidence rate of 18% of new melanomas. This study demonstrated that while immediate CLND impacted staging and regional disease control by reducing the rate of regional nodal recurrence by 69% (HR 0.31; 95% CI 0.24–0.41, p < 0.001) compared to delayed CLND, it did not provide a survival benefit. When compared to the arm/leg subsites, the HR in the head and neck subgroup for melanoma-related deaths was 0.81 (not significant) in the dissection group and 1.6 ( p = 0.07) in the observation group. However, due to the small sample size, this may have reached significance with an increased number of patients (Faries et al. 2017). Another important clinical trial, published in 2016, a German phase three multicenter randomized control trial (RCT) (DeCOG-SLT), while underpowered, also showed no difference in survival in SLNB-positive patients treated with CLND compared with observation only (Leiter et al. 2016). However, the head and neck melanoma subsite was an exclusion criterion for this study. The fact that head and neck melanomas are often underrepresented in or excluded from clinical trials investigating the utility of CLND has led some to believe that head and neck melanoma patients may be more likely to benefit from CLND. If the determination of performing a CLND is to be based on SLNB status, the head and neck region makes this difficult, as the accuracy of the SLNB may be less reliable. There are over 300 lymph nodes in the head and neck (Hyde and Prvulovich 2002), and lymphatic

701

drainage is unpredictable and discordant and can be bilateral to multiple nodal levels (Even-Sapir et al. 2003; Uren 2004; Wells et al. 1994; Chao et al. 2003; Morton et al. 1993; Shah et al. 1991; Garbe et al. 1995; Albertini et al. 1996; O’brien et al. 1995; Leong et al. 1999; Eberbach et al. 1987; Lin et al. 2006; Klop et al. 2011; Macneill et al. 2005; Teltzrow et al. 2007). With the complex anatomy of the head and neck, surgical inexperience can result in an inability to identify the correct node(s) (Scoggins et al. 2010). From an imaging standpoint, the close proximity of the primary lesion to the nodes can mask the radioactive signal, and the parotid uptake of free technetium can lead to a high background signal (Carlson et al. 2000; Eicher et al. 2002; Jansen et al. 2000). A recent publication calculated the false-negative rate (FNR) in the literature for head and neck melanoma SLNB using the formula FN/TP + FN. FNR ranged from 0 to 50%, with an average of 22.4% (Knackstedt et al. 2018). Thus, it is difficult to base the decision on whether or not to perform a CLND based on the status of the SLNB if the reliability of the biopsy is questionable. (See also chapter ▶ “Lymphoscintigraphy in Patients with Melanoma.”)

Conclusion Head and neck melanomas are a surgical challenge due to the complex anatomy of the facial region and its unpredictable nodal drainage. When managing these patients, surgeons must be thoroughly familiar with the anatomy of the head and neck region, possible patterns of nodal drainage, and regions of potential danger to anatomical structures. Head and neck melanomas are treated with the same guiding principles as melanomas of other regions; however, they are often underrepresented in clinical trials. Treatment of head and neck melanoma must balance the need for oncologic control with the risk of surgical morbidity. As clinical trials continue to guide treatment paradigms for melanoma, it will be crucial to ensure that head and neck melanomas are appropriately represented in order to determine the best strategies for patient care.

702

References Albertini JJ, Cruse CW, Rapaport D, Wells K, Ross M, Deconti R, Berman CG, Jared K, Messina J, Lyman G, Glass F, Fenske N, Reintgen DS (1996) Intraoperative radio-lympho-scintigraphy improves sentinel lymph node identification for patients with melanoma. Ann Surg 223:217–224 Balch CM, Soong S, Ross MI, Urist MM, Karakousis CP, Temple WJ, Mihm MC, Barnhill RL, Jewell WR, Wanebo HJ, Harrison R (2000) Long-term results of a multi-institutional randomized trial comparing prognostic factors and surgical results for intermediate thickness melanomas (1.0 to 4.0 mm). Intergroup Melanoma Surgical Trial. Ann Surg Oncol 7:87–97 Balch CM, Soong SJ, Gershenwald JE, Thompson JF, Reintgen DS, Cascinelli N, Urist M, McMasters KM, Ross MI, Kirkwood JM, Atkins MB, Thompson JA, Coit DG, Byrd D, Desmond R, Zhang Y, Liu PY, Lyman GH, Morabito A (2001) Prognostic factors analysis of 17,600 melanoma patients: validation of the American Joint Committee on Cancer melanoma staging system. J Clin Oncol 19:3622–3634 Carlson GW, Murray DR, Greenlee R, Alazraki N, FrySpray C, Poole R, Blais M, Hestley A, Vansant J (2000) Management of malignant melanoma of the head and neck using dynamic lymphoscintigraphy and gamma probe-guided sentinel lymph node biopsy. Arch Otolaryngol Head Neck Surg 126:433–437 Chao C, Wong SL, Edwards MJ, Ross MI, Reintgen DS, Noyes RD, Stadelmann WK, Lentsch E, McMasters KM, Sunbelt Melanoma Trial G (2003) Sentinel lymph node biopsy for head and neck melanomas. Ann Surg Oncol 10:21–26 Chapman BC, Gleisner A, Kwak JJ, Hosokawa P, Paniccia A, Merkow JS, Koo PJ, Gajdos C, Pearlman NW, McCarter MD, Kounalakis N (2016) SPECT/CT improves detection of metastatic sentinel lymph nodes in patients with head and neck melanoma. Ann Surg Oncol 23:2652–2657 Creighton F, Bergmark R, Emerick K (2016) Drainage patterns to nontraditional nodal regions and level IIB in cutaneous head and neck malignancy. Otolaryngol Head Neck Surg (United States) 155:1005–1011 de Wilt JH, Thompson JF, Uren RF, Ka VS, Scolyer RA, McCarthy WH, O’Brien CJ, Quinn MJ, Shannon KF (2004) Correlation between preoperative lymphoscintigraphy and metastatic nodal disease sites in 362 patients with cutaneous melanomas of the head and neck. Ann Surg 239(4):544–52 Divi V, Fatt MA, Teknos TN, Mukherji SK (2005) Use of cross-sectional imaging in predicting surgical location of parotid neoplasms. J Comput Assist Tomogr 29:315–319 Durgut O, Basut O, Demir UL, Ozmen OA, Kasapoglu F, Coskun H (2013) Association between skin flap thickness and Frey’s syndrome in parotid surgery. Head Neck 35:1781–1786 Eberbach MA, Wahl RL, Argenta LC, Froelich J, Niederhuber JE (1987) Utility of lymphoscintigraphy

B. Gastman et al. in directing surgical therapy for melanomas of the head, neck, and upper thorax. Surgery 102:433–442 Eicher SA, Clayman GL, Myers JN, Gillenwater AM (2002) A prospective study of intraoperative lymphatic mapping for head and neck cutaneous melanoma. Arch Otolaryngol Head Neck Surg 128:241–246 Even-Sapir E, Lerman H, Lievshitz G, Khafif A, Fliss DM, Schwartz A, Gur E, Skornick Y, Schneebaum S (2003) Lymphoscintigraphy for sentinel node mapping using a hybrid SPECT/CT system. J Nucl Med 44:1413–1420 Faries MB (2018) Completing the dissection in melanoma: increasing decision precision. Ann Surg Oncol 25:585–587 Faries MB, Thompson JF, Cochran AJ, Andtbacka RH, Mozzillo N, Zager JS, Jahkola T, Bowles TL, Testori A, Beitsch PD, Hoekstra HJ, Moncrieff M, Ingvar C, Wouters M, Sabel MS, Levine EA, Agnese D, Henderson M, Dummer R, Rossi CR, Neves RI, Trocha SD, Wright F, Byrd DR, Matter M, Hsueh E, MackenzieRoss A, Johnson DB, Terheyden P, Berger AC, Huston TL, Wayne JD, Smithers BM, Neuman HB, Schneebaum S, Gershenwald JE, Ariyan CE, Desai DC, Jacobs L, McMasters KM, Gesierich A, Hersey P, Bines SD, Kane JM, Barth RJ, McKinnon G, Farma JM, Schultz E, Vidal-Sicart S, Hoefer RA, Lewis JM, Scheri R, Kelley MC, Nieweg OE, Noyes RD, Hoon DSB, Wang HJ, Elashoff DA, Elashoff RM (2017) Completion dissection or observation for sentinelnode metastasis in melanoma. N Engl J Med 376:2211–2222 Fincher TR, O’brien JC, McCarty TM, Fisher TL, Preskitt JT, Lieberman ZH, Stephens JF, Kuhn JA (2004) Patterns of drainage and recurrence following sentinel lymph node biopsy for cutaneous melanoma of the head and neck. Arch Otolaryngol Head Neck Surg 130:844–848 Garbe C, Buttner P, Bertz J, Burg G, D’hoedt B, Drepper H, Guggenmoos-holzmann I, Lechner W, Lippold A, Orfanos CE et al (1995) Primary cutaneous melanoma. Prognostic classification of anatomic location. Cancer 75:2492–2498 Gershenwald JE, Thompson W, Mansfield PF, Lee JE, Colome MI, Tseng CH, Lee JJ, Balch CM, Reintgen DS, Ross MI (1999) Multi-institutional melanoma lymphatic mapping experience: the prognostic value of sentinel lymph node status in 612 stage I or II melanoma patients. J Clin Oncol 17:976–983 Gregoire V, Ang K, Budach W, Grau C, Hamoir M, Langendijk JA, Lee A, Le QT, Maingon P, Nutting C, O’sullivan B, Porceddu SV, Lengele B (2014) Delineation of the neck node levels for head and neck tumors: a 2013 update. DAHANCA, EORTC, HKNPCSG, NCIC CTG, NCRI, RTOG, TROG consensus guidelines. Radiother Oncol 110:172–181 Hermanek P, Scheibe O, Spiessl B, Wagner G (1988) The TNM classification of malignant tumors: the new 1987 edition. Hautarzt 39:120 Hodi FS, Hwu W-J, Kefford R, Weber JS, Daud A, Hamid O, Patnaik A, Ribas A, Robert C, Gangadhar TC, Joshua AM, Hersey P, Dronca R, Joseph R, Hille D,

Neck Dissection and Parotidectomy for Melanoma Xue D, Li XN, Kang SP, Ebbinghaus S, Perrone A, Wolchok JD (2016) Evaluation of immune-related response criteria and RECIST v1.1 in patients with advanced melanoma treated with pembrolizumab. J Clin Oncol 34:1510–1517 Hyde N, Prvulovich E (2002) Is there a role for lymphoscintigraphy and sentinel node biopsy in the management of the regional lymphatics in mucosal squamous cell carcinoma of the head and neck? Eur J Nucl Med Mol Imaging 29:579–584 Jansen L, Nieweg OE, Kapteijn AE, Valdes Olmos RA, Muller SH, Hoefnagel CA, Kroon BB (2000) Reliability of lymphoscintigraphy in indicating the number of sentinel nodes in melanoma patients. Ann Surg Oncol 7:624–630 Klop WM, Veenstra HJ, Vermeeren L, Nieweg OE, Balm AJ, Lohuis PJ (2011) Assessment of lymphatic drainage patterns and implications for the extent of neck dissection in head and neck melanoma patients. J Surg Oncol 103:756–760 Knackstedt RW, Couto RA, Gastman B (2018) Indocyanine green fluorescence imaging with lymphoscintigraphy for sentinel node biopsy in head and neck melanoma. J Surg Res 228:77–83 Lachiewicz AM, Berwick M, Wiggins CL, Thomas NE (2008) Survival differences between patients with scalp or neck melanoma and those with melanoma of other sites in the Surveillance, Epidemiology, and End Results (SEER) program. Arch Dermatol 144:515–521 Leiter U, Stadler R, Mauch C, Hohenberger W, Brockmeyer N, Berking C, Sunderkotter C, Kaatz M, Schulte KW, Lehmann P, Vogt T, Ulrich J, Herbst R, Gehring W, Simon JC, Keim U, Martus P, Garbe C (2016) Complete lymph node dissection versus no dissection in patients with sentinel lymph node biopsy positive melanoma (DeCOG-SLT): a multicentre, randomised, phase 3 trial. Lancet Oncol 17:757–767 Leong SP, Achtem TA, Habib FA, Steinmetz I, Morita E, Allen RE, Kashani-Sabet M, Sagebiel R (1999) Discordancy between clinical predictions vs lymphoscintigraphic and intraoperative mapping of sentinel lymph node drainage of primary melanoma. Arch Dermatol 135:1472–1476 Lin D, Franc BL, Kashani-Sabet M, Singer MI (2006) Lymphatic drainage patterns of head and neck cutaneous melanoma observed on lymphoscintigraphy and sentinel lymph node biopsy. Head Neck 28:249–255 Macneill KN, Ghazarian D, McCready D, Rotstein L (2005) Sentinel lymph node biopsy for cutaneous melanoma of the head and neck. Ann Surg Oncol 12:726–732 McKean ME, Lee K, McGregor IA (1985) The distribution of lymph nodes in and around the parotid gland: an anatomical study. Br J Plas Surg 38:1–5 Morton DL, Wanek L, Nizze JA, Elashoff RM, Wong JH (1991) Improved long-term survival after lymphadenectomy of melanoma metastatic to regional nodes. Analysis of prognostic factors in 1134 patients from the John Wayne Cancer Clinic. Ann Surg 214:491–499; discussion 499–501

703 Morton DL, Wen DR, Foshag LJ, Essner R, Cochran A (1993) Intraoperative lymphatic mapping and selective cervical lymphadenectomy for early-stage melanomas of the head and neck. J Clin Oncol 11:1751–1756 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Elashoff R, Essner R, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Reintgen DS, Coventry BJ, Glass EC, Wang HJ, Group M (2006) Sentinel-node biopsy or nodal observation in melanoma. N Engl J Med 355:1307–1317 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Puleo CA, Coventry BJ, Kashani-sabet M, Smithers BM, Paul E, Kraybill WG, Mckinnon JG, Wang HJ, Elashoff R, Faries MB, GROUP M (2014) Final trial report of sentinel-node biopsy versus nodal observation in melanoma. N Engl J Med 370: 599–609 O’brien CJ, Uren RF, Thompson JF, Howman-Giles RB, Petersen-Schaefer K, Shaw HM, Quinn MJ, McCarthy WH (1995) Prediction of potential metastatic sites in cutaneous head and neck melanoma using lymphoscintigraphy. Am J Surg 170:461–466 Ollila DW, Foshag LJ, Essner R, Stern SL, Morton DL (1999) Parotid region lymphatic mapping and sentinel lymphadenectomy for cutaneous melanoma. Ann Surg Oncol 6:150–154 Raigani S, Cohen S, Boland GM (2017) The role of surgery for melanoma in an era of effective systemic therapy. Curr Oncol Rep 19:17 Reynolds HM, Smith NP, Uren RF, Thompson JF, Dunbar PR (2009) Three-dimensional visualization of skin lymphatic drainage patterns of the head and neck. Head Neck 31(10):1316–25 Robbins KT, Clayman G, Levine PA, Medina J, Sessions R, Shaha A, Som P, Wolf GT, American Head and Neck Society and the American Academy of Otolaryngology-Head and Neck Surgery (2002) Neck dissection classification update: revisions proposed by the American Head and Neck Society and the American Academy of Otolaryngology-Head and Neck Surgery. Arch Otolaryngol Head Neck Surg 128:751–758 Samra S, Sawh-Martinez R, Tom L, Colebunders B, Salameh B, Truini C, Ariyan S, Narayan D (2012) A targeted approach to sentinel lymph node biopsies in the parotid region for head and neck melanomas. Ann Plast Surg 69:415–417 Scoggins CR, Martin RC, Ross MI, Edwards MJ, Reintgen DS, Urist MM, Gershenwald JE, Sussman JJ, Dirk Noyes R, Goydos JS, Beitsch PD, Ariyan S, Stromberg AJ, Hagendoorn LJ, McMasters KM (2010) Factors associated with false-negative sentinel lymph node biopsy in melanoma patients. Ann Surg Oncol 17:709–717 Shah JP, Strong E, Spiro RH, Vikram B (1981) Surgical grand rounds Neck dissection: current status and future possibilities. Clin Bull 11:25–33 Shah JP, Kraus DH, Dubner S, Sarkar S (1991) Patterns of regional lymph node metastases from cutaneous melanomas of the head and neck. Am J Surg 162:320–323

704 Subramanian S, Chiesa F, Lyubaev V, Aidarbekova A, Brzhezovskiy V (2007) The evolution of surgery in the management of neck metastases. Acta Otorhinolaryngol Ital 27:309–316 Supriya M, Narasimhan V, Henderson MA, Sizeland A (2014) Managing regional metastasis in patients with cutaneous head and neck melanoma – is selective neck dissection appropriate? Am J Otolaryngol Head Neck Med Surg 35:610–616 Suton P, Luksic I, Muller D, Virag M (2012) Lymphatic drainage patterns of head and neck cutaneous melanoma: does primary melanoma site correlate with anatomic distribution of pathologically involved lymph nodes? Int J Oral Maxillofac Surg 41:413–420 Teltzrow T, Osinga J, Schwipper V (2007) Reliability of sentinel lymph-node extirpation as a diagnostic method for malignant melanoma of the head and neck region. Int J Oral Maxillofac Surg 36:481–487 Uren RF (2004) Lymphatic drainage of the skin. Ann Surg Oncol 11:179S–185S

B. Gastman et al. Van Der Ploeg IMC, Valdés Olmos RA, Kroon BBR, Wouters MWJM, Van den brekel MWM, Vogel WV, Hoefnagel CA, Nieweg OE (2009) The yield of SPECT/CT for anatomical lymphatic mapping in patients with melanoma. Ann Surg Oncol 16:1537–1542 Wells KE, Cruse CW, Daniels S, Berman C, Norman J, Reintgen DS (1994) The use of lymphoscintigraphy in melanoma of the head and neck. Plast Reconstr Surg 93:757–761 Wertz AP, Durham AB, Malloy KM, Johnson TM, Bradford CR, Mclean SA (2017) Total versus superficial parotidectomy for stage III melanoma. Head Neck 39:1665–1670 Zender C, Guo T, Weng C, Faulhaber P, Rezaee R (2014) Utility of SPECT/CT for periparotid sentinel lymph node mapping in the surgical management of head and neck melanoma. Am J Otolaryngol Head Neck Med Surg 35:12–18

Local and Recurrent Regional Metastases of Melanoma Matthew C. Perez, Kenneth K. Tanabe, Charlotte E. Ariyan, John T. Miura, Dorotea Mutabdzic, Jeffrey M. Farma, and Jonathan S. Zager

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 706 Local and Regional Recurrence of Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Local Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . In-transit Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Regional Nodal Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

707 707 708 709

Hyperthermic Isolated Limb Perfusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . History and Early Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Patient Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Preoperative Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

710 710 711 712 712 712

M. C. Perez · J. T. Miura · J. S. Zager (*) Department of Cutaneous Oncology, Moffitt Cancer Center, Tampa, FL, USA e-mail: [email protected]; john.miura@moffitt.org; jonathan.zager@moffitt.org K. K. Tanabe Division of Surgical Oncology, Department of Surgery, Massachusetts General Hospital Cancer Center, Boston, MA, USA e-mail: [email protected] C. E. Ariyan Memorial Sloan-Kettering Cancer Center, New York, NY, USA e-mail: [email protected] D. Mutabdzic · J. M. Farma Department of Surgical Oncology, Fox Chase Cancer Center, Philadelphia, PA, USA e-mail: [email protected]; [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_24

705

706

M. C. Perez et al. Leak Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hyperthermia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Specific Toxicities and Management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

714 715 716 717 718

Isolated Limb Infusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Patient Selection and Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Response to Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Survival After ILI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Burden of Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

719 719 719 720 720 721 721 721

Intralesional Therapies for Cutaneous Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bacille Calmette-Guerin (BCG) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interleukin-2 (IL-2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Granulocyte Macrophage-Colony Stimulating Factor (GM-CSF) . . . . . . . . . . . . . . . . . . . . . Velimogene Aliplasmid (Allovectin) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Talimogene Laherparepvec (T-VEC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rose Bengal (PV-10) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Daromun (L19IL2 + L19TNF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coxackievirus A21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Combination with Systemic Immune Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electrochemotherapy (ECT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

722 722 722 723 724 724 724 725 725 725 728 728 728

Neoadjuvant Therapy for Borderline Resectable Nodal Metastasis . . . . . . . . . . . . . . . 729 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731

Abstract

Up to 10% of patients with cutaneous melanoma will develop recurrent locoregional disease. While surgical resection remains the mainstay of treatment for isolated recurrences, locoregional melanoma can often present as bulky, unresectable disease and can pose a significant therapeutic challenge. This chapter focuses on the natural history of local and regionally recurrent metastases and the multiple treatment modalities which exist for advanced locoregional melanoma, including regional perfusion procedures such as hyperthermic isolated limb perfusion and isolated limb infusion, intralesional therapies, and neo-adjuvant systemic therapy strategies for borderline resectable regional disease. Hyperthermic limb perfusion (HILP) and isolated limb infusion (ILI) are generally well-tolerated and have shown overall response rates between 44% and 90%. Intralesional therapies also

appear to be well-tolerated as adverse events are usually limited to the site of injection and minor transient flu-like symptoms. Systemic targeted therapies have shown to have response rates up to 85% when used as neoadjuvant therapy in patients with borderline resectable disease. While combination immunotherapy in the neoadjuvant setting has also shown promising results, this data has not yet matured.

Introduction The incidence of cutaneous melanoma continues to steadily increase and accounts for the majority of skin cancer-associated deaths (Siegel et al. 2017). Over 80% of melanoma patients present with clinically localized disease and can be managed with resection of the primary tumor with or without sentinel lymph node biopsy (Balch et al. 2009). However, up to

Local and Recurrent Regional Metastases of Melanoma

707

Local and Regional Recurrence of Melanoma Local Recurrence

Image 1 In-transit melanoma of the upper extremity

10% of patients will develop recurrent locoregional disease (Borgstein et al. 1999). For isolated and resectable local and regional recurrences, complete excision of disease is the mainstay of therapy. However, patients often present with advanced, bulky, and unresectable disease either locally at the primary tumor site, in the regional nodal basin(s), or within the dermal lymphatic channels between the primary site and regional lymph nodes, and can pose a significant therapeutic challenge, as shown in Image 1 (Pawlik et al. 2005). Locoregionally advanced melanoma has also been shown to have significant adverse effects on patient’s quality of life and emotional wellbeing (Weitman et al. 2018). The last decade has seen a surge in the adoption of new immune and targeted therapies which have improved survival for patients with metastatic melanoma (Robert et al. 2015a, b; Chapman et al. 2011; Hodi et al. 2010). As a result of this prolonged survival, patients are living longer with advanced disease and therefore the importance of locoregional control has also increased. Fortunately, multiple management modalities exist for patients with advanced locoregional disease. This chapter focuses on the regional perfusion procedures, hyperthermic isolated limb perfusion (HILP) and isolated limb infusion (ILI), intralesional therapies, and neoadjuvant systemic therapy strategies for borderline resectable nodal metastases.

A local recurrence is a regrowth of melanoma in close proximity to the anatomic site from which the primary tumor was excised. Contemporary studies define this as regrowth within 2 cm of the surgical scar (Balch et al. 1993). This can be a result of either incomplete excision of the primary tumor or intralymphatic metastases and microscopic satellite lesions which are not contiguous but rather within the immediate area of the primary lesion. It is worth noting that although commonly used, a definition involving distance from the primary excision scar can lead to inconsistencies since primary resection margins vary between 1 and 2 cm, depending on the depth of the primary tumor. Several features of the primary tumor and lymphatic metastases have been examined as potential prognostic factors that may be predictive of melanoma local recurrence. Long-term results of the Intergroup Melanoma Trial designed to evaluate 2 cm versus 4 cm surgical margins demonstrated a higher risk for local recurrence in patients with thicker primary melanomas (Karakousis et al. 1996; Balch et al. 2001). The risk of melanoma local recurrence increased with tumor thickness, despite the wider margins employed as tumor thickness increased (Table 1). The presence of ulceration in the primary melanoma is also an important risk factor for local recurrence, as evidenced by results of the Intergroup Melanoma Trial, as the presence or absence of ulceration had the greatest impact on risk for local recurrence (Balch et al. 2001). In the randomized group of patients with lesions on the trunk and proximal extremity, there was a sixfold increase in local recurrence rates (at any time) in patients with ulcerated primary melanomas, with local recurrence rates of 6.6% in patients with ulcerated primary melanomas compared with 1.1% in patients with nonulcerated primary lesions. In the nonrandomized group with lesions involving the distal extremity and head and neck sites, the local recurrence rate was 16.2% (at any time) in patients with ulcerated lesions compared to 2.1% in patients with

708

M. C. Perez et al.

Table 1 Frequency of subgroups of thickness and incidence of local recurrence and other metastases Thickness (mm) 1.00–2.0 2.01–3.0 3.01–4.0

No. of patients 445 (60%) 215 (29%) 77 (10%)

Total

737 (100%)

Local recurrence (%) 2.3 4.2 11.7 ( p = 0.001) 3.8

In-transit metastases (%) 3.6 8.4 16.9 ( p = 0.001) 6.4

Regional nodal recurrence (%) 9.2 15.8 29.9 ( p = 0.001) 13.3

Distant recurrence (%) 14.4 27.9 44.2 ( p = 0.001) 21.4

From Karakousis et al. (1996)

nonulcerated lesions (p < 0.001) (Balch et al. 2001). Resection margin size has not been shown to correlate with local recurrence as several prospective multi-institutional and large retrospective studies have shown that narrow margins for resection of thin melanomas result in similar rates of local recurrence. Furthermore, for tumors between 1 and 2 mm thick, a recent large retrospective study showed no significant difference in local recurrence or survival with 1 cm resection margins when compared to 2 cm margins (Doepker et al. 2016). Surgery remains the mainstay of treatment for local recurrences of melanoma. In patients who are examined at regular intervals, local recurrences are may be detected at a stage at which they can be managed by surgical resection. Few data exist on appropriate surgical margins to employ during resection of a local recurrence. In the absence of significant data, most experts recommend excision using approximately 1 cm margins to ensure that a grossly clear margin is achieved. Clearly, radical surgery that makes use of extensive margins during surgical extirpation of a local recurrence is not justified. But even when more conservative margins are employed to excise a local recurrence, skin grafts are frequently required to close the resulting defect in an operative field that has already been compromised by a prior wide excision.

In-transit Recurrence In-transit disease represents the clinical manifestation of small tumor emboli trapped within the dermal and subdermal lymphatics between the site of the primary tumor and regional lymph node drainage basin(s). Cascinelli et al. reported

a 13% incidence of recurrent in-transit melanoma observed in a cohort of more than 1500 patients who had clinical stage I and II melanoma (Cascinelli et al. 1986). Dalal et al. reported a 4.8% incidence of in-transit metastases as the first site of recurrence in a cohort of 1046 patients with stage I or II disease at a median follow-up of 36 months (Dalal et al. 2007). A number of factors may predispose patients to subsequent in-transit recurrence after resection of their primary tumors. In a review by Cascinelli et al., patients who had positive nodes had a higher incidence of recurrence than those who had negative nodes. Furthermore, in patients who had one positive node, the incidence of subsequent intransit disease was 11%, whereas in those who had three or more positive nodes, the incidence of subsequent in-transit recurrence as the initial site of failure was 31% (Cascinelli et al. 1986). Calabro et al. reviewed 1001 patients who had positive nodes, observing metastatic in-transit melanoma as the initial site of recurrence in 99 of them (10%) (Calabro et al. 1990). Gershenwald et al. reported a 3.3% incidence of in-transit metastases as the initial site of recurrence among 243 patients with stage I or II melanoma and a negative result on SLN biopsy that were followed for a median of 24 months (Gershenwald et al. 1998). Essner et al. reported a 3.7% incidence of in-transit metastases among 267 patients with clinical stage I to II melanoma who underwent SLN biopsy and were followed for a median of 45 months. For patients who had a negative result on SLN biopsy (n = 225), the incidence of in-transit metastases was 2.7% (n = 6), whereas for patients who had a positive result on SLN biopsy (n = 42), the incidence was 9.5% (n = 4) (Essner et al. 1999). Dalal et al. reported an increase in intransit

Local and Recurrent Regional Metastases of Melanoma

709

recurrences in SLN-positive patients compared with SLN-negative patients (30% versus 21%, respectively; p = NS). Factors predisposing to recurrence in this cohort of 1046 patients with stage I or II melanoma were a positive SLN, ulceration, thickness, and stage II disease (Dalal et al. 2007). The median time to the appearance of in-transit metastatic melanoma is fairly consistent, ranging from 13 to 24 months (Dalal et al. 2007; Lee 1980; McCarthy et al. 1988; Wong et al. 1990). SLNpositive patients appear to have an earlier recurrence at a median of 13 months, compared with SLN-negative patients who have a recurrence at a median of 24 months (Dalal et al. 2007).

recurrence in that series included a history of prior lymph node biopsy, the clinical status of the nodes, or the site of regional lymphadenectomy. The highest reported rate of regional nodal recurrence after lymphadenectomy was from a series by Monsour et al. Of 48 patients undergoing therapeutic lymphadenectomy, 52% developed regional nodal recurrence as their initial site of failure. In that review, nodal relapse was age dependent as 31% in patients less than 50 years old and 66% in patients more than 50 years of age recurred. The authors did not comment on the factors that may have predisposed their patients to such a high rate of regional nodal recurrence (Monsour et al. 1993). The widespread use of SLN biopsy has led to a new clinical scenario: nodal recurrence after both negative and positive SLN biopsy. Gershenwald et al. looked at 602 melanoma patients who underwent successful lymphatic mapping and SLN biopsy, followed for a median of 30 months. Of these patients, 105 (17%) had at least one histologically positive SLN, of whom 101 (96%) underwent completion lymphadenectomy. Of 36 patients who had recurrences, 10 had recurrences in the nodal basin at a median time of 14 months. The nodal basin was not the sole site of recurrence in any of these 10 patients. Reexamination of the sentinel node in patients who had nodal failure after a negative result on SLN biopsy revealed missed micrometastatic disease in eight (Gershenwald et al. 2000). Nodal basin recurrence seems to be the result of aggressive disease biology and not surgical manipulation. Clary et al., at Memorial Sloan-Kettering Cancer Center, reported on the pattern of recurrence among 332 consecutive patients with localized primary cutaneous melanoma who underwent SLN biopsy. The overall incidence of recurrence was greater in the patients who had positive nodes (40% versus 14%), although the distribution between locoregional and systemic recurrences was not statistically different (Clary et al. 2001). Locoregional recurrences constituted 61% of all first-site recurrences, and distant recurrences accounted for 39%. As in Gershenwald’s study, a re-examination of a sentinel node initially thought to be negative in 11 patients who had subsequent nodal

Regional Nodal Recurrence The incidence of regional nodal recurrence after lymphadenectomy is variable and depends on a number of factors, primarily related to the tumor burden in the lymph node basin originally dissected. To a lesser extent, the likelihood of regional nodal recurrence depends on the extent of the procedure performed. Miller et al. found a 12% rate of nodal relapse at the site of 207 patients who underwent prior lymphadenectomy and noted relapse in a regional nodal basin after removal of negative nodes in 6.7% of patients. When one to three nodes were positive, 14% of patients subsequently suffered a relapse in the regional nodal basin; when more than four nodes were positive at the initial dissection, 53% of patients subsequently had a relapse in the regional nodal basin (Miller et al. 1992). Warso and Das Gupta found a 5.6% rate of nodal recurrence among 1030 patients undergoing prior lymphadenectomy (Warso and Das Gupta 1994). Regional nodal failure was often a harbinger of systemic relapse, because fewer than half of those patients had recurrences confined to the nodes alone. Calabro et al. reported 162 nodal relapses among 1001 patients undergoing lymphadenectomy with positive nodes (16%). Factors that predicted nodal basin relapse included the number of positive lymph nodes and the presence of extranodal tumor extension (Calabro et al. 1989). Of note, factors that did not predict nodal

710

recurrence detected metastatic disease in seven (64%). In the MSLT-1 trial, 59 (6.3%) of the 944 patients with tumor-negative SLNs developed regional nodal recurrence at a median follow-up of 54 months (Morton et al. 2006). Of the 59 patients, 48 (81%) had recurrence in the SLN drainage basin and 11 had recurrence in a basin that was not sampled. More recently, Zogakis et al. reported on 773 patients with negative SLNs, of whom 8.9% (n = 69) had a recurrence. Distant metastases were seen in 4.8%: 1.8% were in-transit, 1.7% were nodal, and 1.2% were local recurrences (Zogakis et al. 2007). For low-burden in-transit and recurrent regional nodal disease, complete resection of disease is often preferred when feasible. For unresectable and locoregionally advanced disease, multiple treatment modalities currently exist such as regional chemotherapy and intralesional therapies, as well as systemic immune and targeted therapies, and are discussed in more detail later in this chapter.

Hyperthermic Isolated Limb Perfusion History and Early Clinical Studies In 1956, the Department of Surgery at University of Tulane embarked on regional perfusion studies to increase the uptake of chemotherapy in tumors located in regions of the body whose vascular supply and drainage could be completely isolated (Krementz et al. 1994). Use of a heart-lung machine to support isolated hyperthermic perfusion of the tumor was evaluated in dogs as a strategy to avoid systemic toxic effects and at the same time increase the dose of nitrogen mustard (Ryan et al. 1958). The concept of cannulation of both the arterial inflow and the venous drainage for connection to an extracorporeal oxygenated circuit maintained by a heatlung machine represented an improvement over the technique previously described by Kopp and colleagues in which the chemotherapy was administered into the artery, with the venous drainage left unaltered or clamped (Ryan et al. 1958).

M. C. Perez et al.

In June 1957, a patient with a very high burden of melanoma metastases to the extremity presented to Charity Hospital, 2 years after having been treated for a melanoma of the right ankle. Despite having over 80 satellite lesions, the patient refused the standard therapy at that time, namely, amputation. The team performed an isolated chemotherapy perfusion using melphalan, a chemotherapy agent that was new and under evaluation at the time for metastatic melanoma. The patient experienced a complete clinical response and remained melanoma-free until his death at age 92, some 16 years later. In 1958, Creech presented the results of isolated perfusion in six patients with melanoma and another 18 with other advanced cancers before the American Surgical Association in New York (Creech et al. 1958). For pelvic tumors, the aorta and IVC were occluded below their renal branches and cannulated just above the bifurcation. For perfusion of lung tumors, use of two circuits and caval occlusion were used to prevent mixing between the systemic and pulmonic circuits. And in cases, in which tourniquets could not be applied (e.g., breast), a motor pump was used to create negative pressure in the venous return circuit to minimize systemic mixing. They reported gross or microscopic responses in 18 of 19 cases followed long enough for changes to be evident. And by 1962, they had treated a sufficiently large number of patients to report results of 303 patients, 123 with melanomas (Krementz et al. 1962). Over the ensuing four decades, many hospitals started performing isolated limb perfusion and reporting their results. Unfortunately, an opportunity for progress was lost during this interval because this experience lacked scientific rigor. The studies were generally single arm, absent appropriate control groups, and involved heterogenous patient populations including patients with completely resected tumors and unresectable tumors. Treatment schedules were varied in their dose of melphalan, perfusion duration, and perfusion temperature (Rosin and Westbury 1980; Di Filippo et al. 1989; Jonsson et al. 1983; Lejeune et al. 1983; Skene et al. 1990; Minor et al. 1985; Kroon 1988; Klaase et al. 1994a; Kroon et al.

Local and Recurrent Regional Metastases of Melanoma

711

1993). For example, in a report of 1139 perfusions performed over 35 years, the authors included patients in need of definitive treatment of in-transit metastases, unresectable recurrent or primary tumors, adjunctive therapy to surgical excision for regionally confined melanoma, conversion of advanced unresectable melanoma to resectable, and palliation in noncurable recurrent melanomas by maintaining a functional limb in the presence of systemic metastases (Krementz et al. 1994). However, clinical studies in the past two decades have been of significantly higher quality and with greater scientific rigor. Results of these studies are discussed below. See also chapter ▶ “Hyperthermic Regional Perfusion for Melanoma of the Limbs.”

systemic therapy before resorting to HILP. BRAF V600 mutant melanomas are commonly sensitive to targeted therapy using a BRAF inhibitor combined with a MEK inhibitor, with a response rate of 63% and acceptable toxicity (Flaherty et al. 2012a, b). And for patients without BRAF V600 mutations in their melanoma, immune checkpoint inhibitor therapy to block CTLA-4, PD1, or PDL1 is commonly used. Response rates range from 11% with Ipilimumab to 61% with Ipilimumab and nivolumab (Wolchok et al. 2017). Combined BRAF and MEK inhibitor therapy is typically first-line treatment for unresectable in-transit metastases that are BRAF mutant. And immune checkpoint inhibitor immunotherapy is typically first-line treatment for unresectable in-transit metastases that are wild type. HILP is considered for patients who progress on these therapies. And it is a good approach for patients who have a contraindication to immunotherapy, such as liver transplant, active colitis, and/or unmanageable and severe toxicity to immunotherapy. Adjuvant HILP was once accepted as appropriate adjuvant treatment for high-risk melanomas. A prospective randomized control trial was conducted at the University of Cologne randomized to excision alone or excision with HILP. This trial was stopped early because interim analysis showed a remarkable reduction in recurrences in the HILP arm (Ghussen et al. 1988). But subsequently conducted randomize control trials that are of higher quality and larger patient number have convincingly demonstrated lack of benefit of adjuvant HILP. Thus, the results of the Cologne study – positive and in favor of HILP – are generally discounted and considered an unreliable outlier based on its very small sample size (e.g., 34 patients treated with HILP). The clinical trial considered definitive in this area was conducted by a consortium of EORTC, WHO, and the North American Perfusion Group (NAPG-1) (Koops et al. 1998). Over a period of 10 years, 852 patients were randomized to wide excision alone or wide excision and HILP. HILP-treated patients benefitted from a reduction in incidence of intransit metastases as first site of recurrence (reduced from 6.6% to 3.3%), and of regional lymph node metastases, with a reduction from

Patient Selection Patients with metastatic melanoma confined to an extremity without evidence of distant metastases should be considered for hyperthermic isolated limb perfusion (HILP). The melanoma should be considered unresectable, although the definition of unresectable is subjective. The most common indication for limb perfusion is in-transit metastases, and the frequency and timing of in-transit metastases as well as the number and distribution of metastases are used to define when resection is appropriate. Rapid recurrence of multiple in tumor nodules soon after excision of in-transit metastases indicates that further surgical resection is not warranted. Full staging including PET-CT and head MRI to exclude other metastases should be performed. Patients with peripheral vascular disease are not good candidates for HILP because of a significantly higher risk for toxicity and complications. The presence of peripheral vascular disease is typically evident on preoperative evaluation (see below). The role of HILP has shifted over the years. Prior to effective molecularly targeted immunotherapies, HILP was accepted as the most effective and appropriate treatment for patients with metastases or local recurrence confined to an extremity. However, with advent of effective systemic therapies, most patients are treated with

712

16.7% to 12.6%. But importantly, HILP-treated patients experienced no benefit in overall survival or time to distant metastasis. Adjuvant HILP was also examined as adjuvant to excision of in-transit metastases, and similar to other adjuvant trial results, improvement in regional disease control could be demonstrated but not improvement in overall survival (Hafstrom et al. 1991). In summary, HILP is not beneficial as an adjuvant therapy.

Preoperative Evaluation Preoperative evaluation requires careful staging to exclude metastases outside the limb, determination of preoperative ambulatory status, comorbidities, significant peripheral vascular disease, and patient motivation to handle side effects. Patients whose functional status has declined to a point where they are no longer ambulatory are poor candidates for HILP. Measurement of any preexisting leg edema to establish a baseline is appropriate. Careful evaluation for regional nodal metastases by CT or PET-CT establishes whether elective concomitant lymphadenectomy is required. In cases where the presence of significant peripheral vascular disease is detected on clinical examination, pulse volume recordings (PVR) are a useful noninvasive evaluation to determine the locations and severity of disease, and to establish a baseline. Patients with moderate or severe peripheral vascular disease are at high risk for complications from isolated limb perfusion and require alternative approaches for management of their melanoma. In situations where physical examination and PVR are insufficient to accurately assess the severity of peripheral vascular disease, preoperative angiography or CT angiography is indicated. If drug dosing will be determined based on volume of the limb, several measurement techniques for limb volume are available. One involves physical measurements of the circumference of the leg at 2 cm intervals to calculate the cross-sectional area, and then calculation of the integral of this function. Another technique uses a water displacement. The leg or arm is immersed into a container filled

M. C. Perez et al.

completely to a brim, and the volume of displaced, overflowing water is the volume of the immersed extremity (Rabe et al. 2010). The last involves use of a CT or MRI scan of the entire extremity and use of 3D analytic software to measure leg volume (Brys et al. 2016).

Equipment The operation requires a standard heart-lung bypass device equipped with a roller pump, oxygenator with a gas source (95% oxygen 5% carbon dioxide), heater capable of reaching 42  C and venous reservoir. It is helpful, though not required, to have a machine for activated clotting time measurements in the operating room. A scintillation probe mounted over the chest (precordial) is used to monitor for I-131 or 99 m-Technetium labeled albumin or red cells as an indicator of leak from the circuit into the systemic circulation. An ultraviolet (black) light is used to evaluate for leakage of fluorescein from the extremity. A pulse volume recording machine is used before, during and after the operation. Heating blankets warmed by a heated water circulator are used for external warming of the extremity. Thermistors inserted under the skin are connected to digital temperature monitors to monitor temperature in different locations during the operation. Standard vascular instruments are used during the operation, as well as rummel tourniquets, a hand drill for placement of Steinmann pins, and a Doppler probe. A selection of different size arterial and venous cannulas should be on hand, as well as heparin-saline irrigation. A self-retaining retractor attached to the table is of significant help for approaching iliac vessels.

Operation Isolated limb perfusion is a technically complex operation that requires closely integrated teamwork by a multidisciplinary team including anesthesiologists, perfusionists, nuclear radiologists, pharmacists, nurses, and surgeons. The quality and frequency of communication among these

Local and Recurrent Regional Metastases of Melanoma

713

team members affects outcomes. The procedure involves the use of an extracorporeal circuit attached to a heart-lung machine (oxygenator and blood pump) to heat the circulating blood, increase the oxygen tensions before delivery to the isolated limb and buffer with carbon dioxide. Anesthesia must be prepared for intraoperative fluid shifts between the vascular compartments of the limb and the remainder of the body, hypotension caused by low vascular tone, and sequel of ischemia-reperfusion (Ruschulte et al. 2013). The operation is conducted under general anesthesia. Preparation for an operation of 4–6 h duration is appropriate, depending on which vessels require isolation and whether a concomitant lymphadenectomy is indicated. Anesthesia should be prepared for acute blood loss, particularly if surgical isolation of the vessels is anticipated to be difficult (e.g., iliac vessels, scarred vessels), and the operation should be conducted with two large-bore peripheral IVs. Central venous pressure monitoring is not typically required. An active type and crossmatch in the blood bank is mandatory. An arterial line is useful for repeated activated clotting time (ACT) measurements, and on occasion, close monitoring of blood pressure to enable manipulations necessary to manage leakage between the circuit and systemic circulation. A bladder catheter should be inserted. An oral-gastric tube may be used during the operation. An epidural catheter for post-operative pain management is not typically used. A dose of prophylactic antibiotics is administered prior to the skin incision. After induction of anesthesia, PVR is measured and saved for comparison after the operation. Similarly, peripheral pulses in the affected extremity are carefully assessed and recorded. It is useful to monitor temperatures of the extremity in several locations during the operation. Thermistors are placed in the proximal and distal extremity both medially and laterally (e.g., four thermistors) for real time temperature monitoring. The extremity is then wrapped in heating blankets, leaving the PVR cuff in place. The prep and drape should be wide. It is necessary to place sterile surgical tubing (or esmark bandage) around the root of the extremity for later use as a tourniquet.

The general approach is to use an incision over the vessels, with extension if needed for a lymphadenectomy. Axillary lymphadenectomy and iliac/hypogastric lymphadenectomy are performed as a matter of routine during isolated limb perfusion through the axillary or external iliac vessels, respectively. However, superficial femoral lymphadenectomy is performed at time of isolated limb perfusion only when clinical evidence of nodal metastases is present given that the incision used for this lymphadenectomy has high likelihood of infection or dehiscence, especially in a chemotherapy-treated field. Moreover, perfusion from an iliac approach does effectively treat nodes in the femoral triangle (Koops et al. 1998). The vessels are circumferentially isolated and small collateral vessels distal to the cannulation sites are tied off. A Steinmann pin is placed into the anterior superior iliac spine to serve as a cleat and prevent slippage of the tourniquet around the root of the extremity. Once the dissection is complete, 350 U/kg heparin is administered to achieve an ACT of over 450 sec. The vessels are occluded proximally and distally with either vascular clamps or Rummel tourniquets. The vein and artery are cannulated through transversely oriented incisions in the vessels, and each held in place with a Rummel tourniquet placed around the distal vessel and cannula, taking care to avoid fracturing any atherosclerotic plaque that is present. The tourniquet around the root of the limb is tightened maximally. After confirmation of a therapeutic ACT (typically >450 s), the cannulas are connected to the extracorporeal circuit and the roller pump is gradually brought up to the maximum flow rate at which the reservoir volume does not diminish. In rare patients, typically those that start with a very low hematocrit that also have very small limb volumes and consequently larger hemodilution from the priming volume, the hematocrit in the circuit is unacceptably low (e.g., below 18%). In these cases, a portion of a unit of packed red blood cells is transfused into the circuit. Heparin resistance – defined by the inability to achieve therapeutic ACT with typical heparin doses – is typically successfully treated with additional heparin. However, anti-thrombin III deficiency should be suspected if this maneuver

714

is unsuccessful. In these situations, options include changing to argatroban, or transfusion of fresh frozen plasma or antithrombin (Spiess 2008). Once the extremity has reached the target temperature, melphalan is administered into the arterial side of the circuit based on the planned dose schedule. The heater for the heart-lung machine is adjusted based on the extremity temperatures registered by the thermistors. Isolated perfusion is conducted for the planned time, typically 60 or 90 min, during which time leak monitoring is employed to guide any necessary adjustments (see below). Protocols for drug dosage, drug administration schedule, target temperature, and duration of perfusion differ amongst centers. After the perfusion is complete, the extremity is rinsed with crystalloid and/or colloid, with the drugcontaining venous effluent discarded. The cannulas are removed, and the arteriotomy and venotomy are repaired with fine sutures. PVR measurements in the distal extremity are obtained and upon confirmation of a return to baseline, protamine is administered to reverse the effects of heparin. The wound is closed in multiple layers. Because in-transit metastases occur most commonly in the lower extremity, access for HILP is most commonly achieved via the iliac vessels or the femoral vessels. If in-transit metastases are located within 6 inches of the inguinal crease, perfusion via the iliac vessels is required to achieve perfusion of the proximal thigh. The surgical approach to the iliac vessels starts with an oblique incision in the lower abdominal wall. The external oblique fascia is incised and the internal oblique musculature is separated to reveal the transversalis fascia. This is incised and the abdominal contents are retracted superomedially to expose the iliac vessels. The external iliac and obturator nodes are removed. Note is made of the quality and characteristics of the Doppler signals in the external iliac artery and vein. The hypogastric vein is ligated in situ (not necessary to divide) and a bulldog clip is placed on the hypogastric artery. The external iliac vessels are followed under the inguinal ligament for as far as possible to allow for identification of small branches, which are clipped to prevent collateral flow.

M. C. Perez et al.

Removal of the clips on arterial branches at completion of the operation improves blood flow to portions of the healing wound. A drill is used to place a Steinmann pin in the anterior superior iliac spine to hold the tourniquet in place. For approach to the axillary artery and vein, a generous incision is made in the axilla and flaps are raised to allow a complete axillary lymphadenectomy. The pectoralis minor muscle is divided inferior to its insertion onto the coracoid to allow removal of level III axillary nods and provide additional exposure of the axillary vessels. Branches are tied off and divided. The brachial plexus trunks are carefully pushed aside to provide exposure to the artery with minimal disruption to the nerves. A Steinmann pin is placed to serve as a cleat for the tourniquet. An alternative approach is to use a retractor connected to the table to hold the tourniquet in place (Stamatiou et al. 2017).

Leak Monitoring During isolated limb perfusion it is necessary to assess for ongoing leakage from the circuit into the systemic circulation, or from the systemic circulation into the circuit. The former condition leads to systemic exposure to the drug, and the latter condition leads to lowering of the drug concentration in the perfusion circuit. One commonly used technique to measure leak involves mounting a shielded precordial scintillation detector over the precordium and injecting isotope labelled albumin or red cells into the perfusion circuit. I-131 and Tm-99 are used most commonly. A fraction of the total dose is administered into the systemic circulation to calibrate the system and allow for quantification of the leak, using the assumption that the volumes of the extracorporeal circuit and the systemic vasculature are in the proportion of 1:5. This technique allows determination of the percent fractional leak as a function of time. An alternative approach involves administration of fluorescein into the circuit and then viewing different areas of the body and collected urine with a Woods lamp. This technique can reveal specific spots of skin outside the extremity

Local and Recurrent Regional Metastases of Melanoma

715

that are receiving perfusate, and lead to identification of specific collateral vessels to be tied off. Fluorescein in the urine collection chamber is easily identified with the black light, and increase in intensity over time provides a qualitative sense of the rate of leak. A disadvantage of this technique is that quantification is not possible, and once a significant systemic leak has occurred, it is not possible to confirm correction of the leak. Another technique that has been described but not used widely is administration of 3% desflurane into the bypass circuit using an anesthetic vaporizer. The expired breath is then monitored by standard gas analysis for desflurane as a sign of leakage (Stanley et al. 2000). Leakage of significant amounts of melphalan into the systemic circulation can lead to acute nausea and delayed bone marrow suppression or hair loss. Leak from the circuit into the systemic circulation that occurs later during the perfusion is of less consequence, since most of the melphalan in the extremity will have been taken up by tissue. Leakage of even small amounts of tumor necrosis factor leads to proinflammatory cytokine storm responsible for sepsis-like side effects including intraoperative tachycardia, hypotension and pulmonary edema (Laurenzi et al. 2004). Specific maneuvers are employed to manage leakage between the circuit and systemic circulation (Table 1). Leakage from the circuit into the systemic circulation typically results in loss of volume in the venous reservoir. The leakage may be through venous collaterals, arterial collaterals, or both. Leakage from the extracorporeal circuit that occurs after drug is administered results in systemic exposure to drug, and a lower concentration in the limb. The first step is to lower the flow rate, which results in reduced pressure in collateral arteries and veins. The operating table can be tilted into reverse Tredelenberg position to lower the venous pressure in the leg relative to the IVC. After infusion of fluorescein into the circuit, the skin should be examined with a Woods lamp to search for specific collateral vessels that were missed on initial dissection and can be tied off (e.g., inferior epigastric or circumflex iliac vessels). The systemic mean arterial pressure may be increased by infusion of pressor agents, and the

central venous pressure may be increased by infusion of intravenous fluid. An increase in reservoir volume over time indicates a “steal” of systemic blood into the circuit, resulting in unintended lowering of the drug concentration, as well as discarding more drugcontaminated blood at the end of the procedure than intended. The first step is to increase the flow rate, though carefully monitoring outflow pressure to avoid intimal injury. The operating table can be tilted into Tredelenberg position to raise the venous pressure in the lower limb relative to the IVC. The central venous pressure and the systemic mean arterial pressure may be lowered by infusion of nitroglycerin. A complex situation may arise whereby the precordial scintillation monitor suggests ongoing leak, yet the reservoir volume is stable or increasing. This set of observations indicates bi-directional leak, with blood movement into the limb via one set of collateral vessels (i.e., venous) and out of the limb via different collateral vessels (i.e., arterial). The approach to this condition involves some trial and error (Table 1).

Agents Melphalan is the most widely used agent for HILP for melanoma. It is fortuitous that this was the agent selected for the first patient treated with HILP given the clinical complete and durable response it produced. Melphalan is a phenyl alanine mustard and taken up by melanoma cells (Luck 1956). Phenyl alanine itself is a precursor for melanine biosynthesis, and therefore taken up avidly by melanocytes. The mechanism of action of melphalan is through its ability to interact directly with DNA and cause miscoding. A second mechanism by which alkylating agents cause DNA damage is by formation of cross-bridges in the DNA, thereby preventing strand replication or transcription. Pharmacokinetic studies of melphalan following injection demonstrate that concentrations decline rapidly in a biexponential manner with distribution phase half-life of 10 min, and terminal elimination phase half-life of approximately

716

75 min. Pharmacokinetic studies in HILP demonstrate rapid uptake in tissue in the first 5–10 min, and continual reduction in drug concentration over 60 min to 10–20% of the starting concentration (Scott et al. 1992). Ideal dosing is calculated from limb volume rather than body weight. Limb volume expressed as a percentage of total body weight produced as much as a twofold variation in the population for both lower and upper extremities. This could lead to double the amount of melphalan administered to the same volume of tissue in two different individuals when dosed by weight. When dosed by limb volume, optimal dosages of 10 mg/L limb volume in the leg and 13 mg/L limb volume in the arm have been determined as the highest dose with acceptable risk, and little variation in toxicity (Kroon 1988; Benckhuijsen et al. 1988; Wieberdink et al. 1982). Melphalan is stable in sterile 0.9% sodium chloride for only 90 min at room temperature (Desmaris et al. 2015). Therefore, for an HILP procedure it should be prepared immediately before administration. Melphalan is eliminated from plasma primarily by chemical hydrolysis to inactive monohydroxymelphalan and dihydroxymelphalan. Renal excretion is extremely low. Identification of fluorescein in the urine from a leak test does not equate to a similar amount of melphalan in the urine. Nonetheless, all discarded bodily fluids from an HILP case should be handled as chemotherapy biohazard waste. Side effects of melphalan administration as part of HILP are discussed below. Tumor necrosis factor alpha TNFα is a proinflammatory cytokine produced by multiple different immune cells, and causes rapid and significant hemorrhagic necrosis of tumors. For these reasons there has been great interest in its potential as an anti-cancer agent. However, humans are exquisitely sensitive to toxic effects of TNFα including a septic-like response with fevers, tachycardia, cardiovascular collapse, pulmonary edema and shock. The maximum tolerated systemic dose has no effect on tumors. TNFα therefore is a logical choice of agent for isolated regional perfusion with a goal of achieving antitumor effects in the extremity without systemic side effects. TNFα alone has been used for

M. C. Perez et al.

isolated limb perfusion, with limited benefit observed (Posner et al. 1995). Of six treated patients, partial response of less than 1 month’s duration was seen in two patients and one patient had a complete response of only 7 months’ duration and then progressed. The observation that TNFα increases tumor neovascular permeability suggests that its best use is in combination with other agents. It has been combined most commonly with melphalan and interferon-γ. Other agents used in the past for isolated limb perfusion either alone or in combination with other agents include cisplatin, dacarbazine, actinomycin D, and fotemusine (Sanki et al. 2007).

Hyperthermia In Creech’s original report, isolated limb perfusion with chemotherapy was used without hyperthermia (Creech et al. 1958). Investigators subsequently observed that combined regional chemotherapy with mild hyperthermia produced higher response rates (Stehlin et al. 1975). There are no prospective randomized clinical trial results comparing isolated limb perfusion with versus without hyperthermia to inform this strategic decision. Hyperthermia during HILP affects tumor cells, other cell populations within the tumors including neovasculature and stromal cells, and normal tissues in the extremity. The addition of hyperthermia clearly increases side effects (e.g., effects on normal tissues). In one study, factors associated with a greater toxicity were tissue temperatures 40  C or higher, female gender, low pH in the circuit, and perfusion at a proximal level of isolation (Klaase et al. 1994c). However, it is equally clear that tumor cells are more susceptible to adverse effects of hyperthermia compared to normal cells. Results of animal model studies of isolated limb perfusion with versus without hyperthermia suggest added cytotoxicity and increased efficacy with the addition of the hyperthermia (Abdel-Wahab et al. 2004). These studies implicated a mechanism of enhanced cytotoxicity of l-phenylalanine mustard with hyperthermia rather than improved drug delivery and uptake.

Local and Recurrent Regional Metastases of Melanoma

Results HILP has been enthusiastically embraced over its greater than 60-year history, primarily because of a combination of the unique biology of melanoma in-transit metastases and the extraordinarily high response rate observed with this regional therapy. The primary agent used by nearly all centers has been melphalan, but most centers have developed protocols that differ from one another in drug administration schedule, temperature and duration of perfusion. Accordingly, it is difficult to reach conclusions about which techniques and schedules are optimal. In general, the complete response rate for HILP with melphalan alone is in the range from 40% to 60% (Table 2). The overall response rate (e.g., including partial responses) generally ranges from 60% to 90%. For leg perfusions, the melphalan dose varies from 0.8 to 2 mg/kg of body weight, or when dosed per liter of extremity volume from 6 to 10 mg/L. The dose for arm perfusions is generally less and ranges from 0.45 to 0.8 mg/kg. Perfusion times vary from 50 to 120 min. Target limb temperatures vary range from 37  (normothermia) to 42  . From this heterogeneous collection of reports it is not possible to draw a conclusion about the relationship between dose schedule and response rates. An interesting approach of sequential perfusions was evaluated and involved external iliac and common femoral approaches staged 6 weeks apart (Kroon et al. 1993). While the complete response rate with this approach jumped up to

Table 2 Wieberdink acute limb toxicity scale Grade I II III

IV

V

Clinical characteristics No subjective or objective evidence of reaction Slight erythema or edema Considerable erythema or edema with some blistering; slightly disturbed motility permissible Extensive epidermolysis or obvious damage to the deep tissues causing definite functional disturbances; threatened or manifest compartmental syndromes Reaction that may necessitate amputation

717

77%, no benefit in overall survival was observed relative to patients undergoing a single perfusion. The heterogeneity in procedures used also makes evaluation of the contribution of hyperthermia challenging. One retrospective analysis compared 218 patients treated with mild hyperthermia (39–40  C) to 116 patients perfused under normothermic conditions (37–38  C), in which no benefit in recurrence-free or overall survival was observed (Klaase et al. 1995). However, interpretation of these data are complicated by the observation that treatment schedules varied in many ways beyond temperature, including differences in number of perfusions. Most of the patients receiving normothermic perfusion received a double perfusion, and double perfusions were associated with a higher response rate than single perfusions (Klaase et al. 1994a). Other factors associated with a higher response rate in this study were negative regional lymph nodes and leg as the site of disease rather than the arm or foot. A separate study of 216 patients treated between 1978 and 1990 reported that prognostic factors for survival in order of significance were stage of disease, gender, age, Breslow thickness, Clark level of infiltration of the primary melanoma and the number of metastases (Klaase et al. 1994b). In a similar analysis from Tulane University on 174 patients treated with limb perfusion between 1957 and 1982 – some in the adjuvant setting – the factors associated with decreased survival rates in patients that also underwent elective lymph node dissection were increasing age, presence of subcutaneous or both subcutaneous and dermal metastases, treatment at normothermic temperatures or earlier date of treatment (Sutherland et al. 1987). The addition of tumor necrosis factor alpha (TNFα) and interferon gamma (IFNγ) to melphalan appears to be associated with an increased rate of response. The combination of preoperative subcutaneous interferon combined with a perfusate containing IFNγ 0.2 mg and TNFα 4 mg and melphalan 10 mg/L limb volume for lower extremities, or INFγ 0.2 mg and TNF 3 mg and melphalan 13 mg/L limb volume for upper extremities. The total perfusion treatment time was 90 min, with the melphalan added 30 min

718

into the perfusion. In a phase II study, 90% of melanoma patients treated experienced a complete response, with time to best response achieved in one third of the time compared to that typically observed with melphalan alone (Lienard et al. 1992). The tumors liquefied quickly, as has been observed with TNFα in animal models. Toxicity including shock and ARDS was observed despite use of prophylactic dopamine infusion. A successor phase III trial designed to evaluate the contribution of IFNγ did not reproduce the extremely high response rates even in the IFNγ-TNFα-melphalan arm (Lienard et al. 1999). There was a trend towards lower response rate in absence of IFNγ; however, this did not reach statistical significance. But the addition of TNFα to melphalan appeared to provide superior response rates compared to melphalan alone as observed in historical controls. A phase III randomized control trial performed at the National Cancer Institute (NCI) comparing the triple drug combination as championed by Lienard (Lienard et al. 1992) to melphalan alone and an interim analysis revealed a complete response rate of 80% in the triple-drug regimen compared to 61% for the melphalan-alone arm. The difference was statistically significant, even though the response-rate observed with melphalan alone was higher than typically observed. At the same time as this trial, Fraker and colleagues at the NCI conducted a trial in which the TNFα dose was escalated in combination with the standard melphalan and IFNγ doses (Fraker et al. 1996). The complete response rate in the 26 patients that received 4-mg TNFα was 76%, with an overall objective response rate of 92%. The complete response rate in the 12 patients that received 6 mg TNFα was 36% with an overall objective response rate of 100%. In the TNFα 6 mg group, regional toxicity was dose-limiting and greater in the group that received TNFα 4 mg, particularly skin blistering, painful myopathy and neuropathy. Based on these data the investigators concluded that HILP with TNFα at 4 mg combined with IFN and melphalan was considerably less toxic than TNFα at 6 mg, yet can lead to complete local responses in the majority of patients.

M. C. Perez et al.

Subsequent reports of HILP with TNFα in a three-drug regimen produced a range of observed complete response and survival rates. The American College of Surgeons Oncology Group conducted an important clinical trial evaluating the effects of TNFα in a two-drug regimen. Patients with in-transit metastases were randomized to melanoma combined with TNFα or melphalan alone (Cornett et al. 2006b). HILP was completed in 124 patients of the 133 enrolled. Greater toxicity was observed in the TNFα patients. Grade 4 adverse events were observed in 3 of 64 (4%) patients in the melphalan-alone arm compared to 11 of 65 (16%) patients in the melphalan-plus-TNF-alpha arm ( p = 0.04). The complete response rate at 3 months was 25% in the melphalan-alone arm and 26% in the melphalan-TNFα arm. The complete response rate at 6 months was higher in patients treated with the TNFα-containing regimen (42%) compared to the melphalan-alone regimen (20%), although this difference did not reach statistical significance. These clinical trial results do not support addition of TNFα to melphalan for treatment of in-transit metastases.

Specific Toxicities and Management Regional toxicities from HILP are caused by sensitivity of normal tissue to the high concentrations of toxic agents, hyperthermia, and mild acidemia. These may be in the form of lymphedema, skin blistering, painful neuralgia, or painful myopathy. The latter two conditions are managed conservatively with gabapentin and analgesics. Leg edema is managed with elevation and compression wraps. Skin blistering is self-limiting, and managed conservatively. Muscle injury and swelling is a grave sign because it can lead to compartment syndrome (see below). Postoperative hypotension resulting from “cytokine storm” may be observed even in the absence of TNFα in the perfusate and requires pressor agents for management. Melphalan left in the tissues of the extremity at the completion of perfusion and wash enters the systemic

Local and Recurrent Regional Metastases of Melanoma

circulation once limb vascularization is restored. This may cause acute postoperative nausea and emesis, which can be effectively managed with odansetron. Systemic melphalan may also lead to marrow suppression, manifest by neutropenia or pancytopenia 7–14 days after HILP. Wounds and abrasions on an extremity treated with HILP do not heal well for the first 3 months. It is therefore important for the patient to assiduously avoid cuts or skin abrasions in the first 3 months following HILP. And HILP procedures combined with superficial inguinal lymphadenectomy are at very high risk for wound breakdown. And when wounds do develop on the treated extremity, surgical debridement should be very conservative. Debridement down to healthy tissue is not typically rewarded with subsequent granulation tissue, and rather, most commonly results in simply a larger wound. Surgical debridement should be limited to unroofing areas of purulence. Toxicity can also result from acute vascular compromise. Any period of unrecognized postoperative ischemia that results from vascular inflow compromise potentiates the toxicity of the HILP treatment. Diligence in monitoring distal extremity pulses and perfusion is of paramount importance in the immediate postoperative period for early detection of vascular compromise. An atherosclerotic plaque that is cracked during the operation or creation of an intimal flap may result in vascular compromise post-operatively. Unilateral loss of pulses, cool extremity, or evidence of reduced perfusion should be investigated immediately with noninvasive studies (PVR, Doppler) and angiography or CT angiography. Immediate repair of compromised inflow is indicated. And following restoration of blood flow, careful monitoring for compartment syndrome should be performed by pressure measurements. A two-incision, fourcompartment fasciotomy is performed if indicated. Evidence for rhabdomyolysis should be sought by monitoring muscle tenderness, serum CK, and urine myoglobin. If found, maneuvers commonly employed include administration of large volumes of intravenous fluids, sodium bicarbonate, and potentially mannitol.

719

Isolated Limb Infusion Background Although effective, HILP is an invasive and costly procedure. Additionally, it is difficult to repeat the procedure secondary to the development of scar tissue from the initial procedure and as a result the overall complication rate increases from 28% to 51% for initial and repeat procedures, respectively (Cornett et al. 2006a). In the 1990s, at the now Melanoma Institute of Australia (MIA), Thompson et al. introduced the minimally-invasive and repeatable alternative to HILP known as isolated limb infusion (ILI) (Thompson et al. 1998). See also chapter ▶ “Isolated Limb Infusion for Melanoma.” ILI uses the same principal of high-dose chemotherapy infusion into an isolated limb, however vascular access is gained by percutaneously placed arterial and venous catheters and cytotoxic drugs are instilled at a low-flow under hypoxic conditions (Thompson et al. 1998). Presently, ILI is used throughout the world and has shown favorable response rates for locallyadvanced and in-transit melanoma (Kroon et al. 2014, 2016; Muilenburg et al. 2015; Li et al. 2018).

Patient Selection and Indications Indications for ILI are primarily patients with unresectable locally-advanced or in-transit melanoma of the upper or lower extremity and no evidence of distant metastases. Patients need to be cleared for general anesthesia prior to the procedure, but it is usually well-tolerated in most patients. ILI may be repeated after partial responses, or recurrences and progression following an initial response (Chai et al. 2012). ILI may also be performed in patients who also have distant metastatic disease in a palliative effort to control symptomatic locoregional disease. Kroon et al. reported a limb salvage rate of 86% in a series of 37 patients with symptomatic limb disease and documented distant metastases at the time of ILI (Kroon et al. 2009a). While

720

predominately performed for cutaneous melanoma, ILI has also been used to treat locallyadvanced soft tissue sarcoma and non-melanoma cutaneous malignancies including Merkel cell carcinoma, and squamous cell carcinoma (Mullinax et al. 2017; Turaga et al. 2011; O’Donoghue et al. 2017).

Technique Usually performed in the radiology department on the morning of the procedure, arterial and venous catheters are percutaneously placed under fluoroscopic guidance and advanced into the affected limb with the catheter tips positioned distal to the level of the tourniquet to ensure adequate isolation of the extremity. Limb temperatures are maintained greater than 37  C and usually closer to 40  C by a combination of liquid warming blankets on the affected extremity and overhead heaters which are institution-specific. The infusion portion of the procedure is then performed in the operating room under general anesthesia. Prior to inflating the tourniquet, patients are fully heparinized to achieve an activated clotting time (ACT) greater than 350 or 400 sec, depending on the institution. The tourniquet is inflated once adequate circulation is achieved by manually drawing blood from the venous catheter and reinjecting it into the arterial catheter with a 20 cc syringe. After tourniquet inflation, papaverine is routinely administered through the arterial catheter to maximize vasodilation. Isolation of the limb is checked by confirming cessation of flow in pedal or radial arteries by a Doppler probe. Limb volume measurements are used to determine the dose of chemotherapeutic agents, usually melphalan with or without actinomycin-D, and melphalan dosing is corrected for ideal body weight. Infusion lasts for typically 30 minutes and the chemotherapeutic agent is manually circulated with a syringe connected in line to the closed circuit with a heating source. Following infusion of chemotherapy, the limb is washed out with Hartmann’s or saline solution until the effluent is clear, and protamine is administered to reverse heparinization after the tourniquet is released.

M. C. Perez et al.

Patients are usually admitted to a monitored care unit postoperatively and remain on bed rest for the first 24 h following the procedure. Extremity neurovascular checks are performed at least twice daily along with serum creatinine phosphokinase (CPK) levels. Normal saline and corticosteroids should be administered if the patient’s CPK level exceeds 1000 IU/L. Patients are usually discharged from the hospital after CPK levels start to return towards baseline as long as other standard criteria for discharge have been met.

Response to Therapy Multiple single-center and multi-institution studies throughout the world have been published in the last two decades on the efficacy of ILI. Kroon et al. published the largest single institution experience consisting of 185 melanoma patients undergoing ILI at the MIA and reported an ORR of 84% (Kroon et al. 2008). Duke University (DU) and Moffitt Cancer Center (MCC) subsequently reported their single-institution experiences of 61 procedures and 79 procedures, respectively. The ORR at DU was 44% with 30% of procedures achieving a CR, while MCC reported an ORR of 70% and a CR in 32% of procedures (Beasley et al. 2008; Wong et al. 2013). O’Donoghue et al. later extended the MCC experience to 145 ILI procedures for melanoma and reported an ORR of 59% with a CR in 26% of procedures (O’Donoghue et al. 2017). In an Australian study of five institutions, a 75% ORR (33% CR) was reported in 316 ILI procedures. (Kroon et al. 2016), while two multicenter studies in the United States reported an ORR of 64% (31% CR) and 57% (34% CR) in 128 patients and 160 patients, respectively (Beasley et al. 2009; Muilenburg et al. 2015). Kroon et al. also performed a systematic review of 576 patients around the world who underwent ILI and reported a 73% ORR with 33% of patients achieving a CR (Kroon et al. 2014). Differences in response rates among varying institutions are likely related to multiple factors including heterogeneous patient populations, differences in technique, and inconsistencies in

Local and Recurrent Regional Metastases of Melanoma

721

response criteria used. The MIA series included patients with stage I-II disease, while the DU and MCC reports were comprised of stage III patients only. Drug-exposure times also differ between institutions and only certain institutions routinely use papaverine for vasodilation. For response criteria, DU and MCC determined response to ILI at 3 months using the Response Evaluation Criteria in Solid Tumors (RECIST) criteria version 1.1, while the MIA determined best response using standard World Health Organization (WHO) criteria (Eisenhauer et al. 2009; Wold Health Organization 1979). However, the Australian multicenter analysis showed response rates to be significantly different on multivariate analysis among the included institutions which used the same response criteria (Kroon et al. 2016).

complete response (CR) rates (Steinman et al. 2014; Muilenburg et al. 2015). Muilenburg et al. reported an ORR and CR rate of, respectively, 73% and 50% in 60 patients with low BOD (less than 10 lesions, none greater than 2 cm) and an ORR and CR rate, respectively, of 47% and 24% in those with a high BOD ( p = 0.002) (Muilenburg et al. 2015). At the MIA, patients who underwent ILI for one lesion were found to have an improved OS when compared to those who had multiple lesions (Kroon et al. 2008). In the MIA single-center series, maximal depth of the primary tumor (Breslow depth), a well-known and strong indicator of overall survival in all patients with localized melanoma, was not shown to be a significant predictor of response to ILI (Kroon et al. 2008). However, in the multicenter analysis, a thinner primary melanoma was significantly associated with a higher response rate on multivariate analysis ( p = 0.04) (Kroon et al. 2016).

Survival After ILI Response to therapy has shown a significant association with improved survival in patients who undergo ILI. O’Donoghue et al. reported both infield progression-free survival (IPFS) and overall survival (OS) to be significantly higher in melanoma patients who responded to therapy when compared to those who did not respond to ILI (IPFS 14.1 vs. 3.2 months, p < 0.0001; OS 56.0 vs. 26.7 months, p = 0.0004) (O’Donoghue et al. 2017). Kroon et al. reported a median survival of 38 months for all patients who underwent ILI and showed a significantly higher OS in those patients who achieved a CR (53 months) compared to those who did not (25 months; p = 0.005) (Kroon et al. 2008). Furthermore, Wong et al. showed that resection of residual disease following ILI improves both disease-free and overall survival with rates similar to those who experienced a CR after ILI (Wong et al. 2014).

Burden of Disease Multiple studies have shown burden of disease (BOD) to be associated with response to therapy. Patients with a lower BOD have been shown to have higher overall response rates (ORR) and

Toxicity ILI is generally a well-tolerated procedure with the majority of the toxicity limited to transient edema and hyperpigmentation of the limb postoperatively. The Wieberdink toxicity scale is routinely used to characterize toxicity related to ILI (Table 2) (Wieberdink et al. 1982). At the MIA, 56% of patients experienced grade II toxicity, 39% of patients experienced grade III toxicity, and 3% of patients experienced grade IV toxicity (Kroon et al. 2009b). The Australian multicenter study of 316 procedures reported a rate of 27% grade III toxicity, and 3% grade IV toxicity (Kroon et al. 2016). O’Donoghue et al. reported a grade III toxicity rate of 11.4% in a series of 201 ILI procedures at MCC with 145 of the procedures being performed for melanoma (O’Donoghue et al. 2017). Only one patient in this series developed a grade IV toxicity which resulted in a fasciotomy to treat compartment syndrome (O’Donoghue et al. 2017). DU used the National Cancer Institute Common Technology Criteria for Adverse Events (CTCAE) version 3 and has shown similar toxicity to MCC and the MIA, with grade III toxicities reported in 18%

722

of patients, excluding CPK elevation (Beasley et al. 2008). Although toxicity score was significantly associated with an improved ORR on both univariate and multivariate analysis in the single-center MIA study, increased toxicity has consistently been shown to not improve rates of CR, in-field progression, or survival (Kroon et al. 2008, 2009b). Independent risk factors which have been shown to correlate with limb toxicity include postoperative CPK level, high peak and final melphalan concentration, and tourniquet time. Systemic toxicities related to ILI are rare and mostly comprised of mild nausea and vomiting likely related to the general anesthesia used for the procedure. This is likely a result of the low rate of chemotherapy which leaks into the systemic circulation from the isolated limb. Systemic melphalan was only detected in 11 (6%) patients in the MIA single-center series, and 10 of these patients’ systemic leakage was 6 months) response rate 16.3% of patients

56.8% of lesions

26% of patients; 53% of lesions 25% of patients; 44.4% of lesions 5% of patients; 28.3% of lesions

10.8% of patients

93 days

4 months

Not estimable

54% (7/13) of noninjected lesions

22% of uninjected nonvisceral lesions complete response; 9% uninjected visceral lesions complete response 33% response in uninjected lesions

Local and Recurrent Regional Metastases of Melanoma 727

728

there are ongoing studies combining intralesional therapy with systemic immunotherapy agents.

Combination with Systemic Immune Therapies Recently, the combination of intralesional therapies with systemic immune therapies has shown promising results. By combining different treatment modalities, particularly strategies with immunologic mechanisms, there appears to be a synergistic effect. A phase I trial of T-VEC combined with pembrolizumab in stage IIIB-IV melanoma demonstrated an objective response rate of 48% and complete response rate of 14% (Long et al. 2016). Grade 3 or 4 adverse events were reported in 33% of patients. A phase I trial of T-VEC combined with Ipilimumab established safety of the regimen and demonstrated an objective response rate of 50%. A 6 months durable response was seen in 44% of patients that responded (Puzanov et al. 2016). A subsequent phase II randomized open-label study of T-VEC in combination with Ipilimumab versus Ipilimumab alone demonstrated an overall objective response rate of 39% in the combined arm versus 18% in the Ipilimumab alone arm (Chesney et al. 2018). Visceral lesion response rates were 53% in the combined arm versus 23% of the Ipilimumab alone arm. Incidence of 3 grade adverse events was 45% and 35%, respectively.

Electrochemotherapy (ECT) Electrochemotherapy is the administration of a chemotherapeutic or cytotoxic drug that is normally not permeable to cell membranes, followed by local application of electrical currents which transiently permeabilize cell membranes and allow the cytotoxic drug to enter tumor cells. This therapeutic modality had initially shown promise in animal models of multiple tumor types (Mir et al. 1998). In clinical studies in patients with melanoma the most commonly used cytotoxic drug has been

M. C. Perez et al.

bleomycin and cisplatin administered either intravenously or by intratumoral injection (Mali et al. 2013). In some studies, patients required either local or general anesthesia for the treatments but the treatment was well-tolerated with reported adverse effects including localized erythema, edema, and pain (Mir et al. 1998). In a systematic review and pooled analysis of 22 studies of electroporation used in patients with melanoma, 150 patients and 922 lesions were included. The objective response rate per lesion was 80.6% and the complete response rate was 56.8% (Mali et al. 2013). A phase II study including 19 patients randomized their lesions to intralesional bleomycin injection and electroporation versus intralesional bleomycin alone. In this study, the objective response rate was 78% for patients treated with electroporation compared to 32% with bleomycin alone (x2 = 9.39, 1df, p = 0.002) (Byrne et al. 2005). While these results are promising, larger randomized studies are needed to assess the efficacy of electrochemotherapy.

Conclusion In summary, intralesional therapies have been investigated in melanoma since the beginning of the twentieth century. Over the years, some therapies such as BCG, GM-CSF, and Allovectin had shown promise in earlier reports but demonstrated not to be effective in larger randomized studies. Meanwhile other agents such as T-VEC have demonstrated efficacy in large randomized studies and gained FDA approval and wide-spread use. PV-10, Daromun, and Coxsackievirus A21, and electrochemotherapy are still under investigation but have shown promising results thus far. Finally, perhaps the most promising results of all have been with the combination of intralesional therapies and systemic immunotherapies in which a synergistic effect appears to be observed and could potentially be used both in the adjuvant and neoadjuvant setting. The recent flurry of research on intralesional therapies for melanoma have made this an exciting treatment modality for patients with advanced melanoma.

Local and Recurrent Regional Metastases of Melanoma

Neoadjuvant Therapy for Borderline Resectable Nodal Metastasis Treatment of advanced melanoma now includes several options secondary to advances in systemic treatments with immunotherapy, or checkpoint blockade, and targeted therapy, or BRAF inhibition. While surgical resection remains a basic tenet of treatment, today there are multiple possibilities for the timing of surgery and use of systemic agents. This chapter will look at the role of neoadjuvant treatment in patients with melanoma. See also chapter ▶ “Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients.” The advantage of neoadjuvant treatment in patients with borderline resectable disease is several. First of all, provides an opportunity to decrease the size of the tumor and make a surgical resection more feasible. The pathology of the tumor also provides a biological window into the mechanism of response and/or resistance to treatment which can also facilitate the development of biomarkers. Neoadjuvant treatment offers the opportunity to select out the patients most likely to respond to treatment, or better tumor biology. However, an alternative argument is that the patients who fail neoadjuvant therapy have lost the opportunity regional control through surgical resection. Therefore, it is important to understand the response rate and time to response prior to embarking on neoadjuvant treatment. Several issues in neoadjuvant treatment however, remain unclear. For example, the optimal time of neoadjuvant treatment prior to resection, whether the measurement of response based upon radiologic or pathologic response, and whether the response rate in the neoadjuvant setting supercedes therapy as an adjuvant. Prior to the approval of checkpoint blockade, effective treatments in melanoma consisted of chemotherapy, temodar given as a single agent, or CVD (cisplatinum, vincristine, and dacarbazine), and immunotherapy consisting of IL-2 and interferon, which were also studied as neoadjuvant treatments in patients with Stage III melanoma. Amongst patients with measurable disease, an early trial of CVD given in 2–3 cycles followed by surgery and subsequent CVD if a response demonstrated a 48% response rate (Buzaid et al.

729

1994). This was followed by multiple trials which demonstrated efficacy of biochemotherapy, which includes CVD in combination with IL2 and interferon. Selected series demonstrated response rates of 39–50% (Buzaid et al. 1998; Gibbs et al. 2002). However, a phase II multicenter trial demonstrated a slightly lower response rate of 26% (22% partial and 4% complete) (Lewis et al. 2006). Overall, these response rates were not higher than the 48% response rate demonstrated in a Phase III trial of patients with metastatic melanoma, which unfortunately did not improve survival (Eton et al. 2002). Single agent temodar was also studied in a neoadjuvant trial of resectable Stage III, IVa disease. The overall response rate of 16% with two patients that had a complete response (CR), but this was not different than responses noted in Stage IV disease (Shah et al. 2010). Interferon-α2b (IFN) was also studied in a neoadjuvant/adjuvant fashion in patients with Stage III disease and palpable nodes. Patients underwent 4 weeks of treatment prior to surgery, followed by maintenance for a total of 1 year of treatment, with an impressive clinical response rate 55%, and three patients with a pathological CR (Moschos et al. 2006). The studies above validated the feasibility of neoadjuvant trials, but unfortunately did not improve survival. However, subsequently the landscape of the treatment of melanoma changed with effective systemic therapies. The first drug to improve overall survival in patients with metastatic melanoma was anti-CTLA-4, or Ipilimumab. The response rate was 10% with a similar improvement in survival over standard therapy alone, vaccination with gp100 or dacarbazine (Hodi et al. 2010; Robert et al. 2011). Subsequently, anti-PD1 therapy has demonstrated a response rate of 30% and the combination of CTLA-4 and PD-1 blockade increases responses to 58% (Topalian et al. 2012; Wolchok et al. 2017). With active drug combinations, there was rationale to use these drugs in the neoadjuvant setting. Tarhini et al. looked at the role of neoadjuvant Ipilimumab in patients with Stage IIIB/C melanoma. Patients had a radiologic assessment, pretreatment biopsy, and then received Ipilimumab for two doses, followed by repeat radiologic assessment and surgery. Of the 33 evaluable

730

patients, all had viable melanoma on the pathologic analysis. Investigation of the tumor demonstrated immune activation with CD4 and CD8 cells after treatment, and a low baseline CD20 cell number trended with a poor clinical response. On imaging, 9% had a partial response, 64% had stable disease, and 24% patients has progressive disease (Tarhini et al. 2014). Importantly, no patients lost the opportunity to have surgery, but the response rate was similar to that seen in Stage IV disease, and therefore subsequent trials have also employed anti-PD1 therapy. Early report of the Optimal Neoadjuvant Combination (OpACIN) trial of combination of Ipilimumab and nivolimab(IPI NIVO) demonstrated responses in 8 of 10 patients in the phase Ib portion of the trial. This trial is now being expanded to additional patients, but also noted many patients did not complete treatment because of toxicity (Rozeman et al. 2017). There are currently open trials of neoadjuvant therapy and the early results were pooled and presented in abstract form of the International Neoadjuvant Melanoma Consortium (INMC). Of the combined 21 patients receiving immunotherapy with IPI NIVO or NIVO alone, a pathologic complete response was found in 8 patients (38%), none of which have had a recurrence (Menzies et al. 2017). Further analysis of genomics of responding and non-responding patients is ongoing. Targeted inhibition of BRAF mutated tumors has been the other area of major advance in melanoma. While single agent BRAF inhibition improved survival, resistance developed in many within a year (Chapman et al. 2011). Combination BRAF/MEK has been shown to improve survival in Stage IV patients beyond single agent and is now the preferred drug combination (Flaherty et al. 2012b). Studies utilizing a neoadjuvant approach with targeted therapy have also shown promise. A prospective trial randomized patients in a 2:1 fashion to neoadjuvant BRAF/MEK inhibition for 12 weeks followed by surgery and up to 44 weeks of postop treatment, for a total of 52 weeks. After 18 months of follow-up the trial was stopped because 10 of 14 patients in the treatment arm (71%) remained free of disease while none (0 of 7 patients) in the standard of

M. C. Perez et al.

care arm remained free of disease. The radiologic response rate in this cohort was 85%, and of the 12 patients that went on to have surgery, the pathological CR rate was 58% (Amaria et al. 2018). The toxicity was acceptable and similar to that seen in patients with Stage IV disease, the majority being fevers, chills and headache. Interestingly, a pretreatment biopsy demonstrating less pERK expression, and increased CD8 toxicity was associated with a CR. The trial was not powered to detect survival differences, and has remained open as a phase II single arm trial. However, the results are consistent with a larger experience of the international neoadjuvant melanoma consortium, which included patients from this trial, with a pathologic complete response rate of 55% (Menzies et al. 2017). All of these neoadjuvant targeted therapy trials administer targeted BRAF/ MEK preoperatively and postoperatively for a year, and therefore do not answer the question of whether neoadjuvant therapy alone would be sufficient. However, the progression free survival (PFS) of targeted therapy in the adjuvant setting is 58% for combination therapy and 39% for placebo, so the benefit of the combination neoadjuvant plus adjuvant approach appear to be improved in these early studies. While the results of neoadjuvant studies in the era of effective systemic therapies are still being finalized, the data to date demonstrates several important points. Responses to treatment occur within 6–12 weeks and no trial to date has reported the loss of control of a regional nodal basin by administering neoadjuvant therapy. Therefore, neoadjuvant therapy remains a safe option for patients with borderline resectable disease. The response rate to 12 weeks of targeted therapy with BRAF and MEK inhibition, 85%, with a CR rate of over 50% at surgery, is the highest noted to date, and better tolerated. The early reports of combination Ipilimumab and nivolimab are also encouraging, with responses from 80% after 6 weeks of treatment, and a CR rate at surgery of 38%. However, combination immunotherapy data is not as mature, and treatment is associated with more toxicity and this risk must be weighed on an individual basis. It will be interesting to see if the high response rates seen in

Local and Recurrent Regional Metastases of Melanoma

731

early combination trials persist in larger cohorts. Also, future studies will look at whether surgery can be delayed in responding patients to the point of maximal response. In general, for a patient with borderline resectable disease and a BRAF mutation, neoadjuvant therapy has the highest tolerability and chance of success. Hopefully the results of these trials will elucidate the patients who are most likely to benefit from the neoadjuvant approach in the future.

Hawkins M, Paradise C, Kunkel L, Rosenberg SA (1999) High-dose recombinant interleukin 2 therapy for patients with metastatic melanoma: analysis of 270 patients treated between 1985 and 1993. J Clin Oncol 17:2105–2116 Balch CM, Urist MM, Karakousis CP, Smith TJ, Temple WJ, Drzewiecki K, Jewell WR, Bartolucci AA, Mihm MC Jr, Barnhill R et al (1993) Efficacy of 2-cm surgical margins for intermediate-thickness melanomas (1 to 4 mm). Results of a multi-institutional randomized surgical trial. Ann Surg 218:262–267. discussion 267–9 Balch CM, Soong SJ, Smith T, Ross MI, Urist MM, Karakousis CP, Temple WJ, Mihm MC, Barnhill RL, Jewell WR, Wanebo HJ, Desmond R (2001) Long-term results of a prospective surgical trial comparing 2 cm vs. 4 cm excision margins for 740 patients with 1-4 mm melanomas. Ann Surg Oncol 8:101–108 Balch CM, Gershenwald JE, Soong SJ, Thompson JF, Atkins MB, Byrd DR, Buzaid AC, Cochran AJ, Coit DG, Ding S, Eggermont AM, Flaherty KT, Gimotty PA, Kirkwood JM, Mcmasters KM, Mihm MC Jr, Morton DL, Ross MI, Sober AJ, Sondak VK (2009) Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol 27:6199–6206 Bast RC Jr, Zbar B, Borsos T, Rapp HJ (1974) BCG and cancer. N Engl J Med 290:1458–1469 Beasley GM, Petersen RP, Yoo J, Mcmahon N, Aloia T, Petros W, Sanders G, Cheng TY, Pruitt SK, Seigler H, Tyler DS (2008) Isolated limb infusion for in-transit malignant melanoma of the extremity: a well-tolerated but less effective alternative to hyperthermic isolated limb perfusion. Ann Surg Oncol 15:2195–2205 Beasley GM, Caudle A, Petersen RP, Mcmahon NS, Padussis J, Mosca PJ, Zager JS, Hochwald SN, Grobmyer SR, Delman KA (2009) A multi-institutional experience of isolated limb infusion: defining response and toxicity in the US. J Am Coll Surg 208:706–715 Bedikian AY, Richards J, Kharkevitch D, Atkins MB, Whitman E, Gonzalez R (2010) A phase 2 study of high-dose Allovectin-7 in patients with advanced metastatic melanoma. Melanoma Res 20:218–226 Benckhuijsen C, Kroon BB, Van Geel AN, Wieberdink J (1988) Regional perfusion treatment with melphalan for melanoma in a limb: an evaluation of drug kinetics. Eur J Surg Oncol 14:157–163 Borgstein PJ, Meijer S, Van Diest PJ (1999) Are locoregional cutaneous metastases in melanoma predictable? Ann Surg Oncol 6:315–321 Boyd KU, Wehrli BM, Temple CL (2011) Intra-lesional interleukin-2 for the treatment of in-transit melanoma. J Surg Oncol 104:711–717 Brys AK, Bhatti L, Bashir MR, Jaffe TA, Beasley GM, Nath NS, Salama AK, Tyler DS, Mosca PJ (2016) Computed tomography-based limb volume measurements for isolated limb infusion in melanoma. Ann Surg Oncol 23:1090–1095 Buchbinder EI, Mcdermott DF (2014) Interferon, interleukin-2, and other cytokines. Hematol Oncol Clin North Am 28:571–583

References Abdel-Wahab OI, Grubbs E, Viglianti BL, Cheng TY, Ueno T, Ko S, Rabbani Z, Curtis S, Pruitt SK, Dewhirst MW, Tyler DS (2004) The role of hyperthermia in regional alkylating agent chemotherapy. Clin Cancer Res 10:5919–5929 Agarwala SS, Neuberg D, Park Y, Kirkwood JM (2004) Mature results of a phase III randomized trial of bacillus Calmette-Guerin (BCG) versus observation and BCG plus dacarbazine versus BCG in the adjuvant therapy of American Joint Committee on Cancer Stage I-III melanoma (E1673): a trial of the Eastern Oncology Group. Cancer 100:1692–1698 Amaria RN, Prieto PA, Tetzlaff MT, Reuben A, Andrews MC, Ross MI, Glitza IC, Cormier J, Hwu WJ, Tawbi HA, Patel SP, Lee JE, Gershenwald JE, Spencer CN, Gopalakrishnan V, Bassett R, Simpson L, Mouton R, Hudgens CW, Zhao L, Zhu H, Cooper ZA, Wani K, Lazar A, Hwu P, Diab A, Wong MK, Mcquade JL, Royal R, Lucci A, Burton EM, Reddy S, Sharma P, Allison J, Futreal PA, Woodman SE, Davies MA, Wargo JA (2018) Neoadjuvant plus adjuvant dabrafenib and trametinib versus standard of care in patients with high-risk, surgically resectable melanoma: a single-centre, open-label, randomised, phase 2 trial. Lancet Oncol 19:181–193 Andtbacka RHI, Kaufman HL, Collichio F, Amatruda T, Senzer N, Chesney J, Delman KA, Spitler LE, Puzanov I, Agarwala SS, Milhem M, Cranmer L, Curti B, Lewis K, Ross M, Guthrie T, Linette GP, Daniels GA, Harrington K, Middleton MR, Miller WH, Zager JS, Ye YN, Yao B, Li A, Doleman S, Vanderwalde A, Gansert J, Coffin RS (2015) Talimogene Laherparepvec Improves Durable Response Rate in Patients With Advanced Melanoma. J Clin Oncol 33:2780–2U98 Andtbacka RHI, Curti B, Hallmeyer S, Fox B, Feng Z, Paustian C, Bifulco C, Grose M, Shafren DR (2016) Phase II CALM extension study: intratumoral CAVATAK increases immune-cell infiltrates and upregulates immune-checkpoint molecules in the microenvironment of lesions from advanced melanoma patients. J Immunother Cancer 4:319 Atkins MB, Lotze MT, Dutcher JP, Fisher RI, Weiss G, Margolin K, Abrams J, Sznol M, Parkinson D,

732 Buzaid A, Legha SS, Balch CM, Ross M, Ring S, Plager C, Papadopoulos NE, El-Naggar AK, Benjamin RS (1994) Pilot study of preoperative chemotherapy with cisplatin, vinblastine, and dacarbazine in patients with local-regional recurrence of melanoma. Cancer 74:2476–2482 Buzaid AC, Colome M, Bedikian A, Eton O, Legha SS, Papadopoulos N, Plager C, Ross M, Lee JE, Mansfield P, Rice J, Ring S, Lee JJ, Strom E, Benjamin R (1998) Phase II study of neoadjuvant concurrent biochemotherapy in melanoma patients with localregional metastases. Melanoma Res 8:549–556 Byers BA, Temple-Oberle CF, Hurdle V, Mckinnon JG (2014) Treatment of in-transit melanoma with intralesional interleukin-2: a systematic review. J Surg Oncol 110:770–775 Byrne CM, Thompson JF, Johnston H, Hersey P, Quinn MJ, Michael Hughes T, Mccarthy WH (2005) Treatment of metastatic melanoma using electroporation therapy with bleomycin (electrochemotherapy). Melanoma Res 15:45–51 Calabro A, Singletary SE, Balch CM (1989) Patterns of relapse in 1001 consecutive patients with melanoma nodal metastases. Arch Surg 124:1051–1055 Calabro A, Singletary SE, Carrasco CH, Legha SS (1990) Intraarterial infusion chemotherapy in regionally advanced malignant melanoma. J Surg Oncol 43:239–244 Cascinelli N, Bufalino R, Marolda R, Belli F, Nava M, Galluzzo D, Santinami M, Levene A (1986) Regional non-nodal metastases of cutaneous melanoma. Eur J Surg Oncol 12:175–180 Chai CY, Deneve JL, Beasley GM, Marzban SS, Chen YA, Rawal B, Grobmyer SR, Hochwald SN, Tyler DS, Zager JS (2012) A multi-institutional experience of repeat regional chemotherapy for recurrent melanoma of extremities. Ann Surg Oncol 19:1637–1643 Chapman PB, Hauschild A, Robert C, Haanen JB, Ascierto P, Larkin J, Dummer R, Garbe C, Testori A, Maio M, Hogg D, Lorigan P, Lebbe C, Jouary T, Schadendorf D, Ribas A, O’day SJ, Sosman JA, Kirkwood JM, Eggermont AM, Dreno B, Nolop K, Li J, Nelson B, Hou J, Lee RJ, Flaherty KT, Mcarthur GA, Group B-S (2011) Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N Engl J Med 364:2507–2516 Chesney J, Puzanov I, Collichio F, Singh P, Milhem MM, Glaspy J, Hamid O, Ross M, Friedlander P, Garbe C, Logan TF, Hauschild A, Lebbe C, Chen L, Kim JJ, Gansert J, Andtbacka RHI, Kaufman HL (2018) Randomized, open-label phase II study evaluating the efficacy and safety of talimogene laherparepvec in combination with Ipilimumab versus Ipilimumab alone in patients with advanced, unresectable melanoma. J Clin Oncol 36:1658–1667 Clary BM, Brady MS, Lewis JJ, Coit DG (2001) Sentinel lymph node biopsy in the management of patients with primary cutaneous melanoma: review of a large singleinstitutional experience with an emphasis on recurrence. Ann Surg 233:250–258

M. C. Perez et al. Coley WB (1910) The treatment of inoperable sarcoma by bacterial toxins (the mixed toxins of the Streptococcus erysipelas and the Bacillus prodigiosus). Proc R Soc Med 3:1–48 Cornett WR, Mccall LM, Petersen RP, Ross MI, Briele HA, Noyes RD, Sussman JJ, Kraybill WG, Kane JM 3rd, Alexander HR, Lee JE, Mansfield PF, Pingpank JF, Winchester DJ, White RL Jr, Chadaram V, Herndon JE 2nd, Fraker DL, Tyler DS (2006a) Randomized multicenter trial of hyperthermic isolated limb perfusion with melphalan alone compared with melphalan plus tumor necrosis factor: American College of Surgeons Oncology Group Trial Z0020. J Clin Oncol 24:4196–4201 Cornett WR, Mccall LM, Petersen RP, Ross MI, Briele HA, Noyes RD, Sussman JJ, Kraybill WG, Kane JM 3rd, Alexander HR, Lee JE, Mansfield PF, Pingpank JF, Winchester DJ, White RL Jr, Chadaram V, Herndon JE 2nd, Fraker DL, Tyler DS, American College of Surgeons Oncology Group Trial, Z (2006b) Randomized multicenter trial of hyperthermic isolated limb perfusion with melphalan alone compared with melphalan plus tumor necrosis factor: American College of Surgeons Oncology Group Trial Z0020. J Clin Oncol 24:4196–4201 Creech O, Krementz ET, Ryan RF, Winblad JN (1958) Chemotherapy of cancer: regional perfusion utilizing an extracorporeal circuit. Ann Surg 148:616–632 Dalal KM, Patel A, Brady MS, Jaques DP, Coit DG (2007) Patterns of first-recurrence and post-recurrence survival in patients with primary cutaneous melanoma after sentinel lymph node biopsy. Ann Surg Oncol 14:1934–1942 Danielli R, Patuzzo R, Di Giacomo AM, Gallino G, Maurichi A, Di Florio A, Cutaia O, Lazzeri A, Fazio C, Miracco C, Giovannoni L, Elia G, Neri D, Maio M, Santinami M (2015) Intralesional administration of L19-IL2/L19-TNF in stage III or stage IVM1a melanoma patients: results of a phase II study. Cancer Immunol Immunother 64:999–1009 Desmaris RP, Mercier L, Paci A (2015) Stability of melphalan in 0.9% sodium chloride solutions prepared in polyvinyl chloride bags for intravenous injection. Drugs R&D 15:253–259 Di Filippo F, Calabro A, Giannarelli D, Carlini S, Cavaliere F, Moscarelli F, Cavaliere R (1989) Prognostic variables in recurrent limb melanoma treated with hyperthermic antiblastic perfusion. Cancer 63:2551–2561 Doepker MP, Thompson ZJ, Fisher KJ, Yamamoto M, Nethers KW, Harb JN, Applebaum MA, Gonzalez RJ, Sarnaik AA, Messina JL, Sondak VK, Zager JS (2016) Is a wider margin (2 cm vs. 1 cm) for a 1.01-2.0 mm melanoma necessary? Ann Surg Oncol 23:2336–2342 Doukas J, Rolland A (2012) Mechanisms of action underlying the immunotherapeutic activity of Allovectin in advanced melanoma. Cancer Gene Ther 19:811–817 Eisenhauer EA, Therasse P, Bogaerts J, Schwartz LH, Sargent D, Ford R, Dancey J, Arbuck S, Gwyther S, Mooney M, Rubinstein L, Shankar L, Dodd L, Kaplan R, Lacombe D, Verweij J (2009) New response

Local and Recurrent Regional Metastases of Melanoma

733

evaluation criteria in solid tumours: revised RECIST guideline (version 1.1). Eur J Cancer 45:228–247 Elias EG, Zapas JL, Mccarron EC, Beam SL, Hasskamp JH, Culpepper WJ (2008) Sequential administration of GM-CSF (Sargramostim) and IL-2 +/ autologous vaccine as adjuvant therapy in cutaneous melanoma: an interim report of a phase II clinical trial. Cancer Biother Radiopharm 23:285–291 Essner R, Conforti A, Kelley MC, Wanek L, Stern S, Glass E, Morton DL (1999) Efficacy of lymphatic mapping, sentinel lymphadenectomy, and selective complete lymph node dissection as a therapeutic procedure for early-stage melanoma. Ann Surg Oncol 6:442–449 Eton O, Legha SS, Bedikian AY, Lee JJ, Buzaid AC, Hodges C, Ring SE, Papadopoulos NE, Plager C, East MJ, Zhan F, Benjamin RS (2002) Sequential biochemotherapy versus chemotherapy for metastatic melanoma: results from a phase III randomized trial. J Clin Oncol 20:2045–2052 Faries MB (2016) Intralesional immunotherapy for metastatic melanoma: the oldest and newest treatment in oncology. Crit Rev Oncog 21:65–73 Flaherty KT, Infante JR, Daud A, Gonzalez R, Kefford RF, Sosman J, Hamid O, Schuchter L, Cebon J, Ibrahim N, Kudchadkar R, Burris HA 3rd, Falchook G, Algazi A, Lewis K, Long GV, Puzanov I, Lebowitz P, Singh A, Little S, Sun P, Allred A, Ouellet D, Kim KB, Patel K, Weber J (2012a) Combined BRAF and MEK inhibition in melanoma with BRAF V600 mutations. N Engl J Med 367:1694–1703 Flaherty KT, Robert C, Hersey P, Nathan P, Garbe C, Milhem M, Demidov LV, Hassel JC, Rutkowski P, Mohr P, Dummer R, Trefzer U, Larkin JM, Utikal J, Dreno B, Nyakas M, Middleton MR, Becker JC, Casey M, Sherman LJ, Wu FS, Ouellet D, Martin AM, Patel K, Schadendorf D, Group M S (2012b) Improved survival with MEK inhibition in BRAF-mutated melanoma. N Engl J Med 367:107–114 Fraker DL, Alexander HR, Andrich M, Rosenberg SA (1996) Treatment of patients with melanoma of the extremity using hyperthermic isolated limb perfusion with melphalan, tumor necrosis factor, and interferon gamma: results of a tumor necrosis factor dose-escalation study. J Clin Oncol 14:479–489 Gershenwald JE, Colome MI, Lee JE, Mansfield PF, Tseng C, Lee JJ, Balch CM, Ross MI (1998) Patterns of recurrence following a negative sentinel lymph node biopsy in 243 patients with stage I or II melanoma. J Clin Oncol 16:2253–2260 Gershenwald JE, Berman RS, Porter G, Mansfield PF, Lee JE, Ross MI (2000) Regional nodal basin control is not compromised by previous sentinel lymph node biopsy in patients with melanoma. Ann Surg Oncol 7:226–231 Ghussen F, Kruger I, Groth W, Stutzer H (1988) The role of regional hyperthermic cytostatic perfusion in the treatment of extremity melanoma. Cancer 61:654–659 Gibbs P, Anderson C, Pearlman N, Laclaire S, Becker M, Gatlin K, O’driscoll M, Stephens J, Gonzalez R (2002) A phase II study of neoadjuvant biochemotherapy for stage III melanoma. Cancer 94:470–476

Hafstrom L, Rudenstam CM, Blomquist E, Ingvar C, Jonsson PE, Lagerlof B, Lindholm C, Ringborg U, Westman G, Ostrup L (1991) Regional hyperthermic perfusion with melphalan after surgery for recurrent malignant melanoma of the extremities. Swedish Melanoma Study Group. J Clin Oncol 9:2091–2094 Hersh EM, Gutterman JU, Mavligit GM (1977) BCG as adjuvant immunotherapy for neoplasia. Annu Rev Med 28:489–515 Hodi FS, O’day SJ, Mcdermott DF, Weber RW, Sosman JA, Haanen JB, Gonzalez R, Robert C, Schadendorf D, Hassel JC, Akerley W, Van Den Eertwegh AJ, Lutzky J, Lorigan P, Vaubel JM, Linette GP, Hogg D, Ottensmeier CH, Lebbe C, Peschel C, Quirt I, Clark JI, Wolchok JD, Weber JS, Tian J, Yellin MJ, Nichol GM, Hoos A, Urba WJ (2010) Improved survival with Ipilimumab in patients with metastatic melanoma. N Engl J Med 363:711–723 Hu JCC, Coffin RS, Davis CJ, Graham NJ, Groves N, Guest PJ, Harrington KJ, James ND, Love CA, Mcneish I, Medley LC, Michael A, Nutting CM, Pandha HS, Shorrock CA, Simpson J, Steiner J, Steven NM, Wright D, Coombes RC (2006) A phase I study of OncoVEX (GM-CSF), a second-generation oncolytic herpes simplex virus expressing granulocyte macrophage colony-stimulating factor. Clin Cancer Res 12:6737–6747 Jonsson PE, Hafstrom L, Hugander A (1983) Results of regional hyperthermic perfusion for primary and recurrent melanomas of the extremities. Recent Results Cancer Res 86:277–282 Karakousis CP, Balch CM, Urist MM, Ross MM, Smith TJ, Bartolucci AA (1996) Local recurrence in malignant melanoma: long-term results of the multiinstitutional randomized surgical trial. Ann Surg Oncol 3:446–452 Kaufman HL, Ruby CE, Hughes T, Slingluff CL Jr (2014) Current status of granulocyte-macrophage colonystimulating factor in the immunotherapy of melanoma. J Immunother Cancer 2:11 Klaase JM, Kroon BB, Van Geel AN, Eggermont AM, Franklin HR, Hart AA (1994a) Prognostic factors for tumor response and limb recurrence-free interval in patients with advanced melanoma of the limbs treated with regional isolated perfusion with melphalan. Surgery 115:39–45 Klaase JM, Kroon BB, Van Geel AN, Van Wijk J, Franklin HR, Eggermont AM, Hart AA (1994b) Limb recurrence-free interval and survival in patients with recurrent melanoma of the extremities treated with normothermic isolated perfusion. J Am Coll Surg 178:564–572 Klaase JM, Kroon BB, Van Geel BN, Eggermont AM, Franklin HR, Hart GA (1994c) Patient- and treatmentrelated factors associated with acute regional toxicity after isolated perfusion for melanoma of the extremities. Am J Surg 167:618–620 Klaase JM, Kroon BB, Eggermont AM, Van Geel AN, Schraffordt Koops H, Oldhoff J, Lienard D, Lejeune FJ, Berkel R, Franklin HR (1995) A

734 retrospective comparative study evaluating the results of mild hyperthermic versus controlled normothermic perfusion for recurrent melanoma of the extremities. Eur J Cancer 31A:58–63 Koops HS, Vaglini M, Suciu S, Kroon BB, Thompson JF, Gohl J, Eggermont AM, Di Filippo F, Krementz ET, Ruiter D, Lejeune FJ (1998) Prophylactic isolated limb perfusion for localized, high-risk limb melanoma: results of a multicenter randomized phase III trial. European Organization for Research and Treatment of Cancer Malignant Melanoma Cooperative Group Protocol 18832, the World Health Organization Melanoma Program Trial 15, and the North American Perfusion Group Southwest Oncology Group-8593. J Clin Oncol 16:2906–2912 Krementz ET, Creech O Jr, Ryan RF, Reemtsma K (1962) An appraisal of cancer chemotherapy by regional perfusion. Ann Surg 156:417–428 Krementz ET, Carter RD, Sutherland CM, Muchmore JH, Ryan RF, Creech O (1994) Regional chemotherapy for melanoma. A 35-year experience. Ann Surg 220:520–535 Kroon BB (1988) Regional isolation perfusion in melanoma of the limbs; accomplishments, unsolved problems, future. Eur J Surg Oncol 14:101–110 Kroon BB, Klaase JM, Van Geel BN, Eggermont AM, Franklin HR, Van Dongen JA (1993) Results of a double perfusion schedule with melphalan in patients with melanoma of the lower limb. Eur J Cancer 29A:325–328 Kroon HM, Moncrieff M, Kam PC, Thompson JF (2008) Outcomes following isolated limb infusion for melanoma. A 14-year experience. Ann Surg Oncol 15:3003–3013 Kroon HM, Lin DY, Kam PC, Thompson JF (2009a) Isolated limb infusion as palliative treatment for advanced limb disease in patients with AJCC stage IV melanoma. Ann Surg Oncol 16:1193–1201 Kroon HM, Moncrieff M, Kam PC, Thompson JF (2009b) Factors predictive of acute regional toxicity after isolated limb infusion with melphalan and actinomycin D in melanoma patients. Ann Surg Oncol 16:1184–1192 Kroon HM, Huismans AM, Kam PC, Thompson JF (2014) Isolated limb infusion with melphalan and actinomycin D for melanoma: a systematic review. J Surg Oncol 109:348–351 Kroon HM, Coventry BJ, Giles MH, Henderson MA, Speakman D, Wall M, Barbour A, Serpell J, Paddle P, Coventry AG, Sullivan T, Smithers BM, Thompson JF (2016) Australian multicenter study of isolated limb infusion for melanoma. Ann Surg Oncol 23:1096–1103 Laurenzi L, Natoli S, Di Filippo F, Calamaro A, Centulio F, Anza M, Cavaliere F, Marcelli ME, Garinei R, Arcuri E (2004) Systemic and haemodynamic toxicity after isolated limb perfusion (ILP) with TNF-alpha. J Exp Clin Cancer Res 23:225–231 Lawson DH, Lee S, Zhao F, Tarhini AA, Margolin KA, Ernstoff MS, Atkins MB, Cohen GI, Whiteside TL, Butterfield LH, Kirkwood JM (2015) Randomized,

M. C. Perez et al. placebo-controlled, phase III trial of yeast-derived Granulocyte-Macrophage Colony-Stimulating Factor (GM-CSF) versus peptide vaccination versus GMCSF plus peptide vaccination versus placebo in patients with no evidence of disease after complete surgical resection of locally advanced and/or stage IV melanoma: a trial of the Eastern Cooperative Oncology Group-American College of Radiology Imaging Network Cancer Research Group (E4697). J Clin Oncol 33:4066–4076 Lee YT (1980) Loco-regional recurrent melanoma: I natural history. Cancer Treat Rev 7:59–72 Lejeune FJ, Deloof T, Ewalenko P, Fruhling J, Jabri M, Mathieu M, Nogaret JM, Verhest A (1983) Objective regression of unexcised melanoma in-transit metastases after hyperthermic isolation perfusion of the limbs with melphalan. Recent Results Cancer Res 86:268–276 Lewis KD, Robinson WA, Mccarter M, Pearlman N, O’day SJ, Anderson C, Amatruda TT, Baron A, Zeng C, Becker M, Dollarhide S, Matijevich K, Gonzalez R (2006) Phase II multicenter study of neoadjuvant biochemotherapy for patients with stage III malignant melanoma. J Clin Oncol 24:3157–3163 Li S, Sheng X, Si L, Cui C, Kong Y, Mao L, Lian B, Tang B, Yan X, Wang X, Chi Z, Guo J (2018) Outcomes and predictive factors of isolated limb infusion for patients with in-transit melanoma in China. Ann Surg Oncol 25:885–893 Lienard D, Lejeune FJ, Ewalenko P (1992) In transit metastases of malignant melanoma treated by high dose rTNF alpha in combination with interferongamma and melphalan in isolation perfusion. World J Surg 16:234–240 Lienard D, Eggermont AM, Koops HS, Kroon B, Towse G, Hiemstra S, Schmitz P, Clarke J, Steinmann G, Rosenkaimer F, Lejeune FJ (1999) Isolated limb perfusion with tumour necrosis factor-alpha and melphalan with or without interferon-gamma for the treatment of in-transit melanoma metastases: a multicentre randomized phase II study. Melanoma Res 9:491–502 Liu BL, Robinson M, Han ZQ, Branston RH, English C, Reay P, Mcgrath Y, Thomas SK, Thornton M, Bullock P, Love CA, Coffin RS (2003) ICP34.5 deleted herpes simplex virus with enhanced oncolytic, immune stimulating, and anti-tumour properties. Gene Ther 10:292–303 Long GV, Dummer R, Ribas A, Puzanov I, Vanderwalde A, Andtbacka RHI, Michielin O, Olszanski AJ, Malvehy J, Cebon JS, Fernandez E, Kirkwood JM, Gajewski T, Gause CK, Chen L, Gorski K, Anderson A, Kaufman DR, Chou J, Hodi FS (2016) Efficacy analysis of MASTERKEY-265 phase lb study of talimogene laherparepvec (T-VEC) and pembrolizumab (pembro) for unresectable stage IIIBIV melanoma. J Clin Oncol 34. ASCO 2016 Abstract TPS9598 Luck JM (1956) Action of p-[di(2-chloroethyl)]-amino-lphenylalanine on Harding-Passey mouse melanoma. Science 123:984–985

Local and Recurrent Regional Metastases of Melanoma

735

Mali B, Jarm T, Snoj M, Sersa G, Miklavcic D (2013) Antitumor effectiveness of electrochemotherapy: a systematic review and meta-analysis. Eur J Surg Oncol 39:4–16 Mccarthy WH, Shaw HM, Thompson JF, Milton GW (1988) Time and frequency of recurrence of cutaneous stage I malignant melanoma with guidelines for followup study. Surg Gynecol Obstet 166:497–502 Meloni G, Foa R, Capria S, Tosti S, Vignetti M, Gavosto F, Mandelli F (1992) IL-2 for the treatment of acute leukemias. Leukemia 6(Suppl 2):28–30 Menzies AM, Rozeman EA, Amaria R, Scolyer R, Tetzlaff M, Guminski A, Davies M, Blank C, Wargo J, Long G (2017) Preliminary results from the international neoadjuvant melanoma consortium (INMC), abstract 9581. J Clin Oncol 35:9581 Miller EJ, Daly JM, Synnestvedt M, Schultz D, Elder D, Guerry DT (1992) Loco-regional nodal relapse in melanoma. Surg Oncol 1:333–340 Minor DR, Allen RE, Alberts D, Peng YM, Tardelli G, Hutchinson J (1985) A clinical and pharmacokinetic study of isolated limb perfusion with heat and melphalan for melanoma. Cancer 55:2638–2644 Mir LM, Glass LF, Sersa G, Teissie J, Domenge C, Miklavcic D, Jaroszeski MJ, Orlowski S, Reintgen DS, Rudolf Z, Belehradek M, Gilbert R, Rols MP, Belehradek J Jr, Bachaud JM, Deconti R, Stabuc B, Cemazar M, Coninx P, Heller R (1998) Effective treatment of cutaneous and subcutaneous malignant tumours by electrochemotherapy. Br J Cancer 77:2336–2342 Miura JT, Zager JS (2018) Intralesional therapy as a treatment for locoregionally metastatic melanoma. Expert Rev Anticancer Ther 18:399–408 Monsour PD, Sause WT, Avent JM, Noyes RD (1993) Local control following therapeutic nodal dissection for melanoma. J Surg Oncol 54:18–22 Morgan DA, Ruscetti FW, Gallo R (1976) Selective in vitro growth of T lymphocytes from normal human bone marrows. Science 193:1007–1008 Morton D, Eilber FR, Malmgren RA, Wood WC (1970) Immunological factors which influence response to immunotherapy in malignant melanoma. Surgery 68:158–163. discussion 163–4 Morton DL, Eilber FR, Holmes EC, Hunt JS, Ketcham AS, Silverstein MJ, Sparks FC (1974) BCG immunotherapy of malignant melanoma: summary of a seven-year experience. Ann Surg 180:635–643 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Elashoff R, Essner R, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Reintgen DS, Coventry BJ, Glass EC, Wang HJ (2006) Sentinel-node biopsy or nodal observation in melanoma. N Engl J Med 355:1307–1317 Moschos SJ, Edington HD, Land SR, Rao UN, Jukic D, Shipe-Spotloe J, Kirkwood JM (2006) Neoadjuvant treatment of regional stage IIIB melanoma with highdose interferon alfa-2b induces objective tumor regression in association with modulation of tumor

infiltrating host cellular immune responses. J Clin Oncol 24:3164–3171 Mousavi H, Zhang X, Gillespie S, Wachter E, Hersey P (2006) Rose Bengal induces dual modes of cell death in melanoma cells and has clinical activity against melanoma. Melanoma Res 16:S8 Muilenburg DJ, Beasley GM, Thompson ZJ, Lee JH, Tyler DS, Zager JS (2015) Burden of disease predicts response to isolated limb infusion with melphalan and actinomycin D in melanoma. Ann Surg Oncol 22:482–488 Mullinax JE, Kroon HM, Thompson JF, Nath N, Mosca PJ, Farma JM, Bhati R, Hardmann D, Sileno S, O’donoghue C, Perez M, Naqvi SM, Chen YA, Gonzalez RJ, Zager JS (2017) Isolated limb infusion as a limb salvage strategy for locally advanced extremity sarcoma. J Am Coll Surg 224:635–642 O’donoghue C, Perez MC, Mullinax JE, Hardman D, Sileno S, Naqvi SMH, Kim Y, Gonzalez RJ, Zager JS (2017) Isolated limb infusion: a single-center experience with over 200 infusions. Ann Surg Oncol 24:3842–3849 Old LJ, Clarke DA, Benacerraf B (1959) Effect of Bacillus Calmette-Guerin infection on transplanted tumours in the mouse. Nature 184(Suppl 5):291–292 Pawlik TM, Ross MI, Johnson MM, Schacherer CW, Mcclain DM, Mansfield PF, Lee JE, Cormier JN, Gershenwald JE (2005) Predictors and natural history of in-transit melanoma after sentinel lymphadenectomy. Ann Surg Oncol 12:587–596 Posner MC, Lienard D, Lejeune FJ, Rosenfelder D, Kirkwood J (1995) Hyperthermic isolated limb perfusion with tumor necrosis factor alone for melanoma. Cancer J Sci Am 1:274–280 Puzanov I, Milhem MM, Minor D, Hamid O, Li A, Chen LS, Chastain M, Gorski KS, Anderson A, Chou J, Kaufman HL, Andtbacka RHI (2016) Talimogene laherparepvec in combination with Ipilimumab in previously untreated, unresectable stage IIIB-IV melanoma. J Clin Oncol 34:2619–U109 Rabe E, Stucker M, Ottillinger B (2010) Water displacement leg volumetry in clinical studies – a discussion of error sources. BMC Med Res Methodol 10:5 Ridolfi L, Ridolfi R (2002) Preliminary experiences of intralesional immunotherapy in cutaneous metastatic melanoma. Hepato-Gastroenterology 49:335–339 Robert C, Thomas L, Bondarenko I, O'day S, Weber J, Garbe C, Lebbe C, Baurain JF, Testori A, Grob JJ, Davidson N, Richards J, Maio M, Hauschild A, Miller WH Jr, Gascon P, Lotem M, Harmankaya K, Ibrahim R, Francis S, Chen TT, Humphrey R, Hoos A, Wolchok JD (2011) Ipilimumab plus dacarbazine for previously untreated metastatic melanoma. N Engl J Med 364:2517–2526 Robert C, Long GV, Brady B, Dutriaux C, Maio M, Mortier L, Hassel JC, Rutkowski P, Mcneil C, Kalinka-Warzocha E, Savage KJ, Hernberg MM, Lebbe C, Charles J, Mihalcioiu C, Chiarion-Sileni V, Mauch C, Cognetti F, Arance A, Schmidt H,

736 Schadendorf D, Gogas H, Lundgren-Eriksson L, Horak C, Sharkey B, Waxman IM, Atkinson V, Ascierto PA (2015a) Nivolumab in previously untreated melanoma without BRAF mutation. N Engl J Med 372:320–330 Robert C, Schachter J, Long GV, Arance A, Grob JJ, Mortier L, Daud A, Carlino MS, Mcneil C, Lotem M, Larkin J, Lorigan P, Neyns B, Blank CU, Hamid O, Mateus C, Shapira-Frommer R, Kosh M, Zhou H, Ibrahim N, Ebbinghaus S, Ribas A (2015b) Pembrolizumab versus Ipilimumab in advanced melanoma. N Engl J Med 372:2521–2532 Rosenberg SA, Lotze MT, Yang JC, Aebersold PM, Linehan WM, Seipp CA, White DE (1989) Experience with the use of high-dose interleukin-2 in the treatment of 652 cancer patients. Ann Surg 210:474–484. discussion 484–5 Rosin RD, Westbury G (1980) Isolated limb perfusion for malignant melanoma. Practitioner 224:1031–1036 Rozeman EA, Blank CU, Van Akkooi A, Kvistborg P, Fanchi L, Van Thienen J, Stegenga B, Lamon B, Haanen J, Schumacher T (2017) Neoadjuvant Ipilimumab + nivolumab (IPI+NIVO) in palpable stage III mleanoma: updated data from the OpACIN trial and first immunological analyses, abstract 9586. J Clin Oncol 35:9586 Ruschulte H, Shi S, Tseng WW, Kolodzie K, Crawford PC, Schneider DB, Kashani-Sabet M, Minor D, Apfel C, Leong SP (2013) Anesthesia management of patients undergoing hyperthermic isolated limb perfusion with melphalan for melanoma treatment: an analysis of 17 cases. BMC Anesthesiol 13:15 Ryan RF, Winblad JN, Krementz ET, Creech O (1958) Treatment of malignant neoplasms with chemotherapeutic agents utilizing a pump-oxygenator: technics and early results. Bull Tulane Univ Med Fac 17:133 Sanki A, Kam PC, Thompson JF (2007) Long-term results of hyperthermic, isolated limb perfusion for melanoma: a reflection of tumor biology. Ann Surg 245:591–596 Schrama D, Reisfeld RA, Becker JC (2006) Antibody targeted drugs as cancer therapeutics. Nat Rev Drug Discov 5:147–159 Scott RN, Blackie R, Kerr DJ, Hughes J, Burnside G, Mackie RM, Byrne DS, Mckay AJ (1992) Melphalan concentration and distribution in the tissues of tumourbearing limbs treated by isolated limb perfusion. Eur J Cancer 28A:1811–1813 Senzer NN, Kaufman H, Amatruda T, Nemunaitis M, Daniels G, Glaspy J, Goldsweig H, Coffin RS, Nemunaitis J, Ii OMP (2009) Phase II clinical trial with a second generation, GM-CSF encoding, oncolytic herpesvirus in unresectable metastatic melanoma. J Clin Oncol 27:5763–5771 Shafren DR, Au GG, Nguyen T, Newcombe NG, Haley ES, Beagley L, Johansson ES, Hersey P, Barry RD (2004) Systemic therapy of malignant human melanoma tumors by a common cold-producing enterovirus, coxsackievirus a21. Clin Cancer Res 10:53–60 Shah GD, Socci ND, Gold JS, Wolchok JD, Carvajal RD, Panageas KS, Viale A, Brady MS, Coit DG, Chapman

M. C. Perez et al. PB (2010) Phase II trial of neoadjuvant temozolomide in resectable melanoma patients. Ann Oncol 21:1718–1722 Si Z, Hersey P, Coates AS (1996) Clinical responses and lymphoid infiltrates in metastatic melanoma following treatment with intralesional GM-CSF. Melanoma Res 6:247–255 Siegel RL, Miller KD, Jemal A (2017) Cancer statistics, 2017. CA Cancer J Clin 67:7–30 Skene AI, Bulman AS, Williams TR, Thomas JM, Westbury G (1990) Hyperthermic isolated perfusion with melphalan in the treatment of advanced malignant melanoma of the lower limb. Br J Surg 77:765–767 Sloot S, Rashid OM, Sarnaik AA, Zager JS (2016) Developments in intralesional therapy for metastatic melanoma. Cancer Control 23:12–20 Spiess BD (2008) Treating heparin resistance with antithrombin or fresh frozen plasma. Ann Thorac Surg 85:2153–2160 Spitler LE, Weber RW, Allen RE, Meyer J, Cruickshank S, Garbe E, Lin HY, Soong SJ (2009) Recombinant human granulocyte-macrophage colony-stimulating factor (GM-CSF, sargramostim) administered for 3 years as adjuvant therapy of stages II(T4), III, and IV melanoma. J Immunother 32:632–637 Stamatiou D, Ioannou CV, Kontopodis N, Michelakis D, Perisinakis K, Lasithiotakis K, Zoras O (2017) Hyperthermic isolated limb perfusion. The switch from Steinmann pins to Omni-tract assisted isolation. J Surg Res 213:147–157 Stanley G, Sundarakumaran R, Crowley R, Kavanah M (2000) Desflurane as a marker of limb-to-systemic leak during hyperthermic isolated limb perfusion. Anesthesiology 93:574–576 Stehlin JS, Giovanella BC, De Ipolyi PD, Muenz LR, Anderson RF (1975) Results of hyperthermic perfusion for melanoma of the extremities. Surg Gynecol Obstet 140:339–348 Steinman J, Ariyan C, Rafferty B, Brady MS (2014) Factors associated with response, survival, and limb salvage in patients undergoing isolated limb infusion. J Surg Oncol 109:405–409 Sutherland CM, Mather FJ, Krementz ET (1987) Factors influencing the survival of patients with regional melanoma of the extremity treated by perfusion. Surg Gynecol Obstet 164:111–118 Tarhini AA, Edington H, Butterfield LH, Lin Y, Shuai Y, Tawbi H, Sander C, Yin Y, Holtzman M, Johnson J, Rao UN, Kirkwood JM (2014) Immune monitoring of the circulation and the tumor microenvironment in patients with regionally advanced melanoma receiving neoadjuvant Ipilimumab. PLoS One 9:e87705 Thompson JF, Kam PC, Waugh RC, Harman CR (1998) Isolated limb infusion with cytotoxic agents: a simple alternative to isolated limb perfusion. Semin Surg Oncol 14:238–247 Thompson JF, Hersey P, Wachter E (2008) Chemoablation of metastatic melanoma using intralesional Rose Bengal. Melanoma Res 18:405–411

Local and Recurrent Regional Metastases of Melanoma

737

Thompson JF, Agarwala SS, Smithers BM, Ross MI, Scoggins CR, Coventry BJ, Neuhaus SJ, Minor DR, Singer JM, Wachter EA (2015) Phase 2 study of intralesional PV-10 in refractory metastatic melanoma. Ann Surg Oncol 22:2135–2142 Topalian SL, Hodi FS, Brahmer JR, Gettinger SN, Smith DC, Mcdermott DF, Powderly JD, Carvajal RD, Sosman JA, Atkins MB, Leming PD, Spigel DR, Antonia SJ, Horn L, Drake CG, Pardoll DM, Chen L, Sharfman WH, Anders RA, Taube JM, Mcmiller TL, Xu H, Korman AJ, Jure-Kunkel M, Agrawal S, Mcdonald D, Kollia GD, Gupta A, Wigginton JM, Sznol M (2012) Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. N Engl J Med 366:2443–2454 Turaga KK, Beasley GM, Kane JM 3rd, Delman KA, Grobmyer SR, Gonzalez RJ, Letson GD, Cheong D, Tyler DS, Zager JS (2011) Limb preservation with isolated limb infusion for locally advanced nonmelanoma cutaneous and soft-tissue malignant neoplasms. Arch Surg 146:870–875 Warso MA, Das Gupta TK (1994) Melanoma recurrence in a previously dissected lymph node basin. Arch Surg 129:252–255 Weide B, Eigentler TK, Pflugfelder A, Zelba H, Martens A, Pawelec G, Giovannoni L, Ruffini PA, Elia G, Neri D, Gutzmer R, Becker JC, Garbe C (2014) Intralesional treatment of stage III metastatic melanoma patients with L19-IL2 results in sustained clinical and systemic immunologic responses. Cancer Immunol Res 2:668–678 Weide B, Neri D, Elia G (2017) Intralesional treatment of metastatic melanoma: a review of therapeutic options. Cancer Immunol Immunother 66:647–656 Weitman ES, Perez M, Thompson JF, Andtbacka RHI, Dalton J, Martin ML, Miller T, Gwaltney C, Sarson

D, Wachter E, Zager JS (2018) Quality of life patientreported outcomes for locally advanced cutaneous melanoma. Melanoma Res 28:134–142 Wieberdink J, Benckhuysen C, Braat RP, Van Slooten EA, Olthuis GA (1982) Dosimetry in isolation perfusion of the limbs by assessment of perfused tissue volume and grading of toxic tissue reactions. Eur J Cancer Clin Oncol 18:905–910 Wolchok JD, Chiarion-Sileni V, Gonzalez R, Rutkowski P, Grob JJ, Cowey CL, Lao CD, Wagstaff J, Schadendorf D, Ferrucci PF, Smylie M, Dummer R, Hill A, Hogg D, Haanen J, Carlino MS, Bechter O, Maio M, MarquezRodas I, Guidoboni M, Mcarthur G, Lebbe C, Ascierto PA, Long GV, Cebon J, Sosman J, Postow MA, Callahan MK, Walker D, Rollin L, Bhore R, Hodi FS, Larkin J (2017) Overall survival with combined nivolumab and Ipilimumab in advanced melanoma. N Engl J Med 377:1345–1356 Wong JH, Cagle LA, Kopald KH, Swisher SG, Morton DL (1990) Natural history and selective management of in transit melanoma. J Surg Oncol 44:146–150 Wong J, Chen YA, Fisher KJ, Zager JS (2013) Isolated limb infusion in a series of over 100 infusions: a singlecenter experience. Ann Surg Oncol 20:1121–1127 Wong J, Chen YA, Fisher KJ, Beasley GM, Tyler DS, Zager JS (2014) Resection of residual disease after isolated limb infusion (ILI) is equivalent to a complete response after ILI-alone in advanced extremity melanoma. Ann Surg Oncol 21:650–655 World Health Organization (1979) WHO handbook for reporting results of cancer treatment. World Health Organization, Geneva Zogakis TG, Essner R, Wang HJ, Foshag LJ, Morton DL (2007) Natural history of melanoma in 773 patients with tumor-negative sentinel lymph nodes. Ann Surg Oncol 14:1604–1611

Radiotherapy for Primary and Regional Melanoma Angela M. Hong and Graham Stevens

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 739 Lentigo Maligna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740 Primary Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740 Adjuvant RT After Wide Local Excision of a Primary Melanoma . . . . . . . . . . . . . . . . 741 Adjuvant RT for Resected Stage III Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742 RT for Inoperable Regional Node Metastases and In-Transit Metastases . . . . . . . . 743 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745

Abstract

Radiation therapy (RT) plays a role in the management of cutaneous melanoma in definitive, adjuvant, and palliative settings. The role for RT in early stage primary melanoma, for which the treatment is adequate excision, is limited. Desmoplastic melanoma and those with neurotropism invasion are exceptions, due to frequent local recurrence. The greatest controversy regarding RT lies with its use in stage III

A. M. Hong (*) Radiation Oncology, Melanoma Institute Australia, The University of Sydney, North Sydney, NSW, Australia e-mail: [email protected] G. Stevens Radiation Oncology, Orange General Hospital, Orange, NSW, Australia Bathurst Rural Clinical School, School of Medicine, Western Sydney University, Bathurst, NSW, Australia e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_25

disease, particularly as postoperative treatment after regional lymph node dissection. This chapter presents the evidence and illustrative cases for RT in locoregional melanoma.

Introduction Melanoma was thought to be radioresistant based on in vitro cell studies which showed a broad shoulder in the cell survival curves (Barranco et al. 1971; Dewey 1971; Doss and Memula 1982), implying better response to higher dose per fraction (Rofstad 1986). Early clinical observations using large fractions of radiation per fraction (hypofractionation) supported these laboratory findings. A study of 35 patients showed a complete response rate of 9% when 5 Gy per fraction was used. The response rate increased to 50% when 5 Gy per fraction was used. However, other 739

740

studies did not confirm a difference between various large fraction treatment schedules, and some single institution series reported similar outcomes with conventional fractionation schedules (Overgaard 1986; Bentzen et al. 1989; Chang et al. 2006). These conflicting results led to a multicenter randomized phase III trial by the Radiation Therapy Oncology Group (RTOG). A total of 137 patients with metastatic melanoma at any site other than abdomen or brain were randomly assigned to 32 Gy in four fractions given at weekly intervals (Arm 1) or 50 Gy in 20 fractions given 5 days per week over 4 weeks (Arm 2) (Sause et al. 1991). This study failed to show better tumor control with larger fraction sizes, with no difference in the clinical response rates (24% and 23%, respectively, for complete response and 36% and 34% for partial response). Acute radiation toxicity was greater in Arm 1 (three grade 4 reactions and three grade 3 reactions) compared with Arm 2 (four grade 3 reactions). However, the durations of tumor control, late radiation effects, and survival were not reported. The results from this trial are difficult to interpret due to lack of long-term control rate and toxicity rate. In general RT schedule is moving away from hypofractionated treatment schedules for most indications as the fraction size (dose per fraction) is the most important factor determining late radiation effects; a smaller fraction size (2 Gy per fraction) gives better sparing of the late effect than larger fraction size (>3 Gy per fraction). This is an important consideration especially when delivering RT to skin and soft tissue. However, in the setting of stereotactic RT, extreme hypofractionation with high ablation dose is used.

A. M. Hong and G. Stevens

Surgical resection with a 5 mm margin is the standard treatment approach. However, when lesions involve large areas of the face, excision and reconstruction to achieve satisfactory cosmetic and functional results becomes difficult. Superficial RT is an alternative treatment which has been used in Europe for many years as the primary treatment with a success rate of 95% (Harwood and Cummings 1981). In this early study, very high doses of soft x-rays (Genz rays, 10 to 20 KeV) were used to deliver surface doses of 20 Gy once weekly for four to five treatments. These low-energy x-rays only penetrate tissues superficially, attenuating to 50% of the surface dose at a depth of 1–1.3 mm. Very few centers are now equipped with this type of radiation therapy machine due to the very limited use in other conditions. Using commonly available superficial energy photon (50–125 KeV) or low-energy electron, a typical current dose schedule is 50–54 Gy in 2 Gy fraction treating to a depth of 5 mm. To maximize long-term cosmetic and functional outcome, the standard dose fractionation of 2 Gy per fraction is recommended. An example is shown in Fig. 1. However, the dose fractionation can be varied depending on patient factors (age, comorbidity) and tumor factors (size, location). Hypofractionated schedules (higher dose per fraction in fewer number of fractions) can be tailored to the individual patients. Alternative fractionations are 30–33 Gy in 5–6 fractions (two fractions per week), 40 Gy in 10 fractions, and 45 Gy in 15 fractions. In a large cohort series of 593 patients with lentigo maligna and early lentigo maligna melanoma treated with superficial energy RT, the complete clearance rate after definitive treatment with RT, partial excision followed by RT, and wide excision followed by RT were 83%, 90%, and 97%, respectively (Hedblad and Mallbris 2012).

Lentigo Maligna Lentigo maligna is a subtype of melanoma in situ with a slow growth rate and low potential for progression to invasive disease. It typically occurs on head and neck area in severely sun-damaged skin of the elderly. Lentigo maligna melanoma refers to invasive melanoma within a lentigo melanoma.

Primary Melanoma Wide local resection is the definitive curative treatment of primary cutaneous melanoma. RT has a very limited role as the initial therapy of primary melanoma and is only indicated if complete surgery excision is not feasible (e.g.,

Radiotherapy for Primary and Regional Melanoma

proximity to vital structures such as the eye or inoperability because of medical comorbidities). There are limited modern data on the effectiveness of RT for primary melanoma. In a very old series of 95 cases, high radiation doses (100–110 Gy) were delivered in 6 Gy fractions using superficial energy radiation (60 KeV x-rays) (Hellriegel 1963). The 5-year survival rate was 68%, a value similar to that obtained with wide local excision.

Adjuvant RT After Wide Local Excision of a Primary Melanoma Adjuvant RT may be considered after wide local excision where the local recurrence rate is high. The main risk factors for local recurrence following excision are microsatellitosis, lymphovascular invasion, desmoplasia, perineural invasion, involved margins, and multiple local recurrences. Desmoplastic melanoma accounts for 1–4% of all melanoma. It is frequently associated with perineural spread (neurotropism) and is associated with an increased local failure rate following excision. Although there are no randomized trials defining the role of adjuvant RT after complete

Fig. 1 (a) A 60-year-old lady with a biopsy-proven recurrent lentigo maligna 2 years after local excision. Instead of re-excision and reconstruction, she was treated with 54 Gy

741

resection of desmoplastic melanoma with or without neurotropism, multiple retrospective series support the use of adjuvant RT (Chen et al. 2008; Guadagnolo et al. 2014; Strom et al. 2014; Varey et al. 2017). The largest series, from the Melanoma Institute Australia, compared the outcomes of 671 neurotropic melanomas (72% of which had desmoplastic features) with a matched cohort of 718 non-neurotropic melanomas (Varey et al. 2017). Eighty-two patients with neurotropic melanoma were given adjuvant RT to the primary site after local excision. At a median follow up of 3.5 years, adjuvant RT halved the risk of local recurrence if the pathological margins were less than 8 mm (hazard ratio 0.48, 95% CI 0.27–0.87). In the multivariate analysis, RT was associated with a significant reduction in the overall risk of recurrence (hazard ratio 0.51, 95% 0.29–0.87), particularly at the primary site (hazard ratio 0.30, 95% CI 0.13–0.69) and probably in the regional nodal site (hazard ratio 0.41, 95% CI 0.17–0.98) but not at distant sites (hazard ratio 0.60, 95% CI 0.29–1.24, p = 0.17). In another series of 277 patients with resected desmoplastic melanoma, 113 patients were treated adjuvant RT (Strom et al. 2014). At a median follow-up of 43 months,

in 27 fractions of RT. (b) She remained in complete response 3 years after RT

742

the overall 5-year local control rate for patients treated with adjuvant RT was significantly improved compared with those who did not receive RT (86% versus 46%, hazard ratio 0.15, 95% CI 0.06–0.39). In the multivariable analysis, RT was an independent predictor of local control (hazard ratio 0.15, 95% CI 0.06–0.39). For those patients with positive resection margins, the local recurrence rate was 54%, and the addition of RT reduced the local recurrence rate to 14%. For those with negative margins, local control was also improved with adjuvant RT (95% versus 76%). Based on these retrospective data showing improvement in local control, adjuvant RT should be considered for neurotropic melanoma that has not been widely excised (pathological margin 4 mm) without clinical evidence of lymph node metastasis or stage III nodal metastatic disease. Completion LND was required for patients with a positive SLN. This trial enrolled 880 patients and was unblinded after a median follow-up of 1.3 years, when an interim analysis revealed the superiority of HDI in terms of both RFS (HR 1.47, 95% CI, 1.14–1.90) and OS (HR 1.52, 95% CI, 1.07–2.15). These results

demonstrated a 33% lower risk of relapse and death for IFN-α2b compared with the GMK vaccine (Kirkwood et al. 2001b). At 2.1 years, HDI remained superior to the GMK vaccine in terms of both RFS (HR = 1.33, P = 0.006) and OS (HR = 1.32, p = 0.04) (Kirkwood et al. 2004). Concerns that the GMK vaccine might actually have been detrimental to patient survival were not borne out in a randomized trial. EORTC 18961 compared the efficacy of GM2-KLH/QS-21 vaccination vs. observation and found no significant difference in RFS, distant metastasis-free survival (DMFS), or OS (Eggermont et al. 2013).

Adjuvant Systemic Therapy for High-Risk Melanoma Patients

E2696 attempted to build upon IFN with GMK and randomized 107 patients to treatment with GMK vaccine plus concurrent HDI (arm A), GMK vaccine plus sequential HDI (arm B), or GMK vaccine alone (arm C) (Kirkwood et al. 2001a). At 24 months, the RFS in the HDI-containing arms was higher than in patients who received GMK alone (HR = 1.75 for arm vs. arm A; HR =1.96 for arm C vs. arm B), although this effect lost statistical significance on further follow-up (Kirkwood et al. 2004). Recent meta-analyses have confirmed the previously published benefits of HDI. A metaanalysis of 12 RCTs demonstrated a significant reduction in recurrence in patients treated with IFN compared to observation. No OS benefit was demonstrated (Wheatley et al. 2003). Another meta-analysis of 15 RCTs, 11 of which had individual patient data, demonstrated a significant improvement in EFS (HR = 0.86, p < 0.00001) and OS (HR = 0.90, P = 0.003) with IFN-α, without benefit of higher doses compared to lower doses (Ives et al. 2017). S0008 was an intergroup phase III trial of biochemotherapy that randomized 432 patients with high-risk resected melanoma (stage IIIA (N2a)-IIIC) to treatment with either standard HDI or biochemotherapy. At median follow-up of 7.2 years, biochemotherapy was associated with improved RFS (HR = 0.75; 95% CI, 0.58 to 0.97; p = 0.015) and increased toxicity, but not improved OS (Flaherty et al. 2014).

Role of Dose and Duration of IFN-a Therapy in Melanoma Low-dose (Kleeberg et al. 2004; Cascinelli et al. 2001; Kirkwood et al. 2000; Hancock et al. 2004; Cameron et al. 2001) and intermediate-dose IFN regimens have been tested in the adjuvant setting for treatment of melanoma (Eggermont et al. 2005, 2013). The Sunbelt Melanoma Trial evaluated the role of HDI or completion LND in patients with melanoma staged by SLN biopsy (McMasters et al. 2016). In Sunbelt Protocol A, patients with a single positive SLN without extranodal extension underwent LND and were

753

randomized to observation vs. 1 year of HDI. Patients with more than one positive SLN or extranodal extension were treated with HDI in a non-randomized fashion. In Protocol B, patients with tumor-negative SLN by standard histopathology and immunohistochemistry underwent molecular staging by reverse transcriptase polymerase chain reaction (RT-PCR) to detect melanoma-specific mRNA. RT-PCR-positive patients were then randomized to observation vs. LND vs. LND plus HDI. Primary endpoints were DFS and OS. There were no significant differences in DFS or OS detected in any aspect of the protocol, although the study was underpowered and did not ever meet its accrual goals (McMasters et al. 2016). Several studies have attempted to evaluate the optimal duration of adjuvant therapy. A Hellenic Oncology Group phase III RCT randomized 353 patients to 4 weeks induction (group A) vs. 4 weeks induction followed by 1 year of HDI (group B) (Pectasides et al. 2009). There was no significant difference in OS or DFS between the arms, although only three-quarters of the established dosage was used for IFN induction, and the maintenance regimen utilized a 10 MU/dose rather than the standard 10 MU/m2/ dose (Pectasides et al. 2009) employed in other HDI trials. A DeCOG study of lower-dose IFN-α2a for 5 years vs. 18 months also showed no difference in RFS or OS in 850 randomized patients (Hauschild et al. 2010). Similarly, the EORTC 18952 trial randomized 1388 patients and found no significant difference between two intermediate dosages of IFN given for 2 years vs. 1 year (Eggermont et al. 2005). The Nordic trial randomized 855 patients to observation (group A) vs. 4 weeks induction and 12 months maintenance (group B) vs. 4 weeks induction and 24 months maintenance (group C). This study demonstrated that 1 year of adjuvant intermediate-dose interferon significantly improved RFS compared to 2 years, without OS benefit (Hansson et al. 2011). The EORTC trial 18,991 testing pegylated IFN-α2b (peg-IFN) for up to 5 years vs. observation and demonstrated improved RFS but neither improved OS nor DMFS (Eggermont et al. 2008).

754

Nonetheless, in 2011 the peg-IFN regimen was approved by the FDA as an alternative to the standard HDI regimen, although it never achieved widespread use. A retrospective analysis of EORTC 18592 and EORTC 18991 suggested that tumor ulceration may be predictive of therapeutic benefit of IFN-α (see below). EORTC 18081 will therefore compare the benefit of peg-IFN vs. observation in node-negative patients with ulcerated primaries and Breslow thickness >1 mm. Together, these results established HDI as administered in E1684, E1690, and E1694 (with an intravenous 4-week induction phase and 48 weeks of maintenance therapy) as the standard of care for adjuvant therapy of resected high-risk melanoma, and it remained so until the advent of newer agents. Experience has demonstrated that this regimen is safe and tolerable for most patients. Close follow-up and aggressive management of side effects is mandatory. Persisting functional decrements of more than 25–33% should prompt dose interruption or dosage reduction until day-to-day functions have resumed a level that is at least 60% of normal. Evaluations for toxicity and if necessary dose modification are recommended weekly during induction therapy and monthly during maintenance (for at least 3 months) and at 3 monthly intervals to the conclusion of the year of therapy. Quality-of-life analyses using the Q-TWiST methodology have shown that treatment with HDI results in a significant improvement in quality-adjusted time (Kilbridge et al. 2002), suggesting that in patients with high-risk melanoma, the clinical benefits of HDI can offset the toxic effects. Kilbridge and colleagues reported strikingly poorer valuations of time with asymptomatic relapse than might generally be presumed and rather better valuations of time with toxicity (Kilbridge et al. 2002). These observations provide useful tools to portray the utility of treatment for individual patients. While several newer agents have demonstrated RFS benefit in the adjuvant setting, it is important to note that, to date, none have been evaluated in patients with clinical stage IIB or IIC cutaneous melanoma, and no results have been reported evaluating these newer agents in direct comparison to HDI.

Y. G. Najjar et al.

Ulcerated Primary Melanoma: A Potentially More IFN-Sensitive Population A number of phase III trials, including EORTC 18952 (comparing intermediate-dose IFN for 1 or 2 years vs. observation) and EORTC 18991 (peg-IFN vs. observation), suggested that the beneficial effect of adjuvant IFN treatment was observed predominantly in patients with ulcerated primary tumors (Eggermont et al. 2005, 2008, 2012a, b, 2016b). This was also observed in the Sunbelt Melanoma Trial (McMasters et al. 2010). A meta-analysis of all adjuvant IFN vs. observation trials performed in the late 1990s and recently published (Ives et al. 2017) examined the differences between the benefit of IFN among patients with non-ulcerated primary tumors and those with ulcerated primaries: at 5 years RFS was 8% better and at 10 years 10% better (Ives et al. 2017). It is difficult to know how to interpret these findings, which emerge from trials without central pathology review of the involved studies, since prospective US trials that have uniformly had central pathology review have not shown a difference in the benefit for patients with ulcerated primary tumors. Of interest, more recent trials of checkpoint blockade with the first-generation agent ipilimumab have also shown a difference in favor of those with ulcerated primary melanomas (Eggermont et al. 2015); more perplexing, the even larger trials that have established the superior activity of secondgeneration checkpoint blockade therapy with nivolumab and with pembrolizumab, discussed below, have not shown any difference in benefit in favor of patients with ulcerated primary tumors (Eggermont et al. 2018a; Weber et al. 2017).

Immune Checkpoint Blockade Anti-CTLA-4 CTLA-4 is a negative regulator of T cell activation and is upregulated on the surface of T cells following immune activation. CTLA-4 on the surface of T cells binds to costimulatory receptors B7-1 and B7-2 on antigen presenting cells with a

Adjuvant Systemic Therapy for High-Risk Melanoma Patients

higher affinity than T cell costimulatory receptor CD28. CTLA-4 outcompetes CD28 for binding to B7-1 and B7-2, leading to a decrease in the costimulatory signal and thereby decreasing T cell activation. Ipilimumab is an anti-CTLA-4 monoclonal antibody developed to block the interaction of CTLA-4 with its ligand and therefore upregulate T cell activation and promote antitumor activity. Two phase III studies have demonstrated an OS benefit with ipilimumab compared to a vaccine or dacarbazine (Hodi et al. 2010; Robert et al. 2011), and in 2011 ipilimumab was the first immune checkpoint inhibitor to be approved by the US FDA for the treatment of unresectable metastatic melanoma, at a dose of 3 mg/kg given every 3 weeks for four total doses.

Ipilimumab EORTC 18071 evaluated ipilimumab in the adjuvant setting in a double-blind phase III RCT

755

comparing adjuvant high-dose ipilimumab (10 mg/ kg) for 3 years to placebo after complete resection of high-risk stage III melanoma (Table 2). 951 patients were randomized in a 1:1 ratio to either ipilimumab (n = 475) or placebo (n = 476) every 3 weeks for four doses, then every 3 months for up to 3 years (7). 3-year RFS was 46.5% (95% CI 41.5–51.3) in the ipilimumab group vs. 34.8% (30.1–39.5) in the placebo group (Eggermont et al. 2015). More recently, an updated analysis with a median follow-up of 5.3 years was published and reported the 5-year RFS was 40.8% for ipilimumab vs. 30.3% for placebo, respectively (HR = 0.76; 95% CI, 0.64–0.89; p < 0.001) (Eggermont et al. 2016a). DMFS at 5 years was 48.3% vs. 38.9% for ipilimumab and placebo, respectively (HR = 0.76; 95.8% CI, 0.64 to 0.92; p = 0.002). A significant survival benefit was identified, with survival 65.4% in the ipilimumab group and 54.4% in the placebo group (HR = 0.72; 95.1% CI, 0.58–0.88; p = 0.001). Grade 3 or 4 adverse events (AEs) occurred in 54.1% of patients in the ipilimumab arm compared to 26.2% of patients in the placebo

Table 2 Key trials of adjuvant checkpoint blockade immunotherapy

Study EORTC 18071 (Eggermont et al. 2015, 2016a)

Enrolled patients (N) 945

Checkmate 238 (Weber et al. 2017, Weber, 2018)

906

KEYNOTE054 (Eggermont et al. 2018a)

1,019

Stage (AJCC seventh edition) IIIA, IIIB, IIIC

Treatment arms Ipilimumab Placebo

Median follow-up 2.74 years 5.3 years

Primary endpoint RFS RFS OS

IIIB, IIIC, IV

Nivolumab Ipilimumab

19.5 months

RFS

IIIA, IIIB, IIIC

Pembrolizumab Placebo

15 months

RFS

Summary of key findings Median RFS 26.1 months with ipilimumab vs. 17.1 months with placebo (HR 0.75; p = 0.0013) 5-year RFS 40.8% with ipilimumab vs. 30.3% with placebo (HR 0.76; p < 0.001) 5-year OS 65.4% with ipilimumab vs. 54.4% with placebo (HR 0.72; p = 0.001) 12-month RFS 70.5% with nivolumab vs. 60.8% with ipilimumab (HR 0.65; p < 0.001) 24-month RFS 63% with nivolumab vs. 50% with ipilimumab (HR 0.66, p < 0.0001) 12-month RFS 75.4% with pembrolizumab vs. 61.0% with placebo (HR 0.57; p < 0.001)

756

arm. The dominant toxicity category was immunerelated AEs, which occurred in 41.6% vs. 2.7% of patients in the ipilimumab and placebo groups, respectively, consistent with the drug’s mechanism of action. Treatment discontinuation due to AEs occurred in 52% of patients in the ipilimumab arm, and 1.1% of patients in the ipilimumab arm died due to immune-related AE. Based on the results of EORTC 18071, in October 2015, high-dose ipilimumab received regulatory approval from the US FDA for adjuvant therapy in patients with resected stage III melanoma. Given the increased toxicity profile of highdose ipilimumab over standard low-dose ipilimumab approved for treatment of metastatic melanoma, ECOG-ACRIN conducted US intergroup trial E1609 to evaluate the efficacy of high-dose and standard-dose ipilimumab vs. HDI. Patients with resected high-risk melanoma (stages IIIB, IIIC, M1a, M1b) were randomized to ipilimumab 10 mg/kg or 3 mg/kg vs. HDI. The co-primary endpoints in this study were RFS and OS. Grade 3 or greater AEs were reported in 57% of patients treated with high-dose ipilimumab compared to 36.4% of patients treated with standard-dose ipilimumab. Unplanned analysis of RFS among patients randomized to the two ipilimumab arms was performed at median follow-up of 3.1 years and showed no difference in RFS (Tarhini 2017). This analysis identified RFS of 54% vs. 56% with high-dose ipilimumab and standard-dose ipilimumab, respectively. The final results of this trial, specifically reporting the co-primary endpoints of RFS and OS in relation to HDI, are pending.

Anti-PD1 Suppression of immune surveillance has been identified as barrier to immunotherapy of cancer. Programmed death 1 receptor (PD-1), expressed on T cells, and its ligand programmed death ligand 1 (PD-L1) are involved in tumor-mediated immune suppression. Anti-PD1 monoclonal antibodies nivolumab and pembrolizumab were approved in 2015 by the US FDA for the treatment of unresectable metastatic melanoma. The

Y. G. Najjar et al.

demonstration of an improved response rate with anti-PD1 agents compared to ipilimumab in patients with metastatic melanoma led to the development of clinical trials to evaluate antiPD1 agents in the adjuvant setting.

Nivolumab The clinical benefit of adjuvant nivolumab therapy for patients with stage III-IV resectable melanoma was evaluated in CheckMate 238 (Weber et al. 2017) (Table 2). This double-blind phase III RCT compared nivolumab vs. high-dose ipilimumab following complete resection of stage IIIB, IIIC, or stage IV melanoma. The primary endpoint was RFS, and secondary endpoints included OS, safety, and side effect profile. DMFS was an exploratory endpoint. All patients were required to have LND or resection of metastatic disease within 12 weeks prior to randomization. Notably, patients with resected brain metastases were eligible for this adjuvant trial. Patients were stratified by disease stage (IIIB or IIIC or stage IV M1a or M1b) and PD-L1 status (5% cutoff of PD-L1 staining of tumor cells). Nine hundred six patients were randomly assigned to treatment for up to 1 year with either nivolumab 3 mg/kg every 2 weeks or ipilimumab 10 mg/kg every 3 weeks for 4 doses and then every 12 weeks. At a minimum follow-up interval of 18 months, the 12-month RFS was 70.5% in the nivolumab group and 60.8% in the ipilimumab group (HR = 0.65; 97.65% CI, 0.51 to 0.83; p < 0.001). There was a significant improvement in RFS with nivolumab compared to ipilimumab across disease stages, both among those with stage IIIB or IIIC disease (HR, 0.65; 95% CI, 0.51 to 0.82) and those with stage IV disease (HR 0.70; 95% CI, 0.45 to 1.10). Superior RFS was observed with nivolumab compared to ipilimumab both in patients with PD-L1 expression 1% of cells, which was considered to be indicative of PD-L1 positivity. All patients were required to have undergone LND within 13 weeks before the start of treatment. Patients were stratified by disease stage (IIIA, IIIB, IIIC with one to three positive nodes, or IIIC with four or more positive nodes) and geographic region (17 regions). Patients were randomly assigned to treatment with either pembrolizumab 200 mg IV or placebo every 3 weeks for a total of 18 doses (approximately 1 year) or until disease recurrence or unacceptable toxicity. 1019 patients were randomized, with 514 patients assigned to pembrolizumab and 505 patients assigned to placebo. The 12-month rate of RFS in the intention-to-treat population was 75.4% (95% CI, 71.3 to 78.9) in the pembrolizumab arm vs. 61.0% (95% CI, 56.5 to 65.1) in the placebo arm. At 18 months the rate of RFS was 71.4% (95% CI, 66.8 to 75.4) in the pembrolizumab arm vs. 53.2% (95% CI, 47.9 to 58.2) in the placebo arm. RFS was significantly longer in the pembrolizumab arm than in the placebo arm (HR for recurrence or death, 0.57; 98.4% CI, 0.43 to 0.74; p < 0.001). RFS in the subgroup of patients with PD-L1positive tumors was a co-primary endpoint in this study. 853 patients met criteria for PD-L1-positive tumors, and in this subgroup, the 12-month RFS was 77.1% (95% CI, 72.7 to 80.9) in the pembrolizumab arm and 62.6% (95% CI, 57.7 to 67.0) in the placebo arm. RFS was significantly improved with pembrolizumab compared to placebo in both PD-L1-positive tumors (HR for recurrence or death, 0.54; 95% CI, 0.42 to 0.69; p < 0.001) and PD-L1-negative tumors (HR for recurrence or death, 0.47; 95% CI, 0.26 to 0.85; p = 0.01).

758

Immune-related adverse events of any grade occurred in 37.3% of patients in the pembrolizumab arm and 9.0% of patients in the placebo arm. Treatment-related adverse events of grade 3 through 5 occurred in 14.7% of patients in the pembrolizumab arm and 3.4% in the placebo arm. The pembrolizumab arm did have one treatment-death due to myositis. The incidence of grade 3 or 4 immune-related adverse events was 7.1% and 0.6% in the pembrolizumab and placebo arms, respectively. This grade 3 or 4 immune-related AE included colitis (2.0% in pembrolizumab arm), hypophysitis or hypopituitarism (0.6%), and type 1 diabetes mellitus (1.0%). The majority of these resolved, and nearly half (48.8%) resolved within 2 months after the last dose of pembrolizumab. A double-blind, placebo-controlled phase 3 study of adjuvant pembrolizumab vs. placebo in resected stage IIB and IIC melanoma is currently enrolling (Keynote-716).

Targeted Therapy The mitogen-activated protein kinase (MAPK) pathway has been identified as a therapeutic target in the treatment of advanced melanoma with BRAF V600 E/K driver mutations (see also Targeted Therapies for BRAF Mutant Metastatic Melanoma). Activating mutations in BRAF are found in approximately 40% of advanced melanomas (Long et al. 2011). Several BRAF inhibitors, including vemurafenib (Chapman et al. 2011), dabrafenib (Hauschild et al. 2012), and encorafenib (Dummer et al. 2018), have been evaluated in the metastatic setting. Resistance inevitably develops to BRAF inhibition via reactivation of the MAPK pathway, and therefore the combination of BRAF and MEK inhibition was evaluated and demonstrated increased efficacy in the metastatic setting with improvements in response rate, PFS, and OS (Long et al. 2017a; Rizos et al. 2014). Combined BRAF/MEK inhibition is now standard in the treatment of BRAF mutation-positive metastatic melanoma, and the successes of this treatment strategy in metastatic disease led to its evaluation in the adjuvant setting.

Y. G. Najjar et al.

The clinical efficacy of single-agent vemurafenib in the adjuvant setting was evaluated in the BRIM 8 study, which was a phase 3, doubleblind, placebo-controlled RCT. The study enrolled 498 patients with BRAF V600 mutation-positive melanoma in two separate cohorts of stage IIC, IIIA, or IIIB (cohort 1) or stage IIIC (cohort 2) by AJCC seventh edition (Maio et al. 2018). Patients were randomized to treatment with vemurafenib or placebo for 1 year. The primary endpoint of DFS was evaluated separately in each cohort. Analysis of cohort 2 prior to cohort 1 was prespecified in the protocol, which has important implications for the statistical interpretation of the results. In cohort 2, median DFS was 23.1 months in the treatment group vs. 15.4 months in placebo group (HR, 0.8; 95% CI, 0.54 to 1.18; p = 0.26). Although median DFS was not reached in the vemurafenib group as compared to 36.9 months in the placebo group (HR, 0.54; 95% CI, 0.37 to 0.78; p = 0.001) in cohort 1, this analysis was considered exploratory as the primary endpoint was not met in cohort 2 (Maio et al. 2018). COMBI-AD was a double-blind, placebo-controlled, multicenter phase III trial in which 870 patients with completely resected stage III melanoma with BRAF V600E/K mutations were randomized to oral dabrafenib plus trametinib or placebo for 12 months (Long et al. 2017b) (Table 3). Dabrafenib was administered at a dose of 150 mg twice daily, and trametinib was administered at 2 mg once daily, the same doses that are used in metastatic disease. Patients with stage IIIA (with SLN tumor burden 1 mm), IIIB, or IIIC cutaneous melanoma were eligible, and patients were stratified by stage. The primary endpoint was RFS. Secondary endpoints were OS, DMFS, and safety. At a median follow-up of 2.8 years, the RFS at 3 years was 58% vs. 39% in the dabrafenib plus trametinib arm and placebo arm, respectively (HR 0.47; 95% CI, 0.39 to 0.58, p < 0.001). In the treatment arm vs. placebo arm, respectively, locoregional recurrence occurred in 12% vs. 25%, both local and distant recurrence in 2% vs. 2%, and distant recurrence in 22% vs. 29%. In the first interim analysis of OS, the 3-year estimated OS was 86% compared to 77% in the treatment group vs. placebo group, respectively (HR for death, 0.57;

Adjuvant Systemic Therapy for High-Risk Melanoma Patients

759

Table 3 Key trials of adjuvant targeted therapy

Study COMBI-AD (Long et al. 2017b; Hauschild et al. 2018)

Enrolled patients (N) 870

Stage (AJCC seventh edition) IIIA, IIIB, IIIC

Treatment arms Dabrafenib/ trametinib Placebo

95% CI, 0.42 to 0.79; p = 0.0006). However, the between-group difference did not reach the prespecified threshold of p = 0.000019 required to meet statistical significance in the interim analysis of OS. Toxicities were similar in the adjuvant setting compared to prior trials evaluating this combination in the metastatic setting. Treatment-related grade 3 or 4 adverse events occurred in 41% of patients in the treatment arm and 14% in the placebo arm, and 26% of patients in the treatment arm discontinued therapy due to an adverse event vs. 3% in the placebo arm. The most frequent grade 3 or 4 adverse events were fever, fatigue, hypertension, and transaminase elevations. Results after longer-term follow-up of the COMBI-AD trial were recently reported (Hauschild et al. 2018). At a median follow-up of 3.7 years, median RFS was not reached in the dabrafenib-trametinib arm compared to 16.6 months in the placebo arm. Four-year RFS rates were 54% and 38%, respectively. The apparent durability of the RFS impact of adjuvant dabrafenib-trametinib at year 4 (3 years after the end of adjuvant treatment) is noteworthy, given the high frequency with which resistance develops when metastatic disease is treated with BRAF-targeted therapy. Based on the results of COMBI-AD, in April 2018, the US FDA granted approval to dabrafenib and trametinib in combination for the adjuvant treatment of patients with melanoma with BRAF V600E or V600K mutations as detected by an FDA-approved test and involvement of lymph nodes following complete resection. Following

Median followup 2.8 years

Primary endpoint RFS

Summary of key findings 3-year RFS 58% with combination therapy vs. 39% with placebo (HR 0.47; p < 0.001) 4-year RFS 54% with combination therapy vs. 38% with placebo (HR 0.49; 95% CI, 0.40 to 0.59)

the publication of the results of COMBI-AD and CheckMate 238, which demonstrated clinically significant benefit in the adjuvant setting for the combination of BRAF/MEK inhibition and the single agent application of anti-PD1 therapy for 1 year, respectively, single-agent BRAF inhibitors are not recommended in the adjuvant setting.

Choice of Adjuvant Therapy in the BRAF V600E/K MutationPositive Patient The question of what the ideal adjuvant treatment for patients with resected stage III BRAF V600E/K mutant melanoma remains unanswered. Patients with BRAF mutations were included in the trials of nivolumab and pembrolizumab, and anti-PD1 agents demonstrate efficacy regardless of BRAF mutation status. In metastatic unresectable disease, the relative role of targeted vs. checkpoint blockade therapy (dabrafenib plus trametinib vs. nivolumab plus ipilimumab) for first-line treatment of BRAF V600E/K metastatic disease is being evaluated in the ongoing intergroup US trial EA6134. We recommend that at this time, the decision of whether to utilize immunotherapy or targeted therapy for adjuvant treatment must be based on the individual patient, considering the potential long-term benefits and toxicity to be anticipated. The toxicity profile of immune checkpoint blockade vs. targeted therapy is markedly different, with the toxicity of dabrafenib plus trametinib generally resolving rapidly following cessation of drug, whereas several immune-mediated toxicities of CBI can have

760

longer-term sequelae. Moreover, it is reasonable to select patients for treatment in the adjuvant setting as one would in the metastatic setting. For example, patients with a history of autoimmune disease requiring active treatment or patients who require >10 mg prednisone daily should not be treated with immunotherapy in the adjuvant setting and should rather be treated with targeted therapy. On the other hand, a patient without a history of autoimmune disease but with a decreased cardiac ejection fraction may be considered for immunotherapy rather than targeted therapy, given the higher risk of cardiotoxicity with targeted agents.

Contraindications to Immunotherapy Patients with autoimmune disease other than vitiligo, thyroid disease, or type 1 diabetes have generally been excluded from immunotherapy trials due to the association of these drugs with immune-mediated toxicity (see also ▶ Managing Checkpoint Inhibitor Symptoms and Toxicity for Metastatic Melanoma). As detailed above, it is reasonable to select patients for treatment in the adjuvant setting as one would in the metastatic setting, and patients with autoimmune disease and BRAF V600E/K mutation should be strongly considered for treatment with targeted rather than immunotherapy. There is, however, limited data on the safety of immunotherapeutic agents in patients with autoimmune disease. In a retrospective study of 52 patients with advanced melanoma and pre-existing autoimmune disease who were treated with anti-PD1, 4% of patients required treatment discontinuation, and 38% of patients had a flare of autoimmune disease requiring immune suppression (Menzies et al. 2017). Twenty-nine percent of patients in this study developed immune-related toxicity that was different than their underlying disease, and 8% discontinued treatment due these other immunerelated events. If a patient has a history of autoimmune disease, a thorough discussion of the risks versus benefits of initiation of immunotherapy is necessary, and multidisciplinary care with the physician treating the autoimmune disease is recommended.

Y. G. Najjar et al.

Conclusions Adjuvant treatment with anti-PD1 monoclonal antibodies, or the combination of BRAF/MEK inhibition in patients with BRAF V600E/K mutations, represents the new standard of care in the management of patients with stage III or higher resected melanoma. However, several questions require urgent further study. The planned duration of adjuvant treatment with anti-PD1 or dabrafenib/ trametinib has been arbitrarily set at 1 year. It is unknown whether 1 year is the optimal duration of adjuvant therapy. Indeed, optimal duration of therapy has not yet been established in the treatment of the unresectable metastatic disease setting, and randomized discontinuation trials in both settings may be of utility in exposing patients to shorter durations of therapy and potentially less medication and financial toxicity, if equally efficacious. Furthermore, adjuvant therapy is administered to patients who are without clinical evidence of disease. Indeed, while stage IIB-stage IIID disease is associated with increasing level of recurrence risk, there is a proportion of patients who are cured with surgery alone. The identification of predictive biomarkers is essential to refining the population offered adjuvant treatment. Further improvement may be achieved from neoadjuvant schedules, which are currently a very actively explored field and discussed in Neoadjuvant Therapy. Neoadjuvant therapy has multiple potential benefits, as it may more rapidly return response data and information regarding the mechanism of new combinations; in addition, it may facilitate surgery, reduce the need for postoperative radiotherapy, and increase locoregional control. For example, neoadjuvant BRAF/MEK yielded an objective response rate of 100% in 13 patients and reduced relapse rates compared to surgery and standard postoperative adjuvant approaches (which did not include adjuvant BRAF/MEK inhibition or anti-PD1 immunotherapy) (Amaria et al. 2018). The combination of nivolumab (1 mg/kg) and ipilimumab (3 mg/kg) is active in the neoadjuvant setting as well and, interestingly, induces a greater number and variety of TCR clones (Rozeman et al. 2017). Grade 3–4 toxicities are observed in 80–90% of patients

Adjuvant Systemic Therapy for High-Risk Melanoma Patients

and indicate that low-dose ipilimumab combined with anti-PD1 agents may have a future in the neoadjuvant or adjuvant settings. Expansion of the indications for checkpoint blockade and targeted therapy to the adjuvant setting provides patients with additional options to reduce risk of melanoma occurrence. However, treatment early in the disease course in patients who then may develop recurrence and where the therapy utilized may have unknown effect upon their likelihood of response to later application of these same new therapies. Importantly, the role of re-treatment in the metastatic setting for patients who received adjuvant therapy is yet to be fully explored. Indeed, the fact that we are using these agents in the adjuvant setting makes the development of novel targets for checkpoint inhibition even more pressing. The development of novel combinations and novel immune checkpoint inhibitors will provide further treatment options to these patients who recur on the currently approved agents for adjuvant treatment.

References Amaria RN, Prieto PA, Tetzlaff MT, Reuben A, Andrews MC, Ross MI, Glitza IC, Cormier J, Hwu WJ, Tawbi HA, Patel SP, Lee JE, Gershenwald JE, Spencer CN, Gopalakrishnan V, Bassett R, Simpson L, Mouton R, Hudgens CW, Zhao L, Zhu H, Cooper ZA, Wani K, Lazar A, Hwu P, Diab A, Wong MK, Mcquade JL, Royal R, Lucci A, Burton EM, Reddy S, Sharma P, Allison J, Futreal PA, Woodman SE, Davies MA, Wargo JA (2018) Neoadjuvant plus adjuvant dabrafenib and trametinib versus standard of care in patients with high-risk, surgically resectable melanoma: a single-centre, open-label, randomised, phase 2 trial. Lancet Oncol 19:181–193 Balch CM, Gershenwald JE, Soong SJ, Thompson JF, Ding S, Byrd DR, Cascinelli N, Cochran AJ, Coit DG, Eggermont AM, Johnson T, Kirkwood JM, Leong SP, McMasters KM, Mihm MC Jr, Morton DL, Ross MI, Sondak VK (2010) Multivariate analysis of prognostic factors among 2,313 patients with stage III melanoma: comparison of nodal micrometastases versus macrometastases. J Clin Oncol 28:2452–2459 Balch CM, Smalley RV, Bartolucci AA, Burns D, Presant CA, Durant JR (1982) A randomized prospective clinical trial of adjuvant C. parvum immunotherapy in 260 patients with clinically localized melanoma (stage I): prognostic factors analysis and preliminary results of immunotherapy. Cancer 49:1079–1084

761

Bystryn JC, Zeleniuch-Jacquotte A, Oratz R, Shapiro RL, Harris MN, Roses DF (2001) Double-blind trial of a polyvalent, shed-antigen, melanoma vaccine. Clin Cancer Res 7:1882–1887 Cameron DA, Cornbleet MC, Mackie RM, Hunter JA, Gore M, Hancock B, Smyth JF, Scottish Melanoma G (2001) Adjuvant interferon alpha 2b in high risk melanoma – the Scottish study. Br J Cancer 84:1146–1149 Cascinelli N, Belli F, Mackie RM, Santinami M, Bufalino R, Morabito A (2001) Effect of long-term adjuvant therapy with interferon alpha-2a in patients with regional node metastases from cutaneous melanoma: a randomised trial. Lancet 358:866–869 Chapman PB, Hauschild A, Robert C, Haanen JB, Ascierto P, Larkin J, Dummer R, Garbe C, Testori A, Maio M, Hogg D, Lorigan P, Lebbe C, Jouary T, Schadendorf D, Ribas A, O’day SJ, Sosman JA, Kirkwood JM, Eggermont AM, Dreno B, Nolop K, Li J, Nelson B, Hou J, Lee RJ, Flaherty KT, Mcarthur GA, BRIM-3 Study Group (2011) Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N Engl J Med 364:2507–2516 Coates A, Rallings M, Hersey P, Swanson C (1986) PhaseII study of recombinant alpha 2-interferon in advanced malignant melanoma. J Interf Res 6:1–4 Creagan ET, Ahmann DL, Green SJ, Long HJ, Frytak S, Itri LM (1985) Phase II study of recombinant leukocyte A interferon (IFN-rA) plus cimetidine in disseminated malignant melanoma. J Clin Oncol 3:977–981 Creagan ET, Ahmann DL, Green SJ, Long HJ, Frytak S, O’fallon JR, Itri LM (1984a) Phase II study of low-dose recombinant leukocyte a interferon in disseminated malignant melanoma. J Clin Oncol 2:1002–1005 Creagan ET, Ahmann DL, Green SJ, Long HJ, Rubin J, Schutt AJ, Dziewanowski ZE (1984b) Phase II study of recombinant leukocyte A interferon (rIFN-alpha A) in disseminated malignant melanoma. Cancer 54:2844–2849 Creagan ET, Dalton RJ, Ahmann DL, Jung SH, Morton RF, Langdon RM Jr, Kugler J, Rodrigue LJ (1995) Randomized, surgical adjuvant clinical trial of recombinant interferon alfa-2a in selected patients with malignant melanoma. J Clin Oncol 13:2776–2783 Czarnetzki BM, Macher E, Suciu S, Thomas D, Steerenberg PA, Rumke P (1993) Long-term adjuvant immunotherapy in stage I high risk malignant melanoma, comparing two BCG preparations versus non-treatment in a randomised multicentre study (EORTC protocol 18781). Eur J Cancer 29A:1237–1242 De Veer MJ, Holko M, Frevel M, Walker E, Der S, Paranjape JM, Silverman RH, Williams BR (2001) Functional classification of interferon-stimulated genes identified using microarrays. J Leukoc Biol 69:912–920 Dorval T, Palangie T, Jouve M, Garcia-Giralt E, Israel L, Falcoff E, Schwab D, Pouillart P (1986) Clinical phase II trial of recombinant DNA interferon (interferon alpha 2b) in patients with metastatic malignant melanoma. Cancer 58:215–218

762 Dreno B, Thompson JF, Smithers BM, Santinami M, Jouary T, Gutzmer R, Levchenko E, Rutkowski P, Grob JJ, Korovin S, Drucis K, Grange F, Machet L, HERSEY P, Krajsova I, Testori A, Conry R, Guillot B, Kruit WHJ, Demidov L, Thompson JA, Bondarenko I, Jaroszek J, Puig S, Cinat G, Hauschild A, Goeman JJ, Van Houwelingen HC, Ulloa-Montoya F, Callegaro A, Dizier B, Spiessens B, Debois M, Brichard VG, Louahed J, Therasse P, Debruyne C, Kirkwood JM (2018) MAGE-A3 immunotherapeutic as adjuvant therapy for patients with resected, MAGE-A3-positive, stage III melanoma (DERMA): a double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol 19:916–929 Dummer R, Ascierto PA, Gogas HJ, Arance A, Mandala M, Liszkay G, Garbe C, Schadendorf D, Krajsova I, Gutzmer R, Chiarion-Sileni V, Dutriaux C, DE Groot JWB, Yamazaki N, Loquai C, Moutouh-De Parseval LA, Pickard MD, Sandor V, Robert C, Flaherty KT (2018) Encorafenib plus binimetinib versus vemurafenib or encorafenib in patients with BRAF-mutant melanoma (COLUMBUS): a multicentre, open-label, randomised phase 3 trial. Lancet Oncol 19:603–615 Eggermont AM, Chiarion-Sileni V, Grob JJ, Dummer R, Wolchok JD, Schmidt H, Hamid O, Robert C, Ascierto PA, Richards JM, Lebbe C, Ferraresi V, Smylie M, Weber JS, Maio M, Bastholt L, Mortier L, Thomas L, Tahir S, Hauschild A, Hassel JC, Hodi FS, Taitt C, DE Pril V, DE Schaetzen G, Suciu S, Testori A (2016a) Prolonged survival in stage III melanoma with Ipilimumab adjuvant therapy. N Engl J Med 375:1845–1855 Eggermont AM, Chiarion-Sileni V, Grob JJ, Dummer R, Wolchok JD, Schmidt H, Hamid O, Robert C, Ascierto PA, Richards JM, Lebbe C, Ferraresi V, Smylie M, Weber JS, Maio M, Konto C, Hoos A, DE Pril V, Gurunath RK, DE Schaetzen G, Suciu S, Testori A (2015) Adjuvant ipilimumab versus placebo after complete resection of high-risk stage III melanoma (EORTC 18071): a randomised, double-blind, phase 3 trial. Lancet Oncol 16:522–530 Eggermont AM, Suciu S, Mackie R, Ruka W, Testori A, Kruit W, Punt CJ, Delauney M, Sales F, Groenewegen G, Ruiter DJ, Jagiello I, Stoitchkov K, Keilholz U, Lienard D, EORTC Melanoma Group (2005) Post-surgery adjuvant therapy with intermediate doses of interferon alfa 2b versus observation in patients with stage IIb/III melanoma (EORTC 18952): randomised controlled trial. Lancet 366:1189–1196 Eggermont AM, Suciu S, Rutkowski P, Kruit WH, Punt CJ, Dummer R, Sales F, Keilholz U, De Schaetzen G, Testori A, EORTC Melanoma Group (2016b) Long term follow up of the EORTC 18952 trial of adjuvant therapy in resected stage IIB-III cutaneous melanoma patients comparing intermediate doses of interferonalpha-2b (IFN) with observation: ulceration of primary is key determinant for IFN-sensitivity. Eur J Cancer 55:111–121

Y. G. Najjar et al. Eggermont AM, Suciu S, Rutkowski P, Marsden J, Santinami M, Corrie P, Aamdal S, Ascierto PA, Patel PM, Kruit WH, Bastholt L, Borgognoni L, Bernengo MG, Davidson N, Polders L, Praet M, Spatz A (2013) Adjuvant ganglioside GM2-KLH/QS-21 vaccination versus observation after resection of primary tumor >1.5 mm in patients with stage II melanoma: results of the EORTC 18961 randomized phase III trial. J Clin Oncol 31:3831–3837 Eggermont AM, Suciu S, Santinami M, Testori A, Kruit WH, Marsden J, Punt CJ, Sales F, Gore M, Mackie R, Kusic Z, Dummer R, Hauschild A, Musat E, Spatz A, Keilholz U, EORTC Melanoma Group (2008) Adjuvant therapy with pegylated interferon alfa-2b versus observation alone in resected stage III melanoma: final results of EORTC 18991, a randomised phase III trial. Lancet 372:117–126 Eggermont AM, Suciu S, Testori A, Kruit WH, Marsden J, Punt CJ, Santinami M, Sales F, Schadendorf D, Patel P, Dummer R, Robert C, Keilholz U, Yver A, Spatz A (2012a) Ulceration and stage are predictive of interferon efficacy in melanoma: results of the phase III adjuvant trials EORTC 18952 and EORTC 18991. Eur J Cancer 48:218–225 Eggermont AM, Suciu S, Testori A, Santinami M, Kruit WH, Marsden J, Punt CJ, Sales F, Dummer R, Robert C, Schadendorf D, Patel PM, DE Schaetzen G, Spatz A, Keilholz U (2012b) Long-term results of the randomized phase III trial EORTC 18991 of adjuvant therapy with pegylated interferon alfa-2b versus observation in resected stage III melanoma. J Clin Oncol 30:3810–3818 Eggermont AMM, Blank CU, Mandala M, Long GV, Atkinson V, Dalle S, Haydon A, Lichinitser M, Khattak A, Carlino MS, Sandhu S, Larkin J, Puig S, Ascierto PA, Rutkowski P, Schadendorf D, Koornstra R, Hernandez-Aya L, Maio M, Van Den Eertwegh AJM, Grob JJ, Gutzmer R, Jamal R, Lorigan P, Ibrahim N, Marreaud S, Van Akkooi ACJ, Suciu S, Robert C (2018a) Adjuvant Pembrolizumab versus placebo in resected stage III melanoma. N Engl J Med 378:1789–1801 Eggermont AMM, Robert C, Ribas A (2018b) The new era of adjuvant therapies for melanoma. Nat Rev Clin Oncol 15:535–536 Ernstoff MS, Reiss M, Davis CA et al (1983) Intravenous (IV) recombinant alpha-2 interferon in metastatic melanoma. Proc Am Soc Clin Oncol 2:57 Faries MB, Mozzillo N, Kashani-Sabet M, Thompson JF, Kelley MC, Deconti RC, Lee JE, Huth JF, Wagner J, Dalgleish A, Pertschuk D, Nardo C, Stern S, Elashoff R, Gammon G, Morton DL, MMAIT-IV Clinical Trial Group (2017) Long-term survival after complete surgical resection and adjuvant immunotherapy for distant melanoma metastases. Ann Surg Oncol 24:3991–4000 Fisher RI, Terry WD, Hodes RJ, Rosenberg SA, Makuch R, Gordon HG, Fisher SG (1981) Adjuvant immunotherapy or chemotherapy for malignant

Adjuvant Systemic Therapy for High-Risk Melanoma Patients melanoma. Preliminary report of the National Cancer Institute randomized clinical trial. Surg Clin North Am 61:1267–1277 Flaherty LE, Othus M, Atkins MB, Tuthill RJ, Thompson JA, Vetto JT, Haluska FG, Pappo AS, Sosman JA, Redman BG, Moon J, Ribas A, Kirkwood JM, Sondak VK (2014) Southwest oncology group S0008: a phase III trial of high-dose interferon alfa-2b versus cisplatin, vinblastine, and dacarbazine, plus interleukin-2 and interferon in patients with high-risk melanoma – an intergroup study of cancer and leukemia group B, Children’s oncology group, eastern cooperative oncology group, and southwest oncology group. J Clin Oncol 32:3771–3778 Gasser O, Sharples KJ, Barrow C, Williams GM, Bauer E, Wood CE, Mester B, Dzhelali M, Caygill G, Jones J, Hayman CM, Hinder VA, Macapagal J, Mccusker M, Weinkove R, Painter GF, Brimble MA, Findlay MP, Dunbar PR, Hermans IF (2017) A phase I vaccination study with dendritic cells loaded with NY-ESO-1 and alpha-galactosylceramide: induction of polyfunctional T cells in high-risk melanoma patients. Cancer Immunol Immunother 67:285–298 Gershenwald JE, Scolyer RA, Hess KR, Sondak VK, Long GV, Ross MI, Lazar AJ, Faries MB, Kirkwood JM, McArthur GA, Haydu LE, Eggermont AMM, Flaherty KT, BALCH CM, Thompson JF, For members of the AmericanJoint Committee on Cancer Melanoma Expert Panel, The International Melanoma Database, Discovery Platform (2017) Melanoma staging: evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67:472–492 Grander D, Sangfelt O, Erickson S (1997) How does interferon exert its cell growth inhibitory effect? Eur J Haematol 59:129–135 Grob JJ, Dreno B, DE LA Salmoniere P, Delaunay M, Cupissol D, Guillot B, Souteyrand P, Sassolas B, Cesarini JP, Lionnet S, Lok C, Chastang C, Bonerandi JJ (1998) Randomised trial of interferon alpha-2a as adjuvant therapy in resected primary melanoma thicker than 1.5 mm without clinically detectable node metastases. French cooperative group on melanoma. Lancet 351:1905–1910 Hancock BW, Wheatley K, Harris S, Ives N, Harrison G, Horsman JM, Middleton MR, Thatcher N, Lorigan PC, Marsden JR, Burrows L, Gore M (2004) Adjuvant interferon in high-risk melanoma: the AIM HIGH study – United Kingdom coordinating Committee on Cancer research randomized study of adjuvant low-dose extended-duration interferon alfa-2a in high-risk resected malignant melanoma. J Clin Oncol 22:53–61 Hansson J, Aamdal S, Bastholt L, Brandberg Y, Hernberg M, Nilsson B, Stierner U, Von Der Maase H, Nordic Melanoma Cooperative, Group (2011) Two different durations of adjuvant therapy with intermediate-dose interferon alfa-2b in patients with high-risk melanoma (Nordic IFN trial): a randomised phase 3 trial. Lancet Oncol 12:144–152

763

Hauschild A, Dummer R, Schadendorf D, Santinami M, Atkinson V, Mandalá M, Chiarion-Sileni V, Larkin J, Nyakas M, Dutriaux C, Haydon A, Robert C, Mortier L, Schachter J, Lesimple T, Plummer R, Dasgupta K, Haas T, Shilkrut M, Gasal E, Kefford R, Kirkwood JM, Long GV (2018). Longer follow-up confirms relapse-free survival benefit with adjuvant Dabrafenib plus Trametinib in patients with resected BRAFV600–mutant stage III melanoma. J Clin Oncol 36(35):3441–3449 Hauschild A, Grob JJ, Demidov LV, Jouary T, Gutzmer R, Millward M, Rutkowski P, Blank CU, Miller WH Jr, Kaempgen E, Martin-Algarra S, Karaszewska B, Mauch C, Chiarion-Sileni V, Martin AM, Swann S, Haney P, Mirakhur B, Guckert ME, Goodman V, Chapman PB (2012) Dabrafenib in BRAF-mutated metastatic melanoma: a multicentre, open-label, phase 3 randomised controlled trial. Lancet 380:358–365 Hauschild A, Weichenthal M, Rass K, Linse R, Berking C, Bottjer J, Vogt T, Spieth K, Eigentler T, Brockmeyer NH, Stein A, Naher H, Schadendorf D, Mohr P, Kaatz M, Tronnier M, Hein R, Schuler G, Egberts F, Garbe C (2010) Efficacy of low-dose interferon {alpha}2a 18 versus 60 months of treatment in patients with primary melanoma of >= 1.5 mm tumor thickness: results of a randomized phase III DeCOG trial. J Clin Oncol 28:841–846 Hersey P, Coates AS, McCarthy WH, Thompson JF, Sillar RW, McLeod R, Gill PG, Coventry BJ, McMullen A, Dillon H, Simes RJ (2002) Adjuvant immunotherapy of patients with high-risk melanoma using vaccinia viral lysates of melanoma: results of a randomized trial. J Clin Oncol 20:4181–4190 Hodi FS, O’day SJ, McDermott DF, Weber RW, Sosman JA, Haanen JB, Gonzalez R, Robert C, Schadendorf D, Hassel JC, Akerley W, Van Den Eertwegh AJ, Lutzky J, Lorigan P, Vaubel JM, Linette GP, Hogg D, Ottensmeier CH, Lebbe C, Peschel C, Quirt I, Clark JI, Wolchok JD, Weber JS, Tian J, Yellin MJ, Nichol GM, Hoos A, Urba WJ (2010) Improved survival with ipilimumab in patients with metastatic melanoma. N Engl J Med 363:711–723 Ives NJ, Suciu S, Eggermont AMM, Kirkwood J, Lorigan P, Markovic SN, Garbe C, Wheatley K, International Melanoma Meta-Analysis Collaborative Group (IMMCG) (2017) Adjuvant interferon-alpha for the treatment of high-risk melanoma: an individual patient data meta-analysis. Eur J Cancer 82:171–183 Kilbridge KL, Cole BF, Kirkwood JM, Haluska FG, Atkins MA, Ruckdeschel JC, Sock DE, Nease RF Jr, Weeks JC (2002) Quality-of-life-adjusted survival analysis of high-dose adjuvant interferon alpha-2b for high-risk melanoma patients using intergroup clinical trial data. J Clin Oncol 20:1311–1318 Kirkwood JM, Ibrahim J, Lawson DH, Atkins MB, Agarwala SS, Collins K, Mascari R, Morrissey DM, Chapman PB (2001a) High-dose interferon alfa-2b does not diminish antibody response to GM2 vaccination in patients with resected melanoma: results of the

764 multicenter eastern cooperative oncology group phase II trial E2696. J Clin Oncol 19:1430–1436 Kirkwood JM, Ibrahim JG, Sondak VK, Richards J, Flaherty LE, Ernstoff MS, Smith TJ, Rao U, Steele M, Blum RH (2000) High- and low-dose interferon alfa-2b in high-risk melanoma: first analysis of intergroup trial E1690/S9111/C9190. J Clin Oncol 18:2444–2458 Kirkwood JM, Ibrahim JG, Sosman JA, Sondak VK, Agarwala SS, Ernstoff MS, Rao U (2001b) High-dose interferon alfa-2b significantly prolongs relapse-free and overall survival compared with the GM2-KLH/ QS-21 vaccine in patients with resected stage IIB-III melanoma: results of intergroup trial E1694/S9512/ C509801. J Clin Oncol 19:2370–2380 Kirkwood JM, Manola J, Ibrahim J, Sondak V, Ernstoff MS, Rao U, Eastern Cooperative Oncology Group (2004) A pooled analysis of eastern cooperative oncology group and intergroup trials of adjuvant high-dose interferon for melanoma. Clin Cancer Res 10:1670–1677 Kirkwood JM, Strawderman MH, Ernstoff MS, Smith TJ, Borden EC, Blum RH (1996) Interferon alfa-2b adjuvant therapy of high-risk resected cutaneous melanoma: the eastern cooperative oncology group trial EST 1684. J Clin Oncol 14:7–17 Kleeberg UR, Suciu S, Brocker EB, Ruiter DJ, Chartier C, Lienard D, Marsden J, Schadendorf D, Eggermont AM, EORTCMelanoma Group in cooperation with the German Cancer Society (DKG) (2004) Final results of the EORTC 18871/DKG 80-1 randomised phase III trial. rIFN-alpha2b versus rIFN-gamma versus ISCADOR M versus observation after surgery in melanoma patients with either high-risk primary (thickness >3 mm) or regional lymph node metastasis. Eur J Cancer 40:390–402 Koops HS, Vaglini M, Suciu S, Kroon BB, Thompson JF, Gohl J, Eggermont AM, Di Filippo F, Krementz ET, Ruiter D, Lejeune FJ (1998) Prophylactic isolated limb perfusion for localized, high-risk limb melanoma: results of a multicenter randomized phase III trial. European Organization for Research and Treatment of Cancer Malignant Melanoma Cooperative Group Protocol 18832, the World Health Organization Melanoma Program Trial 15, and the North American Perfusion Group Southwest Oncology Group-8593. J Clin Oncol 16:2906–2912 Larkin J (2018) Dabrafenib plus trametinib (D + T) as adjuvant treatment of resected BRAF-mutant stage III melanoma: findings from the COMBI-AD trial analyzed based on AJCC 8 classification. J Clin Oncol 36:9591 Larkin J, Chiarion-Sileni V, Gonzalez R, Grob JJ, Cowey CL, Lao CD, Schadendorf D, Dummer R, Smylie M, Rutkowski P, Ferrucci PF, Hill A, Wagstaff J, Carlino MS, Haanen JB, Maio M, Marquez-Rodas I, McArthur GA, Ascierto PA, Long GV, Callahan MK, Postow MA, Grossmann K, Sznol M, Dreno B, Bastholt L, Yang A, Rollin LM, Horak C, Hodi FS, Wolchok JD

Y. G. Najjar et al. (2015) Combined Nivolumab and Ipilimumab or monotherapy in untreated melanoma. N Engl J Med 373:23–34 Legha SS, Papadopoulos NE, Plager C, Ring S, Chawla SP, Evans LM, Benjamin RS (1987) Clinical evaluation of recombinant interferon alfa-2a (Roferon-A) in metastatic melanoma using two different schedules. J Clin Oncol 5:1240–1246 Lipton A, Harvey HA, Balch CM, Antle CE, Heckard R, Bartolucci AA (1991) Corynebacterium parvum versus bacille Calmette-Guerin adjuvant immunotherapy of stage II malignant melanoma. J Clin Oncol 9:1151–1156 Livingston PO, Wong GY, Adluri S, Tao Y, Padavan M, Parente R, Hanlon C, Calves MJ, Helling F, Ritter G et al (1994) Improved survival in stage III melanoma patients with GM2 antibodies: a randomized trial of adjuvant vaccination with GM2 ganglioside. J Clin Oncol 12:1036–1044 Long GV, Flaherty KT, Stroyakovskiy D, Gogas H, Levchenko E, De Braud F, Larkin J, Garbe C, Jouary T, Hauschild A, Chiarion-Sileni V, Lebbe C, Mandala M, Millward M, Arance A, Bondarenko I, Haanen J, Hansson J, Utikal J, Ferraresi V, Mohr P, Probachai V, Schadendorf D, Nathan P, Robert C, Ribas A, Davies MA, Lane SR, Legos JJ, Mookerjee B, Grob JJ (2017a) Dabrafenib plus trametinib versus dabrafenib monotherapy in patients with metastatic BRAF V600E/K-mutant melanoma: long-term survival and safety analysis of a phase 3 study. Ann Oncol 28:1631–1639 Long GV, Hauschild A, Santinami M, Atkinson V, Mandala M, Chiarion-Sileni V, Larkin J, Nyakas M, Dutriaux C, Haydon A, Robert C, Mortier L, Schachter J, Schadendorf D, Lesimple T, Plummer R, Ji R, Zhang P, Mookerjee B, Legos J, Kefford R, Dummer R, Kirkwood JM (2017b) Adjuvant Dabrafenib plus Trametinib in stage III BRAF-mutated melanoma. N Engl J Med 377:1813–1823 Long GV, Menzies AM, Nagrial AM, Haydu LE, Hamilton AL, Mann GJ, Hughes TM, Thompson JF, Scolyer RA, Kefford RF (2011) Prognostic and clinicopathologic associations of oncogenic BRAF in metastatic melanoma. J Clin Oncol 29:1239–1246 Maio M, Lewis K, Demidov L, Mandala M, Bondarenko I, Ascierto PA, Herbert C, Mackiewicz A, Rutkowski P, Guminski A, Goodman GR, Simmons B, Ye C, Yan Y, Schadendorf D, brim8 Investigators (2018) Adjuvant vemurafenib in resected, BRAF(V600) mutationpositive melanoma (BRIM8): a randomised, doubleblind, placebo-controlled, multicentre, phase 3 trial. Lancet Oncol 19:510–520 Markovic S, Suman VJ, Dalton RJ, Woods JE, Fitzgibbons RJ Jr, Wold LE, Buckner JC, Kugler JW, Mailliard JA, Rowland KM, Krook JE, Brown DW, Tirona MT, Creagan ET (2002) Randomized, placebo-controlled, phase III surgical adjuvant clinical trial of megestrol acetate (Megace) in selected patients with malignant melanoma. Am J Clin Oncol 25:552–556

Adjuvant Systemic Therapy for High-Risk Melanoma Patients McCarty MF, Bielenberg D, Donawho C, Bucana CD, Fidler IJ (2002) Evidence for the causal role of endogenous interferon-alpha/beta in the regulation of angiogenesis, tumorigenicity, and metastasis of cutaneous neoplasms. Clin Exp Metastasis 19:609–615 McMasters KM, Edwards MJ, Ross MI, Reintgen DS, Martin RC 2nd, Urist MM, Noyes RD, Sussman JJ, Stromberg AJ, Scoggins CR (2010) Ulceration as a predictive marker for response to adjuvant interferon therapy in melanoma. Ann Surg 252:460–465. Discussion 465–6 Mcmasters KM, Egger ME, Edwards MJ, Ross MI, Reintgen DS, Noyes RD, Martin RC 2nd, Goydos JS, Beitsch PD, Urist MM, Ariyan S, Sussman JJ, Davidson BS, Gershenwald JE, Hagendoorn LJ, Stromberg AJ, Scoggins CR (2016) Final results of the Sunbelt Melanoma trial: a multi-institutional prospective randomized Phase III study evaluating the role of adjuvant high-dose interferon Alfa-2b and completion Lymph node dissection for patients staged by sentinel Lymph node biopsy. J Clin Oncol 34:1079–1086 Meisenberg BR, Ross M, Vredenburgh JJ, Jones R, Shpall EJ, Seigler HF, Coniglio DM, Wu K, Peters WP (1993) Randomized trial of high-dose chemotherapy with autologous bone marrow support as adjuvant therapy for high-risk, multi-node-positive malignant melanoma. J Natl Cancer Inst 85:1080–1085 Menzies AM, Johnson DB, Ramanujam S, Atkinson VG, Wong ANM, Park JJ, Mcquade JL, Shoushtari AN, Tsai KK, Eroglu Z, Klein O, Hassel JC, Sosman JA, Guminski A, Sullivan RJ, Ribas A, Carlino MS, Davies MA, Sandhu SK, Long GV (2017) Anti-PD-1 therapy in patients with advanced melanoma and preexisting autoimmune disorders or major toxicity with ipilimumab. Ann Oncol 28:368–376 Meyskens FL Jr, Liu PY, Tuthill RJ, Sondak VK, Fletcher WS, Jewell WR, Samlowski W, Balcerzak SP, Rector DJ, Noyes RD et al (1994) Randomized trial of vitamin A versus observation as adjuvant therapy in high-risk primary malignant melanoma: a southwest oncology group study. J Clin Oncol 12:2060–2065 Miller LL, Spitler LE, Allen RE, Minor DR (1988) A randomized, double-blind, placebo-controlled trial of transfer factor as adjuvant therapy for malignant melanoma. Cancer 61:1543–1549 Neefe JR, Legha SS, Markowitz A, Salmon S, Meyskens F, Groopman J, Campion M, Evans L (1990) Phase II study of recombinant alpha-interferon in malignant melanoma. Am J Clin Oncol 13:472–476 Paterson AH, Willans DJ, Jerry LM, Hanson J, McPherson TA (1984) Adjuvant BCG immunotherapy for malignant melanoma. Can Med Assoc J 131:744–748 Pectasides D, Dafni U, Bafaloukos D, Skarlos D, Polyzos A, Tsoutsos D, Kalofonos H, Fountzilas G, Panagiotou P, Kokkalis G, Papadopoulos O, Castana O, Papadopoulos S, Stavrinidis E, Vourli G, Ioannovich J, Gogas H (2009) Randomized phase III study of 1 month versus 1 year of adjuvant high-dose

765

interferon alfa-2b in patients with resected high-risk melanoma. J Clin Oncol 27:939–944 Pehamberger H, Soyer HP, Steiner A, Kofler R, Binder M, Mischer P, Pachinger W, Aubock J, Fritsch P, Kerl H, Wolff K (1998) Adjuvant interferon alfa-2a treatment in resected primary stage II cutaneous melanoma. Austrian Malignant Melanoma Cooperative Group. J Clin Oncol 16:1425–1429 Quirt IC, Shelley WE, Pater JL, Bodurtha AJ, McCulloch PB, McPherson TA, Paterson AH, Prentice R, Silver HK, Willan AR et al (1991) Improved survival in patients with poor-prognosis malignant melanoma treated with adjuvant levamisole: a phase III study by the National Cancer Institute of Canada clinical trials group. J Clin Oncol 9:729–735 Rizos H, Menzies AM, Pupo GM, Carlino MS, Fung C, Hyman J, Haydu LE, Mijatov B, Becker TM, Boyd SC, Howle J, Saw R, Thompson JF, Kefford RF, Scolyer RA, Long GV (2014) BRAF inhibitor resistance mechanisms in metastatic melanoma: spectrum and clinical impact. Clin Cancer Res 20:1965–1977 Robert C, Long GV, Brady B, Dutriaux C, Maio M, Mortier L, Hassel JC, Rutkowski P, McNeil C, Kalinka-Warzocha E, Savage KJ, Hernberg MM, Lebbe C, Charles J, Mihalcioiu C, Chiarion-Sileni V, Mauch C, Cognetti F, Arance A, Schmidt H, Schadendorf D, Gogas H, Lundgren-Eriksson L, Horak C, Sharkey B, Waxman IM, Atkinson V, Ascierto PA (2015a) Nivolumab in previously untreated melanoma without BRAF mutation. N Engl J Med 372:320–330 Robert C, Schachter J, Long GV, Arance A, Grob JJ, Mortier L, Daud A, Carlino MS, McNeil C, Lotem M, Larkin J, Lorigan P, Neyns B, Blank CU, Hamid O, Mateus C, Shapira-Frommer R, Kosh M, Zhou H, Ibrahim N, Ebbinghaus S, Ribas A, KEYNOTE-006 investigators (2015b) Pembrolizumab versus Ipilimumab in Advanced Melanoma. N Engl J Med 372:2521–2532 Robert C, Thomas L, Bondarenko I, O'day S, Weber J, Garbe C, Lebbe C, Baurain JF, Testori A, Grob JJ, Davidson N, Richards J, Maio M, Hauschild A, Miller WH Jr, Gascon P, Lotem M, Harmankaya K, Ibrahim R, Francis S, Chen TT, Humphrey R, Hoos A, Wolchok JD (2011) Ipilimumab plus dacarbazine for previously untreated metastatic melanoma. N Engl J Med 364:2517–2526 Rozeman EA, Fanchi L, Van Akkooi ACJ et al (2017) (Neo-)adjuvant ipilimumab 1 nivolumab (IPI1NIVO) in palpable stage 3 melanoma – updated relapse free survival (RFS) data from the OpACIN trial and first biomarker analyses. Ann Oncol 28(Suppl 5):1221 Siegel RL, Miller KD, Jemal A (2018) Cancer statistics, 2018. CA Cancer J Clin 68:7–30 Sondak VK, Liu PY, Tuthill RJ, Kempf RA, Unger JM, Sosman JA, Thompson JA, Weiss GR, Redman BG, Jakowatz JG, Noyes RD, Flaherty LE (2002) Adjuvant immunotherapy of resected, intermediate-thickness, node-negative melanoma with an allogeneic tumor

766 vaccine: overall results of a randomized trial of the southwest oncology group. J Clin Oncol 20:2058–2066 Spitler LE (1991) A randomized trial of levamisole versus placebo as adjuvant therapy in malignant melanoma. J Clin Oncol 9:736–740 Tarhini A (2017) A phase III randomized study of adjuvant ipilimumab (3 or 10 mg/kg) versus high-dose interferon alfa-2b for resected high-risk melanoma (U.S. Intergroup E1609): preliminary safety and efficacy of the ipilimumab arms. ASCO 35:9500 Thyrell L, Erickson S, Zhivotovsky B, Pokrovskaja K, Sangfelt O, Castro J, Einhorn S, Grander D (2002) Mechanisms of interferon-alpha induced apoptosis in malignant cells. Oncogene 21:1251–1262 Veronesi U, Adamus J, Aubert C, Bajetta E, Beretta G, Bonadonna G, Bufalino R, Cascinelli N, Cocconi G, Durand J, De Marsillac J, Ikonopisov RL, Kiss B, Lejeune F, Mackie R, Madej G, Mulder H, Mechl Z, Milton GW, Morabito A, Peter H, Priario J, Paul E, Rumke P, Sertoli R, Tomin R (1982) A randomized trial of adjuvant chemotherapy and immunotherapy in cutaneous melanoma. N Engl J Med 307:913–916 Wallack MK, Sivanandham M, Balch CM, Urist MM, Bland KI, Murray D, Robinson WA, Flaherty L, Richards JM, Bartolucci AA, Rosen L (1998) Surgical adjuvant active specific immunotherapy for patients with stage III melanoma: the final analysis of data from a phase III, randomized, double-blind, multicenter vaccinia melanoma oncolysate trial. J Am Coll Surg 187:69–77. Discussion 77–9

Y. G. Najjar et al. Wallack MK, Sivanandham M, Balch CM, Urist MM, Bland KI, Murray D, Robinson WA, Flaherty LE, Richards JM, Bartolucci AA et al (1995) A phase III randomized, double-blind multiinstitutional trial of vaccinia melanoma oncolysate-active specific immunotherapy for patients with stage II melanoma. Cancer 75:34–42 Weber J, Mandala M, Del Vecchio M, Gogas HJ, Arance AM, Cowey CL, Dalle S, Schenker M, Chiarion-SileniV, Marquez-Rodas I, Grob JJ, Butler MO, Middleton MR, Maio M, Atkinson V, Queirolo P, Gonzalez R, Kudchadkar RR, Smylie M, Meyer N, Mortier L, Atkins MB, Long GV, Bhatia S, Lebbe C, Rutkowski P, Yokota K, Yamazaki N, Kim TM, De Pril V, Sabater J, Qureshi A, Larkin J, Ascierto PA, Checkmate C (2017) Adjuvant Nivolumab versus Ipilimumab in resected stage III or IV melanoma. N Engl J Med 377:1824–1835 Weber JS (2018) Adjuvant therapy with nivolumab versus ipilimumab after complete resection of stage III/IV melanoma: updated results from a phase III trial (CheckMate 238). J Clin Oncol 36:9502. Abstract presented at ASCO 2018 Wheatley K, Ives N, Hancock B, Gore M, Eggermont A, Suciu S (2003) Does adjuvant interferon-alpha for high-risk melanoma provide a worthwhile benefit? A meta-analysis of the randomised trials. Cancer Treat Rev 29:241–252

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients Emily Z. Keung, Rodabe N. Amaria, Vernon K. Sondak, Merrick I. Ross, John M. Kirkwood, and Jennifer A. Wargo

Contents The Current Landscape of Systemic Therapy for Stage III and IV Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 768 Patients with Clinical Stage III Melanoma Are Ideal Candidates for Neoadjuvant Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 769 Rationale for Neoadjuvant Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 770 The History of Neoadjuvant Therapy Use in Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . 772 Neoadjuvant Biochemotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 772 Neoadjuvant High Dose Interferon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 772 The Current State of Neoadjuvant Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 772 Neoadjuvant Targeted Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774

E. Z. Keung · M. I. Ross · J. A. Wargo (*) Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected]; mross@mdanderson. org; [email protected] R. N. Amaria Department of Melanoma Medical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] V. K. Sondak Department of Cutaneous Oncology, H. Lee Moffitt Cancer Center, Tampa, FL, USA e-mail: vernon.sondak@moffitt.org J. M. Kirkwood Departments of Medicine, Dermatology, and Translational Science, University of Pittsburgh and UPMC Hillman Cancer Center, Pittsburgh, PA, USA e-mail: [email protected] © This is a U.S. Government work and not under copyright protection in the U.S.; foreign copyright protection may apply 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_70

767

768

E. Z. Keung et al. Neoadjuvant Immune Checkpoint Inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774 Neoadjuvant Checkpoint Inhibitor Therapy in Combination with Other Therapies . . . 784 Neoadjuvant Local Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785 Neoadjuvant Therapy for Borderline Resectable or Unresectable Melanoma . . . 786 Targeted Therapies for Borderline Resectable/Unresectable Melanoma . . . . . . . . . . . . . . . 786 Immune Checkpoint Blockade for Borderline Resectable/Unresectable Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 786 Predicting Response to Neoadjuvant Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 786 Prognostic Biomarkers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 786 Predictive Biomarkers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 787 Unmet Needs in Neoadjuvant Therapy for Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 788 Therapeutic Goals and Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 789 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 790 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 790

Abstract

The landscape of treatment options and outcomes for patients with locally advanced and metastatic melanoma has changed dramatically in the past decade, with the introduction of molecularly targeted antitumor therapy and immunotherapy. Despite these recent advances for advanced stage melanoma, stage III melanoma patients have heterogeneous outcomes and have highly variable prognosis with respect to risk of loco-regional and distant recurrence and survival. Patients with clinical stage III disease, defined as those with palpable nodes with or without in-transit metastases, represent a high-risk population with poor outcomes even with the recent progress in adjuvant therapy. The full potential of targeted and immunotherapy, as well as other novel therapies, in the neoadjuvant setting is unknown and are only beginning to be investigated. Neoadjuvant therapy conveys advantages over administering the same treatment postoperatively as, in some cases, favorable response to treatment can make surgery less morbid and more effective. Additionally, response to therapy can be assessed and postoperative therapy adjusted accordingly depending on degree of response. Neoadjuvant therapy is associated with disadvantages as well; the exact stage of disease may be unknown when treatment begins, and adverse effects of treatment can negatively impact planned surgery or

increase the risk of postoperative complications. The results of ongoing and future neoadjuvant clinical trials will undoubtedly shape the standard of care for patients with locally advanced melanoma at high risk of recurrence. Ultimately, enhanced selection of patients for systemic therapy prior to surgery could delay or prevent distant metastatic disease and improve survival for melanoma patients.

The Current Landscape of Systemic Therapy for Stage III and IV Melanoma The landscape of treatment options and outcomes for patients with locally advanced and metastatic melanoma has changed dramatically in the past decade with the introduction of molecularly targeted therapies and immunotherapies (see also chapters ▶ “Targeted Therapies for BRAFMutant Metastatic Melanoma,” ▶ “Checkpoint Inhibitors in the Treatment of Metastatic Melanoma,” ▶ “Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma,” ▶ “Managing Checkpoint Inhibitor Symptoms and Toxicity for Metastatic Melanoma,” and ▶ “Sequencing and Combinations of Molecularly Targeted and Immunotherapy for BRAFMutant Melanoma”). Patients with stage IV metastatic melanoma, who previously were considered

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

to have a largely fatal disease without effective systemic therapies, may now benefit from targeted and immune checkpoint blockade which is associated with significant objective response rates (ORR) and have improved overall survival (OS) (Long et al. 2014, 2015; Larkin et al. 2015; Robert et al. 2015a, b, and c). More recently, the use of targeted therapies and immune checkpoint inhibitors in the adjuvant setting following complete resection of stage III melanoma has demonstrated significant improvements in recurrence-free survival (RFS) (Long et al. 2017; Weber et al. 2017; Eggermont et al. 2018). Three key randomized controlled trials (EORTC 18071, COMBI-AD and Keynote054) demonstrated the superiority of high-dose ipilimumab (Eggermont et al. 2015), dabrafenib and trametinib (Long et al. 2017), and pembrolizumab (Eggermont et al. 2018), respectively, to placebo in terms of RFS and – in the case of ipiliumab – OS as well. In Checkmate-238, patients with resected stage IIIB/C and IV melanoma who received adjuvant nivolumab had significantly longer RFS at 1 year compared to those who received adjuvant ipilimumab (70.5% vs. 60.8%, HR 0.65, 97.5%, CI 0.51–0.83,

P < 0.001) with lower rates of grade 3 and 4 adverse events (Weber et al. 2017). Whether these improvements in 1 year RFS translate to longer-term RFS or melanoma-specific survival (MSS) benefits remain to be seen.

Patients with Clinical Stage III Melanoma Are Ideal Candidates for Neoadjuvant Treatment Despite recent success in the development of therapies for the metastatic and adjuvant setting, it is important to recognize that patients with stage III melanoma are heterogeneous and have highly variable prognosis with respect to risk of locoregional and distant recurrence as well as MSS (Fig. 1) (Gershenwald et al. 2017). In the recently reported 8th edition analyses of the American Joint Committee on Cancer (AJCC) melanoma staging system, 5-year MSS range from as high as 93% for patients with stage IIIA melanoma (T1-2aN1aM0) to as low as 32% for those with stage IIID (T4bN3M0) disease. Patients with clinically detected nodal disease without in transit metastases (N1b, N2b, N3b)

1.0

Melanoma-Specific Survival Probability

Fig. 1 Melanoma-specific survival of patients by stage III subgroups. (Reproduced with permission from Gershenwald et al. 2017)

769

0.8

0.6

0.4

N 0.2

0.0 0

2

5-YR 10-YR

lllA

1006

93%

88%

lllB

1170

83%

77%

lllC

2201

69%

60%

lllD

205

32%

24%

4 6 Years since diagnosis

8

10

770

have 5-year MSS rates of 64–76% (Gershenwald et al. 2017) and relapse rates of 58–89% (Balch et al. 2009). Although both adjuvant BRAF/MEK inhibitors (dabrafenib/trametinib) (Long et al. 2017) and immune checkpoint inhibitors [ipilimumab (Eggermont et al. 2015, 2016), pembrolizumab (Eggermont et al. 2018), nivolumab (Weber et al. 2017)] are associated with improved RFS following complete resection of stage III melanoma, patients with clinical stage III melanoma in these studies remained at high risk for recurrence despite postoperative adjuvant immunotherapy. In the patients with macroscopic nodal involvement enrolled on the EORTC 18071 trial of adjuvant ipilimumab versus placebo, the hazard ratio (HR) for progression events was only 0.83 (0.63–1.10) compared to 0.68 (0.47–0.99) for those with microscopic nodal involvement. Thus, although adjuvant therapy has improved patient outcomes and can provide cure in some patients, most high risk patients with clinical stage III disease do not achieve cure with these modalities and novel treatment options and strategies are needed. The patient population for neoadjuvant therapy as well as the duration of neoadjuvant therapy has varied within the cadre of recent neoadjuvant trials. Most neoadjuvant trials to date have enrolled completely resectable stage IIIB/IIIC (AJCC 7th edition) or IIIB/IIIC/IIID (AJCC 8th edition) patients with Response Evaluation Criteria in Solid Tumors (RECIST) measurable disease (https://www.clinicaltrialsgov/ct2/show/ study/NCT02519322. Date accessed: 05/30/18; https://www.clinicaltrialsgov/ct2/show/study/NC T02977052. Date accessed: 05/30/18; https:// www.clinicaltrialsgov/ct2/show/NCT02858921. Date accessed: 05/30/18; https://www.clinical trialsgov/ct2/show/study/NCT03259425. Date accessed: 05/30/18; Huang et al.; Rozeman et al. 2017; Amaria et al. 2018a) (nodal disease with short axis of at least 15 mm), although some trials have focused on attempting to render unresectable disease resectable with neoadjuvant therapy (https://www.clinicaltrialsgov/ct2/show/study/NC T02036086. Accessed 30 May 2018; Haanen et al. 2017). Currently there is not enough clinical

E. Z. Keung et al.

data to suggest which patients will optimally benefit from neoadjuvant therapy based on baseline factors such as tumor burden or mutational status, and collection of these data with plans for pooling across trials is important for future analyses. Most recent neoadjuvant trials have utilized 6–12 week durations of neoadjuvant treatment prior to surgery. This timeframe has been utilized in order to balance the need for time for treatment to take effect without losing the surgical window of opportunity. Ultimately, standardizing the design of neoadjuvant trials will be important moving forward in being able to meaningfully pool data across trials to allow for analysis of patient outcomes.

Rationale for Neoadjuvant Therapy We define neoadjuvant therapy as the preoperative administration of systemic therapy (cytotoxic chemotherapy, molecularly targeted agents, or immunotherapy) for a defined length of time with a plan to carry out definitive resection thereafter, in the absence of tumor progression, metastasis, or treatment-related toxicity that precludes surgery. Postoperative adjuvant therapy with the same or different agents may or may not be offered, and decisions regarding postoperative therapy may or may not be based on the clinical and histologic response to preoperative treatment. We contrast this with “upfront” systemic therapy, in which the primary treatment of potentially resectable tumor is systemic (Sondak and Khushalani 2017), without a predefined duration of administration and with no specific commitment to surgery, which nonetheless might be carried out after maximal response to therapy or tumor progression had been observed. Cytotoxic chemotherapy administered in the preoperative setting is widely accepted as a standard of care for a number of locally advanced solid tumors including breast (National Comprehensive Cancer Network 2018a; Fisher and Brown 1997; Mauri et al. 2005; Rastogi et al. 2008; Von Minckwitz et al. 2012; Gianni et al. 2012; Schneeweiss et al. 2013; Cortazar et al. 2014), head and neck (National Comprehensive

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

771

Cancer Network 2018b; Salvador-Coloma and Cohen 2016), bladder (National Comprehensive Cancer Network 2018f; Grossman et al. 2003; International Collaboration of Trialists et al. 2011), gastro-esophageal (National Comprehensive Cancer Network 2018d; National Comprehensive Cancer Network 2018e; van Hagen et al. 2012; Shapiro et al. 2015; Al-Batran et al. 2016; Franco et al. 2017; Anderegg et al. 2017; Noordman et al. 2018), and rectal cancers (National Comprehensive Cancer Network 2018c; Fernández-Martos et al. 2010; Cercek et al. 2014; Allegra et al. 2015). There are a number of potential advantages to the neoadjuvant approach. One major advantage is the potential to downstage or decrease tumor size preoperatively (Fig. 2). This may allow for decreased morbidity and in some cases a more limited extent of surgery required to achieve complete resection of tumor. A notable example of this is the successful application of neoadjuvant therapy for breast conservation therapy for breast cancer (Killelea et al. 2015; Golshan et al. 2015; Keung et al. 2018).

Neoadjuvant therapy uniquely provides the opportunity to deeply assess an individual patient’s response to the applied therapy. Evaluation of a patient’s radiographic and pathologic response to neoadjuvant therapy may provide information to help tailor postoperative treatment recommendations. The potential to assess the mechanism of treatment response or resistance to neoadjuvant therapy offers an opportunity to explore the relative efficacy of different agents and/or regimens, which may become particularly important given the emergence of new therapies for melanoma. Introduction of systemic therapy earlier, without waiting for surgery and adequate recovery from the procedure and its complications (potentially including a period of relative immunosuppression) (Hogan et al. 2011; Alieva et al. 2018), may also be beneficial in preventing development of drug resistance in residual micrometastases, and potentially improve RFS and OS compared to adjuvant therapy or waiting for disease recurrence to initiate the same therapy (Davar et al. 2013; Liu et al. 2016; Melero et al. 2016). Additionally, in the case of immunotherapies,

Fig. 2 56-year-old man presented with multiple clinically detected lymph nodes in the right neck (a). He was treated with single agent nivolumab 240 mg IVevery 14 days for 4 doses and had a 33% tumor reduction on imaging at the

completion of neoadjuvant therapy (b). At the time of surgery, no viable tumor was identified, consistent with a pathologic complete response

772

treatment given in the neoadjuvant setting in the presence of a higher burden of tumor and tumor antigens may potentially engender a greater and more specific antitumor immune response (Liu et al. 2016). Finally, the neoadjuvant approach to systemic therapy provides an opportunity to collect longitudinal blood and tumor specimens across the duration of treatment and to investigate mechanisms of response and resistance to the applied regimen. This approach of biospecimen collection and analysis can facilitate biomarker assessment and provide critical information to investigators on whether the treatment is achieving its pharmacodynamic goals. Additionally, regulatory agency approval from a neoadjuvant therapy trial has been achieved since the FDA approved use of Pertuzumab (Perjeta, Genentech) on the basis of robust clinical results and biomarker analysis for breast cancer (Gianni et al. 2012; Schneeweiss et al. 2013).

The History of Neoadjuvant Therapy Use in Melanoma Neoadjuvant Biochemotherapy In the era prior to the approval of new agents with proven overall survival benefit (i.e., before 2011), biochemotherapy had historically been the treatment for unresectable metastatic melanoma with the highest observed ORR (Table 1) (Tarhini et al. 2011; Amaral et al. 2018). The most widely studied biochemotherapy regimen was a multidrug combination of cytotoxic chemotherapy with cisplatin 20 mg/m2 IV on days 1–4, vinblastine 1.5–1.6 mg/m2 on days 1–4, and dacarbazine 800 mg/m2 on day 1, plus biologic therapy (immunotherapy) with interleukin-2 9 MIU/m2/ day IV by 96 h continuous infusion on days 1–4 and interferon alfa (IFN-α) 5 MIU/m2 subcutaneously on days 1–5. Two cycles over 6 weeks were traditionally administered in the neoadjuvant setting, with the potential for two postoperative doses. In two phase II studies of this regimen in the neoadjuvant setting, 28 of 64 patients (44%) (Buzaid et al. 1998) and 14 of 36 patients (38.9%)

E. Z. Keung et al.

(Gibbs et al. 2002) achieved objective clinical responses, virtually all partial responses. The significant treatment related toxicities (hematologic toxicities, infection risk) limited widespread use of this regimen.

Neoadjuvant High Dose Interferon Several phase II studies have evaluated high dose IFN-α2b (HDI) in the neoadjuvant setting (Moschos et al. 2006; Amaral et al. 2018). In one trial, patients with clinical stage III disease received IFN-α2b IV (20 MIU/m2/d, 5 days per week) for 4 weeks followed by complete resection and postoperative subcutaneous IFN-α2b (10 MIU/m2 3 times per week) for 48 weeks (Moschos et al. 2006). Eleven of 20 patients (55%) with clinical stage III melanoma had an objective clinical response and 3 (15%) had a pathologic complete response (pCR). This trial included detailed molecular signaling and immunological corollary analyses of the pre- and post-treatment tissues, which revealed immunological effects along with potent suppression of STAT3 with this therapy, associated with dendritic cell and T cell infiltration of the tumor tissue. This has been followed by neoadjuvant trials of HDI and ipilimumab (Tarhini et al. 2016), as well as HDI and pembrolizumab (https://www.clinicaltrialsgov/show/NCT02339324. Date accessed: 05/30/18), which have recently been reported and results will be discussed in the immune checkpoint inhibitors section. The NAM-trial (NCT01341158) is a neoadjuvant phase II trial assessing the efficacy and tolerability of different IFN-α doses. Patients in this study received IFN-α given subcutaneously as flat dosages (3, 9, or 18 MIU) 5 days per week for 4 weeks. This trial has been completed but the results are not yet available.

The Current State of Neoadjuvant Therapy Following recent advances with targeted therapies and immunotherapies for unresectable metastatic melanoma and in the postoperative adjuvant setting for stage III melanoma, neoadjuvant trials are

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

773

Table 1 Clinical trials of neoadjuvant systemic therapy in melanoma of historic interest

Study name A phase II study of neoadjuvant concurrent biochemotherapy in melanoma patients with localregional metastases

Phase II

Number of Patients 65

A phase II study of neoadjuvant biochemotherapy for stage III melanoma

II

48

Neoadjuvant treatment of regional stage IIIB melanoma with high-dose interferon alfa-2b induces objective tumor regression in association with modulation of tumor infiltrating host cellular immune responses

II

20

Study arms 2–4 cycles of biochemotherapy prior to surgery. Each cycle was comprised of cisplatin 20 mg/m2 IV on days 1–4; vinblastine 1.5 mg/ m2 IV on days 1–4; dacarbazine 800 mg/m2 IV on day 1; interleukin-2 9106 IU/m2/day IV over 96 h continuous infusion on days 1–4; and interferon-alfa 2a 5  106 IU/m2/day SC on days 1–5 every 3 weeks Two cycles of biochemotherapy prior to and after complete lymph node dissection. Each cycle was comprised of cisplatin 20 mg/m2 IV on days 1–4; vinblastine 1.6 mg/ m2 IV on days 1–4; dacarbazine 800 mg/m2 IV on day 1; interleukin-2 9106 IU/m2/day IVover 24 h on days 1–4; and interferonalfa 5  106 IU/m2/ day SC on days 1–5 every 3 weeks Standard IV HDI 20  106 units/m2 5 days per week for 4 weeks followed by complete lymphadenectomy and standard maintenance SC HDI 10  106 units/ m2 3 per week for 48 weeks

Publication year 1998

PMID 9918417

Clinical responses observed in 38.9% of patients (14/36) with 13 PRs (36.1%), 1 CR (2.8%), and 4 pCRs (11.1%).

2002

11900232

ORR 55% (11/20), 3 cPRs (15%)

2006

16809739

Main outcomes PR in 28/64 clinically evaluable patients (44%). Among 62 patients whose response was assessed histologically, 27 PRs (43.5%) and 4 pCRs (6.5%).

CR complete response, HDI high dose interferon alfa-2b, IU international units, ORR objective response rate, pCR pathologic complete response, PR partial response, SC subcutaneous

774

underway to evaluate these agents in patients with high-risk resectable or oligometastatic melanoma (Table 2). To date, however, there are few completed neoadjuvant trials. In the sections to follow, we will present the available published results of contemporary neoadjuvant therapy trials for melanoma.

Neoadjuvant Targeted Therapies A number of phase II studies of neoadjuvant targeted therapy, either alone or in conjunction with postoperative adjuvant therapy, have been initiated for patients with high-risk resectable melanoma (Table 2). These trials are predominantly evaluating BRAF/MEK combination therapy and are summarized below.

Dabrafenib Plus Trametinib There are several studies evaluating the combination of dabrafenib (150 mg twice daily) and trametinib (2 mg daily) in the neoadjuvant setting. In the Combi-Neo trial (NCT02231775), patients with stage IIIB/C (AJCC 7th edition) or oligometastatic stage IV BRAF-mutant melanoma were randomized to either surgery followed by standard of care adjuvant therapy (which could be observation alone, HDI for 1 year, pegylated IFN-α2b for up to 5 years, high-dose ipilimumab, or biochemotherapy) or neoadjuvant combination dabrafenib and trametinib (8 weeks) followed by 44 weeks of the same combination postoperatively. Interim analysis of the initial 21 patients enrolled demonstrated that neoadjuvant dabrafenib/trametinib was associated with a high response rate (77%) and high pathological complete response rate (pCR rate, 58%) (Fig. 3). At a median followup of 18.6 months, neoadjuvant targeted therapy was associated with significantly better RFS than patients randomized to surgery followed by possible postoperative adjuvant therapy (HR 60.2, 95% CI 6.7-7965, p < 0.0001) (Fig. 4) (Amaria et al. 2018a). This regimen also showed a manageable safety profile for dabrafenib and

E. Z. Keung et al.

trametinib, similar to that observed in patients with metastatic melanoma, with pyrexia and chills as the most commonly reported adverse events (AEs), none of which interfered with the ability to carry out planned surgery. These encouraging early results led to an early stop to the surgery only arm with trial continuation as a single-arm study of neoadjuvant dabrafenib and trametinib. A second ongoing phase II clinical trial neoadjuvant trial (NCT01972347) was presented at ESMO 2017. In this study, patients with histologically confirmed, resectable bulky stage IIIB/C (AJCC 7th Edition) BRAF V600–mutant melanoma received dabrafenib plus trametinib for 12 weeks before and 40 weeks after resection. With a median follow-up time of 17.6 months, 17 of 35 patients (49%) achieved pCR and 18 patients (51%) achieved a metabolic complete response (CR). However, 37% of patients developed recurrent melanoma (median time to recurrence, 11.7 months), and two patients died due to melanoma at 12.1 and 15.2 months. The median RFS was 20.1 months (95% CI, 17.7-not reached).

Neoadjuvant Immune Checkpoint Inhibitors Clinical investigations evaluating immune checkpoint inhibitors in the neoadjuvant setting in patients with high-risk melanoma are also underway or completed. There are currently three ongoing phase II clinical trials to evaluate either antiPD-1 antibody alone (NCT02306850) (https:// www.clinicaltrialsgov/show/NCT02339324. Date accessed: 05/30/18) or in combination with antiCTLA-4 antibody in the neoadjuvant setting for patients with stage III/IV melanoma (NCT02519322, NCT02977052) (https://www. clinicaltrialsgov/ct2/show/study/NCT02519322. Date accessed: 05/30/18; https://www.clinical trialsgov/ct2/show/study/NCT02977052. Date accessed: 05/30/18) (Table 2). These studies should help establish the benefit of and delineate

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

775

Table 2 Clinical trials of neoadjuvant systemic therapy in melanoma Clinical trials. gov Identifier Study name Targeted therapy NCT01972347 Neoadjuvant dabrafenib and trametinib for AJCC stage IIIB/C BRAF V600 mutation positive melanoma NCT02231775 Neoadjuvant and adjuvant dabrafenib and trametinib in patients with clinical stage III melanoma (Combi-Neo) NCT02036086 Study of neoadjuvant use of vemurafenib plus cobimetinib for BRAF mutant melanoma with palpable lymph nodes metastases NCT02303951 Neoadjuvant vemurafenib + cobimetinib in melanoma: NEO-VC

Immune checkpoint therapy NCT00972933 Immunogenicity and biomarker analysis of neoadjuvant ipilimumab for melanoma

NCT02434354

A tissue collection study of pembrolizumab (MK3475) in subjects with resectable advanced melanoma

Phase

Trial status (estimated accrual)

Study arms

Main outcomes

Dabrafenib 150 mg BID orally and Trametinib 2 mg QD orally for 52 weeks (12 weeks neoadjuvant) Dabrafenib 150 mg BID orally and Trametinib 2 mg QD orally for 52 weeks (8 weeks neoadjuvant)

1. Viable melanoma tissue postoperatively 2. Relapse-free survival 3. OS

II

Active, not recruiting (35)

II

Recruiting (78)

II

Recruiting (20)

Vemurafenib 960 mg BID orally and Cobimetinib 60 mg QID orally for 52 weeks (8 weeks neoadjuvant)

1. Post-therapy resectability rate 2. DMFS 3. DFS 4. OS

II

Recruiting (110)

Vemurafenib 960 mg BID orally and Cobimetinib 60 mg (3 weeks on drug, 1 week off) QID orally (up to 18 weeks neoadjuvant)

1. Post-therapy resectability rate 2. PFS 3. ORR 4. OS

I

Completed (59)

Induction: Ipilimumab 10 mg/kg IV day 0, 21 (baseline, week 3) Maintenance: Ipilimumab 10 mg/kg IV days 63 (+28 days) and, 84 (+28 days) – (3 weeks apart, starting 2–4 weeks following definitive lymphadenectomy)

I

Recruiting (30)

1 dose of neoadjuvant pembrolizumab 200 mg followed by complete resection and then a year of adjuvant pembrolizumab

1. Study the effects of ipilimumab upon the host immune response in nodal metastatic melanoma and in the peripheral blood 2. Evaluate the clinical efficacy of preoperative neoadjuvant therapy with ipilimumab in high-risk clinically and pathologically node-positive melanoma patients 1. Number of adverse events

1. Relapse-free survival with and without pCR

(continued)

776

E. Z. Keung et al.

Table 2 (continued) Clinical trials. gov Identifier NCT02306850

NCT02519322

NCT02977052

Trial status (estimated accrual) Recruiting (15)

Study name Neoadjuvant pembrolizumab for unresectable stage III and unresectable stage IV melanoma (NeoPembroMel) Neoadjuvant and adjuvant checkpoint blockade in patients with clinical stage III or oligometastatic stage IV melanoma

Phase II

II

Recruiting (40)

Optimal neo-adjuvant combination scheme of ipilimumab and nivolumab (OpACINneo)

II

Recruiting (90)

Study arms Pembrolizumab 200 mg infusions every 3 weeks for at least 24 weeks and up to 2 years depending on response Arm A: Neoadjuvant phase: Nivolumab 3 mg/kg IV every 2 weeks on weeks 1, 3, 5, and 7 prior to surgical excision. Adjuvant Phase: Nivolumab 3 mg/kg IV every 2 weeks postoperatively for 6 months. Arm B: Neoadjuvant Phase: Nivolumab 1 mg/kg IV combined with Ipilimumab 3 mg/kg IV every 3 weeks on weeks 1, 4, and 7 prior to surgical excision. Adjuvant Phase: Nivolumab 3 mg/kg IV every 2 weeks postoperatively for 6 months. Arm C: Nivolumab 480 mg and Relatlimab 160 mg every 28 days for 2 doses prior to surgical excision, then 10 doses of Nivolumab 480 mg with Relatlimab 160 mg, postoperatively. Arm A: Two courses standard combination of ipilimumab 3 mg/ kg and nivolumab 1 mg/kg q3wk prior to surgery at week 6. Arm B: Two courses ipilimumab 1 mg/kg and nivolumab 3 mg/ kg q3wk prior to

Main outcomes 1. Resectability rate 2. Response by RECIST criteria

1. Pathologic response 2. Immunologic response

1. Safety 2. Response by RECIST criteria 3. Pathological response 4. RFS

(continued)

Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

777

Table 2 (continued) Clinical trials. gov Identifier

Study name

Phase

Trial status (estimated accrual)

Study arms

Main outcomes

surgery at week 6. Arm C: Two courses of ipilimumab 3 mg/kg q3wks, directly followed (>2 h and 41  C can be achieved) General anesthesia (GA) required

Isolated limb infusion Technically simple Percutaneous vascular catheter insertion in radiology department Approximately 1 h No perfusionist required and fewer total staff Equipment requirements modest Well-tolerated by medically compromised, frail, and elderly patients Can be performed selectively in occlusive vascular disease Not difficult to perform a repeat procedure Systemic metastases not a contraindication Low pressure system, effective vascular isolation with tourniquet Progressive hypoxia and acidosis Usually not possible to raise limb temperature above 40  C Possible with regional anesthesia, GA preferred

doses than systemic regimens) can be utilized. In ILI, dactinomycin, an inhibitor of DNA transcription, is often administered in addition to melphalan, based on results from MIA utilizing the combination during HILP, where an OR rate of 73% was achieved, without increasing limb toxicity (Lindner et al. 2002). Recently, there has been interest in using temozolomide, an imidazotetrazine derivative of dacarbazine, during ILI (Ueno et al. 2004). Like melphalan, the mechanism of action of temozolomide involves disruption of DNA replication through alkylation. A multicenter phase I dose-escalation trial of ILI using temozolomide demonstrated a favorable safety profile. OR rates were low (15.8%), but the study population was small and not designed to evaluate drug efficacy (Beasley et al. 2015). Therefore, further studies

835

involving large sample sizes are required to assess the efficacy of temozolomide during ILI.

Pharmacokinetics of Melphalan During Isolated Limb Infusion The plasma concentration of melphalan in the limb during ILI falls in a monoexponential fashion, suggesting rapid uptake by the tissues (Roberts et al. 2001b). This is in keeping with in vitro studies, demonstrating uptake of melphalan into melanoma cells as a rapid, active, temperature-dependent process that achieves saturation after 10 min (Parsons et al. 1981). The mean residence time and elimination half-life of melphalan during ILI were 21–35 min and 15–25 min, respectively (Roberts et al. 2001b). In vitro studies have shown that the hypoxic conditions during ILI enhance the cytotoxic effect of melphalan by a factor of approximately 1.5 and the combination of hypoxia and acidosis can increase the effect by a factor of 3 (Siemann et al. 1991; Skarsgard et al. 1995). HILP studies have also shown that by administering glucose to the isolated circuit, the intracellular pH in the tumor can be decreased, with a concomitant increase in the response rate (Van der Merwe et al. 1993). However, conflicting data from a HILP in vitro model, examining the effect of a variety of factors on the sensitivity of melanoma cells to melphalan, showed that a pH as low as 6.0 had no significant impact on cell survival (Clark et al. 1994). On the basis of these reports, it appears that the significance of hypoxia and acidosis during ILI remains to be fully elucidated. Studies of the pharmacokinetics of HILP and ILI have demonstrated similar wide variations of plasma melphalan concentrations. Cheng et al. examined pharmacokinetics by obtaining plasma melphalan drug levels during HILP (Cheng et al. 2003). Five of 14 patients suffered Wieberdink limb toxicity grade III/IV (Wieberdink et al. 1982), and marked differences in melphalan plasma concentrations were observed despite using similar dosing guidelines. The strongest predictor of toxicity was the ratio of estimated limb volume (Vesti) to steady-state (Vss) limb

836

drug volume of distribution, whereas the area under the curve and peak plasma concentration failed to correlate with toxicity. All toxicity was seen in patients whose Vesti/Vss ratio was over 4: five of seven patients with a ratio greater than 4 had grade III/IV limb toxicity. In an initial experience with drug pharmacokinetics at Duke University Medical Center (Duke UMC), similar variability was found. Toxicity was also related to overestimation of limb volume compared to steady-state limb drug volume of distribution (Beasley et al. 2008). An analysis of 185 ILIs from MIA found that patients with a body mass index (BMI) of >25 kg/m2 experienced greater limb toxicity (grade III/IV) (Kroon et al. 2009c). This finding may also indicate an overestimation of the volume of distribution. Since melphalan uptake is higher in muscle as opposed to fat, the skin and subcutaneous tissues are exposed to a relatively higher dose of melphalan when concentrations are based on limb volumes only because overweight patients have a lower muscle-to-fat ratio (Kroon et al. 2008, 2009c; Klaase et al. 1994; Scott et al. 1990).

Melphalan Dosage and Ideal Body Weight The observations discussed above suggest that the therapeutic index of melphalan could be optimized through a better understanding of its pharmacokinetics in individual patients, with patients who fit the profile for a high Vesti/Vss given a lower melphalan dose. In view of this, some centers have modified melphalan dosage according to IBW (Beasley et al. 2008). This calculation is performed by multiplying the melphalan dose (7.5 mg/L for a lower limb; 10 mg/L for an upper limb) by the ratio of IBW to actual body weight. Patients at Duke UMC who had their melphalan dose corrected for IBW experienced less variability in melphalan plasma concentrations, and a significant decrease in toxicity was observed ( p = 0.024) when melphalan dose was corrected for IBW without adversely affecting tumor response (McMahon et al. 2009).

G. M. Beasley et al.

Use of Microdialysis During Isolated Limb Infusion Microdialysis is a technique that enables drug concentrations to be monitored in various tissues to investigate the relationships between melphalan concentrations in plasma, the interstitium, and tumor tissue (Wu et al. 2000). In patients undergoing ILI at MIA, microdialysis catheters (CMA60/CMA70; CMA, Solna, Sweden) were inserted subcutaneously into normal and tumor tissues before the start of ILI (Thompson et al. 2001). A microdialysis pump (CMA 106; CMA) maintained a constant infusion of fluids while melphalan concentrations in the samples were measured using high-performance liquid chromatography (Wu et al. 1995). The study showed that the peak melphalan concentrations in plasma were higher than in subcutaneous tissues and tumor tissues. This technique enables melphalan concentrations to be monitored in subcutaneous tissues and tumor deposits and therefore may help to optimize ILI conditions and improve tumor response; however, further studies involving larger sample sizes are required.

Toxicity and Side Effects Following Isolated Limb Infusion Locoregional Side Effects of Isolated Limb Infusion In general, ILI is a well-tolerated procedure. As with HILP, superficial desquamation of the skin often occurs after 2–3 weeks, and residual pigmentation of the limb may persist for months. If the foot or hand has not been excluded by an Esmarch bandage or distal pneumatic tourniquet, as is often possible, loss of the epidermis of the sole of the foot or palm of the hand may occur, leaving a delicate and sensitive new skin surface exposed for weeks until the area is again covered by normal plantar or palmar skin. Furthermore, loss of toe or fingernails of the treated limb may occur 3–4 months after treatment, as well as the loss of hair in the limb (Thompson et al. 1994b).

Isolated Limb Infusion for Melanoma

Limb Toxicity Following Isolated Limb Infusion The Wieberdink toxicity scale, historically used for HILP, is also applicable after ILI (Table 3) (Wieberdink et al. 1982). At MIA, ILI usually results in mild-to-moderate limb toxicity; 56% and 39% of patients experienced Wieberdink grade II and grade III limb toxicity, respectively, while only 3% experienced grade IV toxicity (Kroon et al. 2009c). Although this incidence of limb toxicity is at the higher end of the spectrum of that reported following HILP, long-term morbidity is less frequently observed and less severe after ILI compared to HILP. Fasciotomies due to threatened or actual severe limb toxicity, for instance, are rarely necessary after ILI, and from all reported series, only one patient has required an amputation due to toxicity following ILI (grade V limb toxicity) (Kroon et al. 2009c, 2014a; 2016; Beasley et al. 2008, 2012; Brady et al. 2009; Dossett et al. 2016; Santillan et al. 2009; O’Donoghue et al. 2017). Large, contemporary series have reported grade III limb toxicity or higher in less than 30% (Kroon et al. 2016; Beasley et al. 2012; O’Donoghue et al. 2017). An Australian multicenter study evaluating 316 ILI procedures reported grade III limb toxicity in 27% and grade IV toxicity in 3% of the patients, with no amputations due to toxicity, and a recent single-center study from Moffitt Cancer Center (MCC) reported grade III or higher in 12% of the patients. At Duke UMC, toxicity has been assessed according to the National Cancer Institute Common Terminology Criteria for Adverse Events Table 3 Wieberdink toxicity grading (Wieberdink et al. 1982) Grade I Grade II Grade III Grade IV

Grade V

No visible effect Slight erythema and/or edema Considerable erythema and/or edema with blistering Extensive epidermolysis and/or obvious damage to deep tissues with a threatened or actual compartment syndrome Severe tissue damage necessitating amputation

837

Version 3 (CTCAEv3; Table 4) (Beasley et al. 2008; Common Terminology Criteria for Adverse Events 2006). Using the CTCAEv3, the severity of limb toxicity was similar to that reported by other series using the Wieberdink toxicity scale (Table 5). At Duke UMC, ILI was associated with significantly less limb toxicity compared to HILP, after which more patients experienced grade III limb toxicity or higher (44% after HILP vs. 18% after ILI; p = 0.009) including nine compartment syndromes and two amputations. They reported that limb toxicity was further reduced by melphalan dose correction for IBW (Beasley et al. 2008). Following ILI, no relationship has been found between more severe limb toxicity and complete response (CR), duration of response, or overall survival (OS), but a relationship was observed between Wieberdink grade III/IV limb toxicity and overall response (OR) at MIA (Beasley et al. 2008; Kroon et al. 2008, 2009c). It is interesting to note that the partial response (PR) rate at Duke UMC (14%), which largely accounted for their lower OR rate (44%), was much lower than the PR rate at MIA (46%; OR 84%). One hypothesis that would explain this lower PR rate is that, while the major determinant of a CR is tumor chemosensitivity, a PR may be related to maximal chemotherapy, a delivery which is associated with more limb toxicity. Since at Duke UMC the majority of patients who had chemotherapy dose correction for IBW experienced less toxicity, it is possible that they may have had a favorable response with a higher melphalan dose. Various pharmacokinetic variables have been shown to predict limb toxicity after ILI. An analysis at MIA revealed that a high peak melphalan level, a high final melphalan level, and a larger melphalan concentration area under the curve in the isolated limb were significantly associated with more severe limb toxicity (Kroon et al. 2009c). Also, a smaller increase of the CO2 level in the isolated circuit during ILI was found to be significantly associated with increased limb toxicity. This finding was surprising given the fact that a bigger rise in CO2 had earlier shown to improve response rates (Kroon et al. 2008). These results, demonstrating the abovementioned effect of hypoxia and acidosis both on toxicity and

Musculoskeletal/ soft tissue

Dermatology/ skin

Short name Dermatitis

Ulceration

Myositis

Adverse event Rash

Ulceration

Myositis (inflammation/ damage of muscle) Mild pain not interfering with function



Toxicity grade 1 Faint erythema or dry desquamation

Pain interfering with function, but not interfering with activities of living

Toxicity grade 2 Moderate to brisk erythema; patchy, moist desquamation, mostly confined to skin folds and creases; moderate edema Superficial ulceration 1,000 U/L in an attempt to avoid severe limb toxicity (Kroon et al. 2014b).

Systemic Toxicity and Complications of Isolated Limb Infusion Serious systemic side effects are rare after ILI, with no occurrence of bone marrow depression and only occasional mild postoperative nausea and vomiting, which resolves quickly with conservative management. The reasons for the small number of patients suffering from systemic toxicity after ILI are mainly attributed to the low influx of chemotherapy from the isolated limb to the systemic circulation. Influx of chemotherapy to the systemic circulation is prevented by the reliability of limb isolation with the pneumatic

tourniquet, the thorough flushing of the limb after completion of the ILI, and the low pressure in the isolated circuit, which is much lower than the systemic blood pressure. At MIA, systemic melphalan was detected at a very low rate in a minority of the patients: 5 lesions had improved OS rates ( p = 0.010) (Kroon et al. 2008). In a US multicenter study, BOD was defined as being low if patients had less than ten tumors with none greater than 2 cm, and patients were classified as having a high BOD if they had more than ten

Isolated Limb Infusion for Melanoma

843

Fig. 6 (a) Overall survival (months) following isolated limb infusion in 316 Australian patients (Kroon et al. 2016). (b) Survival (months) of patients after a complete response (CR; solid line) compared with a partial response (PR; dotted line) after isolated limb infusion ( p = 0.014; HR 2.42; 95% CI 1.67–3.09) (Kroon et al. 2016)

lesions or any single lesion larger than 2 cm (Muilenburg et al. 2015). On multivariate analysis, patients with a low BOD were 3.5 times more likely than high BOD patients to have a favorable response to ILI (odds ratio 3.5, p < 0.001). Moreover, patients with a low BOD experienced a significantly increased median LRFI of 6.9 months compared to 3.8 months for high BOD patients ( p = 0.047), although this finding did not translate into improved OS. The findings that several of the patient factors were independent predictors of LRFI and OS, including depth of tumor infiltration, number of lesions/BOD, and Breslow thickness of the primary melanoma, can possibly be explained by the fact that they are derivatives of stage of disease, which has been shown to be a prognostic factor for duration of response and OS (Kroon et al. 2008, 2016). The longer OS of patients who achieved a CR following ILI could

be explained by the fact that they tended to have a lower stage of disease and suggests that the underlying tumor biology may be associated with greater chemosensitivity (Kroon et al. 2008; Aloia et al. 2005; Sanki et al. 2007). At Duke UMC it was shown that patients who responded to ILI had significantly smaller limb volumes (6.4  2.2 L) than those who did not respond (8.1  3.4 L) ( p = 0.043), and at MCC an increased response was seen after upper limb ILI compared to lower limb ILI (Beasley et al. 2008; O’Donoghue et al. 2017). These observations again suggest an important role of drug distribution, as seen in the microdialysis studies that demonstrated that drug distribution was better achieved in smaller limbs than larger limbs, leading to a better drug delivery to the skin and subcutaneous tissues in patients with smaller limbs (Kroon et al. 2009c; Klaase et al. 1994; Thompson

844

et al. 2001). Ultimately, the identification of both favorable tumor and patient-related factors that influence outcomes will help to stratify patients into groups likely to derive the greatest benefit from ILI. In the MIA series, several intraoperative factors with a predictive value for response were identified. A higher melphalan concentration at the conclusion of the procedure was associated with significantly improved CR ( p = 0.013) and OR ( p = 0.022) (Kroon et al. 2008). A greater difference in pCO2 in the isolated circuit between the commencement and the conclusion of the procedure was also associated with an improved CR rate ( p = 0.017), and a tourniquet time >40 min was a prognostic factor for OS and showed a trend toward an increased OR rate ( p = 0.074) (Lindner et al. 2002; Kroon et al. 2008). As previously mentioned, these findings could possibly be related to the synergism of hypoxia and acidosis with melphalan. There was a trend toward a higher OR in patients who had a larger increase from the initial to the final subcutaneous temperatures in the limb during ILI ( p = 0.062). Hyperthermia and its synergistic cytotoxic effect with melphalan were first described over 40 years ago (Cavaliere et al. 1967; Stehlin 1969). During ILI, however, it is usually not possible to achieve true hyperthermia (i.e., tissue temperatures exceeding 41.0  C) in the treated limb because of the low-flow rate due to the high resistance of the small-caliber catheters in the isolated circuit. However, previous HILP studies have suggested that it is mainly the maintenance of normothermia and the avoidance of hypothermia that are most important, rather than the attainment of high limb temperatures. In view of this, the mildly hyperthermic limb temperatures of 38–39  C achieved during ILI may actually be beneficial, as hyperthermic limb temperatures have the disadvantage of causing more rapid melphalan degradation and greater limb toxicity (Chang et al. 1978; Kroon 1988). Postoperatively, a high CK level showed a significant association with OR rate ( p = 0.029), and limb toxicity grade predicted the OR rate ( p = 0.002) but not the CR rate, LRFI, or OS. In contrast, the Duke UMC series reported

G. M. Beasley et al.

that neither limb temperatures, tourniquet time, nor postoperative CK level predicted outcomes (Beasley et al. 2008). Larger studies will be needed to further investigate these relationships.

Special Isolated Limb Infusion Regimens and Indications Special ILI regimens include a planned double ILI, a repeat ILI procedure for disease recurrence, ILI for palliation in patients with AJCC stage IV disease, and ILI as induction therapy. One study examined a planned double ILI at a median of 2–8 weeks apart (Lindner et al. 2004). This protocol achieved a CR of 41% and a PR of 47%, with a median duration of response of 18 months in 47 patients. Because a planned double ILI increased limb toxicity without a significant increasing efficacy, it was concluded that performance of a single ILI remained the preferred treatment option for melanoma confined to a limb (Lindner et al. 2004). A repeat procedure after an initial ILI in patients with disease recurrence or progression may be beneficial, especially if there had been a favorable response to the initial ILI. For patients who did respond to an initial ILI, a repeat procedure is unlikely to provide much benefit. In those cases, systemic therapies will need to be considered. As mentioned before, the minimally invasive character of ILI allows a repeat procedure to be performed with relative ease, in contrast to HILP (Kroon et al. 2009a; Chai et al. 2012; Raymond et al. 2011). A multi-institutional series by Chai et al. evaluated the use of repeat HILP versus repeat ILI after disease progression: 3 patients (7%) had a repeat HILP, 10 (23%) had a HILP following an initial ILI, and 12 (27%) had an ILI following an initial HILP (Chai et al. 2012). Most patients tolerated repeat regional chemotherapy well without increased toxicity or LOS, and no statistical difference in response rates or OS was noted when comparing repeat ILI or HILP procedures. Another series showed that patients who experienced regional recurrences after an initial regional treatment were more likely to achieve a CR after repeat HILP (50%, n = 10)

Isolated Limb Infusion for Melanoma

compared with repeat ILI (28%, n = 18) (Raymond et al. 2011). However, the likelihood of grade IV limb toxicity was greater after HILP (2 of 62) than after ILI (0 of 122). Given that response rates and duration of response following ILI and HILP are generally similar and ILI results in fewer complications due to its minimally invasive nature and causes less limb toxicity grades, many centers use ILI and consider repeat ILI for patients who had a good response to the initial ILI (Dossett et al. 2016). Because of its minimally invasive nature, low limb toxicity, and low complication rates, ILI can also be considered as a palliative procedure to avoid limb amputation in patients with both symptomatic limb disease and distant melanoma metastases in order to achieve limb salvage and increase quality of life. In an ILI MIA study in patients with AJCC stage IV disease, limb preservation was achieved in 86% of the patients (n = 37) (Kroon et al. 2009b). In this time of effective systemic therapy for melanoma metastases, combination treatment of locoregional therapy through ILI and systemic therapy will very likely be considered in the near future in this subset of patients for increased efficacy both on systemic and limb disease. In the two paragraphs below, ILI in combination with systemic therapies is discussed. Another mechanism by which ILI can help achieve limb preservation is by using it as induction therapy. This can convert unresectable disease into resectable disease, and simple local treatment of the remaining lesions, such as excision, laser ablation, electrodessication, and injection with Rose Bengal (PV-10), can then achieve effective disease control (Thompson et al. 2015; Huismans et al. 2016). After a PR, for instance, resection of residual limb disease achieved LRFI and OS rates similar to those observed following a CR after ILI alone (Wong et al. 2014). Finally, ILI can safely and effectively be used in the elderly and in patients with upper extremity melanoma. Particularly in elderly patients, ILI appears to be an attractive and safe procedure compared to HILP, since older patients experienced less limb toxicity compared with younger patients (Wieberdink grade III/IV toxicity 36%

845

vs. 51%; p = 0.009), but efficacy, systemic toxicity, complications, and long-term morbidity were similar in a recent multicenter Australian study (Kroon et al. 2017). ILI for upper extremity melanoma is associated with similar CR rates but lower toxicity than lower limb ILI despite comparable methods, suggesting a particularly important role for ILI in the management of upper extremity disease (Beasley et al. 2012; O’Donoghue et al. 2017).

Novel Isolated Limb Infusion Regimens Given its minimally invasive character and the easy visual assessment of tumor response and access for biopsies of in-transit metastases, ILI is an ideal model to explore novel therapeutic agents and therapy approaches (Lidsky et al. 2014). Using temozolomide as an alternative to melphalan was explored in a phase I study in patients who had previously failed ILI with melphalan. Patients treated with the maximum tolerated temozolomide dose experienced low regional toxicity, but without an increase in OR (Beasley et al. 2015). Another interesting strategy is the use of systemic modulators to augment the cytotoxic effects of regional chemotherapy administered by ILI. In a prospective multicenter phase II trial, 45 patients received 2 doses of systemic ADH-1 (N-cadherin antagonist) in combination with a standard melphalan ILI. CR was seen in 38% of the patients and an OR in 60% without increased toxicity, compared with an OR of 40% with melphalan only previously achieved at the same institution (Beasley et al. 2011a). Following promising results of systemic use of the multityrosine kinase inhibitor sorafenib in combination with dacarbazine, another study used systemic sorafenib in combination with ILI. Results, however, were disappointing: in 20 patients the addition of sorafenib did not augment the response to ILI, and an increase in limb toxicity was observed (McDermott et al. 2008; Beasley et al. 2011b). Over the last few years, immune checkpoint inhibitors have improved the prognosis for patients with advanced melanoma, making ILI

846

plus checkpoint inhibition a novel strategy for patients with limb melanoma (Callahan et al. 2018; Howie et al. 2015). A phase II trial by Ariyan et al. explored the use of ILI followed by the CTLA-4 blocking antibody, ipilimumab (Ariyan et al. 2018). The concept is that ILI can generate immune cell infiltration and increase the efficacy of CTLA-4 blockade. In 26 patients a CR was seen in 62% and a PR in 23% with a 58% progression-free survival at 1 year. Although these results are promising, 38% of the patients experienced significant ipilimumab systemic side effects, similar to the 45% reported in large trials (Weber et al. 2017).

Future of Isolated Limb Infusion The approval of multiple new, effective systemic therapies for melanoma has dramatically changed treatment strategies for patients with metastatic melanoma. To understand the role of ILI in this new era, a brief review of these new therapies is necessary. Because of the side effects of CTLA-4 inhibitors as mentioned in the previous paragraph, potentially safer strategies may involve ILI in combination with PD-1 blockade or with more specific tyrosine kinase inhibition, such as BRAF+MEK inhibitors that have been proven effective clinically (McArthur et al. 2014; Chapman et al. 2011; Hauschild et al. 2012). Examples include mutation-based targeted therapies such as vemurafenib, a BRAF inhibitor, in combination with trametinib, a MEK inhibitor (Long et al. 2017). For the 50% of melanoma patients with a BRAF mutation, impressive initial responses occur in the majority, but resistance usually develops within 6–9 months (Grob et al. 2015). Immune checkpoint inhibitors represent another category of recently developed systemic therapies that appear to have greater long-term efficacy, as they do not appear to be as susceptible to the development of resistance as mutationbased therapy. Also, they are not restricted to patients harboring specific mutations. Therefore, PD-1 based therapy has now become the cornerstone of systemic therapy for patients with advanced melanoma, yet the development of

G. M. Beasley et al.

both primary and late resistance remains a problem (Franklin et al. 2017). Dual checkpoint inhibitor therapy can result in higher response rates, but toxicity can be considerable and is a concern (Callahan et al. 2018). Another strategy particularly for cutaneous melanoma recurrences are injectable therapies including oncolytic immunotherapy. Talimogene laherparepvec (T-VEC) was found to be effective in a phase III randomized trial showing durable response rates of 16.3% in the T-VEC arm compared to 2.1% in the GM-CSF arm (Andtbacka et al. 2015). Also, intralesional injection of cutaneous and subcutaneous melanoma deposits with PV-10 resulted in a CR of 26% and an OR of 51% (Thompson et al. 2015). Additionally, many other novel injectable agents including other oncolytic viruses and immunocytokines are currently being developed (Brown et al. 2017). Despite these important advances in the treatment of metastatic melanoma, the durable response rates achieved are suboptimal, and novel combination therapies with higher response rates may come with unacceptable toxicity. To date, ILI for metastatic limb melanoma is still an effective option, with only low regional toxicity (Grünhagen et al. 2015). Series from Australian and US centers in nearly 800 patients in total report consistent CR rates following ILI of 35%, with a median durability of 12 months (Kroon et al. 2014a, 2016; Beasley et al. 2009; O’Donoghue et al. 2017). Thus, while much in terms of durability remains unknown for novel therapies, ILI remains an important option for patients with unresectable melanoma confined to the limb. However, if in the future systemic therapies do achieve more durable results with less toxicity, ILI may become important as second-line therapy after failure of systemic therapy, with the advantages of a single, minimally invasive treatment associated with low toxicity and minimal systemic side effects. Additionally, as discussed above, combination strategies of ILI in addition to systemic therapies may increase the efficacy of both procedures when administered individually. Finally, for patients with major comorbidities or elderly patients who are thought to be unfit for systemic treatment or HILP, ILI will remain an effective treatment option.

Isolated Limb Infusion for Melanoma

Conclusions Therapeutic options for patients with melanoma metastases confined to a limb continue to evolve, in parallel with systemic therapies, now being used also as adjuvant and neoadjuvant treatments. Since its first introduction by Thompson et al. over 25 years ago as a minimally invasive alternative to HILP, ILI has been widely applied in patients with in-transit melanoma confined to a limb. In this chapter, we have discussed indications, patient selection, technique, toxicity, and results following ILI. The technique has been used to study the role of hypoxia and hyperthermia in cancer therapeutics and as a model to explore novel combination strategies. In recent years, many effective therapies for patients with metastatic melanoma have been developed and have dramatically increased treatment options. While ILI remains an important and effective option for patients with unresectable limb melanoma, selection is more than ever of utmost importance to select those patients who will benefit most from the procedure, either as an initial therapeutic option, as a second-line treatment, or in combination with systemic therapies.

Cross-References ▶ Adjuvant Systemic Therapy for High-Risk Melanoma Patients ▶ Axillary and Epitrochlear Lymph Node Dissection for Melanoma ▶ Classification and Histopathology of Melanoma ▶ Evolving Role of Chemotherapy-Based Treatment of Metastatic Melanoma ▶ Hyperthermic Regional Perfusion for Melanoma of the Limbs ▶ Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma ▶ Local and Recurrent Regional Metastases of Melanoma ▶ Melanoma Prognosis and Staging ▶ Models for Predicting Melanoma Outcome ▶ Neoadjuvant Systemic Therapy for High-Risk Melanoma Patients

847

▶ Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma ▶ Sequencing and Combinations of Molecularly Targeted and Immunotherapy for BRAFMutant Melanoma ▶ Targeted Therapies for BRAF-Mutant Metastatic Melanoma

References Aloia TA, Grubbs E, Onaitis M et al (2005) Predictors of outcome after hyperthermic isolated limb perfusion: role of tumor response. Arch Surg 140:1115–1120 Andtbacka RH, Kaufman HL, Collichio F et al (2015) Talimogene laherparepvec improves durable response rate in patients with advanced melanoma. J Clin Oncol 33:2780–2788 Ariyan CE, Brady MS, Siegelbaum RH et al (2018) Robust antitumor responses result from local chemotherapy and CTLA-4 blockade. Cancer Immunol Res 6:189–200 Beasley GM, Petersen RP, Yoo J et al (2008) Isolated limb infusion for in-transit malignant melanoma of the extremity: a well-tolerated but less effective alternative to hyperthermic isolated limb perfusion. Ann Surg Oncol 15:2195–2205 Beasley GM, Caudle A, Petersen RP et al (2009) A multiinstitutional experience of isolated limb infusion: defining response and toxicity in the US. J Am Coll Surg 208:706–715 Beasley GM, Riboh JC, Augustine CK et al (2011a) Prospective multicenter phase II trial of systemic ADH-1 in combination with melphalan via isolated limb infusion in patients with advanced extremity melanoma. J Clin Oncol 29:1210–1215 Beasley GM, Coleman AP, Raymond A et al (2011b) A phase I multi-institutional study of systemic sorafenib in conjunction with regional melphalan for in-transit melanoma of the extremity. Ann Surg Oncol 19: 3896–3905 Beasley GM, Sharma K, Wong J et al (2012) A multiinstitution experience comparing the clinical and physiologic differences between upper extremity and lower extremity melphalan-based isolated limb infusion. Cancer 118:6136–6143 Beasley GM, Speicher P, Augustine CK et al (2015) A multicenter phase I dose escalation trial to evaluate safety and tolerability of intra-arterial temozolomide for patients with advanced extremity melanoma using normothermic isolated limb infusion. Ann Surg Oncol 22:287–294 Bland KI, Kimura AK, Brenner DE et al (1989) A phase II study of the efficacy of diamminedichloroplatinum (cisplatin) for the control of locally recurrent and in-transit malignant melanoma of the extremities using tourniquet outflow-occlusion techniques. Ann Surg 209:73–80

848 Bonenkamp JJ, Thompson JF, de Wilt JH et al (2004) Isolated limb infusion with fotemustine after dacarbazine chemosensitisation for inoperable locoregional melanoma recurrence. Eur J Surg Oncol 30:1107–1112 Brady MS, Brown K, Patel A et al (2006) A phase II trial of isolated limb infusion with melphalan and dactinomycin for regional melanoma and soft tissue sarcoma of the extremity. Ann Surg Oncol 13:1123–1129 Brady MS, Brown K, Patel A et al (2009) Isolated limb infusion with melphalan and dactinomycin for regional melanoma and soft-tissue sarcoma of the extremity: final report of a phase II clinical trial. Melanoma Res 19:106–111 Brown MC, Holl EK, Boczkowski D et al (2017) Cancer immunotherapy with recombinant poliovirus induces IFN-dominant activation of dendritic cells and tumor antigen-specific CTLs. Sci Transl Med 9(408). http://stm.sciencemag.org/content/9/408/ eaan4220.short Callahan MK, Kluger H, Postow MA et al (2018) Nivolumab plus ipilimumab in patients with advanced melanoma: updated survival, response, and safety data in a phase I dose-escalation study. J Clin Oncol 36:391–398 Cassumbhoy R, Pitman AG (2007) Isolated limb infusion for local control of lower limb melanoma: radiologic aspects. Australas Radiol 51:543–549 Cavaliere R, Ciocatto EC, Giovanella BC et al (1967) Selective heat sensitivity of cancer cells. Biochemical and clinical studies. Cancer 20:1351–1381 Chai CY, Deneve JL, Beasley GM et al (2012) A multiinstitutional experience of repeat regional chemotherapy for recurrent melanoma of extremities. Ann Surg Oncol 19:1637–1643 Chang SY, Alberts DS, Farquhar D et al (1978) Hydrolysis and protein binding of melphalan. J Pharm Sci 67: 682–684 Chapman PB, Hauschild A, Robert C et al (2011) Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N Engl J Med 364: 2507–2516 Cheng TY, Grubbs E, Abdul-Wahab O et al (2003) Marked variability of melphalan plasma drug levels during regional hyperthermic isolated limb perfusion. Am J Surg 186:460–467 Clark J, Grabs AJ, Parsons PG et al (1994) Melphalan uptake, hyperthermic synergism and drug resistance in a human cell culture model for the isolated limb perfusion of melanoma. Melanoma Res 4:365–370 Common Terminology Criteria for Adverse Events v3.0 (CTCAE) (2006) https://ctep.cancer.gov/protocolDeve lopment/electronic_applications/docs/ctcaev3.pdf. Accessed 22 Aug 2018 Cornett WR, McCall LM, Petersen RP et al (2006) Randomized multicenter trial of hyperthermic isolated limb perfusion with melphalan alone compared with melphalan plus tumor necrosis factor: American College of Surgeons Oncology Group Trial Z0020. J Clin Oncol 24:4196–4201

G. M. Beasley et al. Coventry BJ, Kroon HM, Giles MH et al (2014) Australian multi-center experience outside of the Sydney Melanoma Unit of isolated limb infusion chemotherapy for melanoma. J Surg Oncol 109:780–785 Damian DL, Barnetson RS, Rose BR et al (2001) Treatment of refractory hand warts by isolated limb infusion with melphalan and actinomycin D. Australas J Dermatol 42:106–109 Damian DL, Barnetson RS, Thompson JF (2006) Treatment of refractory chromomycosis by isolated limb infusion with melphalan and actinomycin D. J Cutan Med Surg 10:48–51 Defty CL, Marsden JR (2012) Melphalan in regional chemotherapy for locally recurrent metastatic melanoma. Curr Top Med Chem 12:53–60 Dossett LA, Ben-Shabat I, Olofsson Bagge R et al (2016) Clinical response and regional toxicity following isolated limb infusion compared with isolated limb perfusion for in-transit melanoma. Ann Surg Oncol 23:2330–2335 Duprat Neto JP, Mauro AC, Molina AS et al (2014) Isolated limb infusion with hyperthermia and chemotherapy for advanced limb malignancy: factors influencing toxicity. ANZ J Surg 84:677–682 Eisenhauer E, Therasse P, Bogaerts J et al (2009) New response evaluation criteria in solid tumours: revised RECIST guideline (version 1.1). Eur J Cancer 45: 228–247 Elhassadi E, Egan E, O’Sullivan G et al (2006) Isolated limb infusion with cytotoxic agent for treatment of localized refractory cutaneous T-cell lymphoma. Clin Lab Haematol 28:279–281 Fletcher WS, Pommier R, Small K (1994) Results of cisplatin hyperthermic isolation perfusion for stage IIIA and IIIAB extremity melanoma. Melanoma Res 4(Suppl 1):17–19 Franklin C, Livingstone E, Roesch A et al (2017) Immunotherapy in melanoma: recent advances and future directions. Eur J Surg Oncol 43:604–661 Gershenwald JE, Scolyer RA, Hess KR et al (2017) Melanoma staging: evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67:472–492 Grob JJ, Amonkar MM, Karaszewska B et al (2015) Comparison of dabrafenib and trametinib combination therapy with vemurafenib monotherapy on health-related quality of life in patients with unresectable or metastatic cutaneous BRAF Val600-mutation-positive melanoma (COMBI-v): results of a phase 3, open-label, randomised trial. Lancet Oncol 16:1389–1398 Grünhagen DJ, Kroon HM, Verhoef C (2015) Perfusion and infusion for melanoma in-transit metastases in the era of effective systemic therapy. Am Soc Clin Oncol Educ Book 35:e528–e534 Hansson J, Lewensohn R, Ringborg U et al (1987) Formation and removal of DNA cross-links induced by melphalan and nitrogen mustard in relation to drug-induced cytotoxicity in human melanoma cells. Cancer Res 47:2631–2637

Isolated Limb Infusion for Melanoma Hauschild A, Grob JJ, Demidov LVet al (2012) Dabrafenib in BRAF-mutated metastatic melanoma: a multicentre, open-label, phase 3 randomised controlled trial. Lancet 380:358–365 Howie LJ, Tyler DS, Salama AK (2015) Neoadjuvant use of ipilimumab in locally advanced melanoma. J Surg Oncol 112:841–843 Huismans AM, Kroon HM, Kam PC et al (2011) Does increased experience with isolated limb infusion for advanced limb melanoma influence outcome? A comparison of two treatment periods at a single institution. Ann Surg Oncol 18:1877–1883 Huismans AM, Kroon HM, Kam PC et al (2016) Isolated limb infusion. In: Aigner KR, Stephens FO (eds) Regional therapy of malignant tumors, 2nd edn. Springer, Heidelberg/Berlin, pp 375–390 Jiang BS, Speicher PJ, Thomas S et al (2015) Quality of life after isolated limb infusion for in-transit melanoma of the extremity. Ann Surg Oncol 22:1694–1700 Karakousis CP, Kanter PM, Lopez R et al (1979) Modes of regional chemotherapy. J Surg Res 26:134–141 Karakousis CP, Kanter DM, Park PC et al (1982) Tourniquet infusion versus hyperthermic perfusion. Cancer 49:850–858 Katsarelias D, Rådbo E, Ben-Shabat I et al (2018) The effect of temperature and perfusion time on response, toxicity, and survival in patients with in-transit melanoma metastases treated with isolated limb perfusion. Ann Surg Oncol. https://doi.org/10.1245/s104340186 4599 Klaase JM, Kroon BBR, Beijnen JH et al (1994) Melphalan tissue concentrations in patients treated with regional isolated perfusion for melanoma of the lower limb. Br J Cancer 70:151–153 Koops HS, Lejeune KJ, Kroon BBR et al (2004) Isolated limb perfusion for melanoma: technical aspects. In: Thompson JF, Morton DL, Kroon BBR (eds) Textbook of melanoma, 1st edn. Martin Dunitz, London, pp 404–409 Kroon BB (1988) Regional isolation perfusion in melanoma of the limbs; accomplishments, unsolved problems, future. Eur J Surg Oncol 14:101–110 Kroon HM, Moncrieff M, Kam PC et al (2008) Outcomes following isolated limb infusion for melanoma. A 14-year experience. Ann Surg Oncol 15:3003–3013 Kroon HM, Lin DY, Kam PC et al (2009a) Efficacy of repeat isolated limb infusion with melphalan and actinomycin D for recurrent melanoma. Cancer 115: 1932–1940 Kroon HM, Lin DY, Kam PC et al (2009b) Isolated limb infusion as palliative treatment for advanced limb disease in patients with AJCC stage IV melanoma. Ann Surg Oncol 16:1193–1201 Kroon HM, Moncrieff M, Kam PC et al (2009c) Factors predictive of acute regional toxicity after isolated limb infusion with melphalan and actinomycin D in melanoma patients. Ann Surg Oncol 16:1184–1192 Kroon HM, Huismans AM, Kam PC et al (2014a) Isolated limb infusion with melphalan and actinomycin

849 D for melanoma: a systematic review. J Surg Oncol 109:348–351 Kroon HM, Huismans A, Waugh RC et al (2014b) Isolated limb infusion: technical aspects. J Surg Oncol 109:352–356 Kroon HM, Coventry BJ, Giles MH et al (2016) Australian multicentre study of isolated limb infusion for melanoma. Ann Surg Oncol 23:1096–1103 Kroon HM, Coventry BJ, Giles MH et al (2017) Safety and efficacy of isolated limb infusion chemotherapy for advanced locoregional melanoma in elderly patients: an Australian multicentre study. Ann Surg Oncol 24:3245–3251 Lidsky ME, Speicher PJ, Jiang B et al (2014) Isolated limb infusion as a model to test new agents to treat metastatic melanoma. J Surg Oncol 109:357–365 Lindner P, Doubrovsky A, Kam PC et al (2002) Prognostic factors after isolated limb infusion with cytotoxic agents for melanoma. Ann Surg Oncol 9:127–136 Lindner P, Thompson JF, De Wilt JH et al (2004) Double isolated limb infusion with cytotoxic agents for recurrent and metastatic limb melanoma. Eur J Surg Oncol 30:433–439 Long GV, Hauschild A, Santinami M et al (2017) Adjuvant dabrafenib plus trametinib in stage III BRAF-mutated melanoma. N Engl J Med 377:1813–1823 Marsden J, Samarasinghe V, Duddy M et al (2008) Regional chemotherapy for inoperable limb cancer using isolated limb infusion. Br J Dermatol 159:10 McArthur GA, Chapman PB, Robert C et al (2014) Safety and efficacy of vemurafenib in BRAF(V600E) and BRAF(V600K) mutation-positive melanoma (BRIM3): extended follow-up of a phase 3, randomised, openlabel study. Lancet Oncol 15:323–332 McDermott DF, Sosman JA, Gonzalez R et al (2008) Double-blind randomized phase II study of the combination of sorafenib and dacarbazine in patients with advanced melanoma: a report from the 11715 Study Group. J Clin Oncol 26:2178–2185 McMahon N, Cheng TY, Beasley GM et al (2009) Optimizing melphalan pharmacokinetics in regional melanoma therapy: does correcting for ideal body weight alter regional response or toxicity? Ann Surg Oncol 16:953–961 Mian R, Henderson MA, Speakman D et al (2001) Isolated limb infusion for melanoma: a simple alternative to isolated limb perfusion. Can J Surg 44:189–192 Moreno-Ramirez D, de la Cruz-Merino L, Ferrandiz L et al (2010) Isolated limb perfusion for malignant melanoma: systematic review on effectiveness and safety. Oncologist 15:416–427 Muilenburg DJ, Beasley GM, Thompson ZJ et al (2015) Burden of disease predicts response to isolated limb infusion with melphalan and actinomycin D in melanoma. Ann Surg Oncol 22:482–488 Mullinax JE, Kroon HM, Thompson JF et al (2017) Isolated limb infusion as a limb salvage strategy for locally advanced extremity sarcoma. J Am Coll Surg 224:635–642

850 Nieweg OE, Kroon BB (2014) Isolated limb perfusion with melphalan for melanoma. J Surg Oncol 109:332–337 Noorda EM, Vrouenraets BC, Nieweg OE et al (2006) Repeat isolated limb perfusion with TNFalpha and melphalan for recurrent limb melanoma after failure of previous perfusion. Eur J Surg Oncol 32:318–324 O’Donoghue C, Perez MC, Mullinax JE et al (2017) Isolated limb infusion: a single-center experience with over 200 infusions. Ann Surg Oncol 24:3842–3849 Parsons PG, Carter FB, Morrison L et al (1981) Mechanism of melphalan resistance developed in vitro in human melanoma cells. Cancer Res 41:1525–1534 Raigani S, Cohen S, Boland GM (2017) The role of surgery for melanoma in an era of effective systemic therapy. Curr Oncol Rep 19:17 Raymond AK, Beasley GM, Broadwater G et al (2011) Current trends in regional therapy for melanoma: lessons learned from 225 regional chemotherapy treatments between 1995 and 2010 at a single institution. J Am Coll Surg 213:306–316 Roberts MS, Wu ZY, Siebert GA et al (2001a) Saturable dose-response relationships for melphalan in melanoma treatment by isolated limb infusion in the nude rat. Melanoma Res 11:611–618 Roberts MS, Wu ZY, Siebert GA et al (2001b) Pharmacokinetics and pharmacodynamics of melphalan in isolated limb infusion for recurrent localized limb malignancy. Melanoma Res 11:423–431 Sanki A, Kam PC, Thompson JF (2007) Long-term results of hyperthermic, isolated limb perfusion for melanoma: a reflection of tumor biology. Ann Surg 245:591–596 Santillan AA, Delman KA, Beasley GM et al (2009) Predictive factors of regional toxicity and serum creatine phosphokinase levels after isolated limb infusion for melanoma: a multi-institutional analysis. Ann Surg Oncol 16:2570–2578 Scott RN, Blackie R, Kerr DJ et al (1990) Melphalan in isolated limb perfusion for malignant melanoma, bolus or divided dose, tissue levels, the pH effect. In: Jakesz R, Rainer H (eds) Progress in regional cancer therapy. Springer, Berlin/Heidelberg, pp 195–200 Siemann DW, Chapman M, Beikirch A (1991) Effects of oxygenation and pH on tumor cell response to alkylating chemotherapy. Int J Radiat Oncol Biol Phys 20:287–289 Skarsgard LD, Skwarchuk MW, Vinczan A et al (1995) The cytotoxicity of melphalan and its relationship to pH, hypoxia and drug uptake. Anticancer Res 15: 219–223 Stehlin JS Jr (1969) Hyperthermic perfusion with chemotherapy for cancers of the extremities. Surg Gynecol Obstet 129:305–308 Thompson JF, Waugh RC, Saw RPM et al (1994a) Isolated limb infusion with melphalan for recurrent limb melanoma: a simple alternative to isolated limb perfusion. Reg Cancer Treat 7:188–192

G. M. Beasley et al. Thompson JF, Lai DT, Ingvar C et al (1994b) Maximizing efficacy and minimizing toxicity in isolated limb perfusion for melanoma. Melanoma Res 4(Suppl 1):45–50 Thompson JF, Siebert GA, Anissimov YG et al (2001) Microdialysis and response during regional chemotherapy by isolated limb infusion of melphalan for limb malignancies. Br J Cancer 85:157–165 Thompson JF, Agarwala SS, Smithers BM et al (2015) Phase 2 study of intralesional PV-10 in refractory metastatic melanoma. Ann Surg Oncol 22:2135–2142 Turaga KK, Beasley GM, Kane JM 3rd et al (2011) Limb preservation with isolated limb infusion for locally advanced nonmelanoma cutaneous and soft-tissue malignant neoplasms. Arch Surg 146:870–875 Ueno T, Ko SH, Grubbs E et al (2004) Temozolomide is a novel regional infusion agent for the treatment of advanced extremity melanoma. Am J Surg 188: 532–537 Van der Merwe SA, Van den Berg AP, Kroon BB et al (1993) Modification of human tumour and normal tissue pH during hyperthermic and normothermic antiblastic regional isolation perfusion for malignant melanoma: a pilot study. Int J Hyperthermia 9:205–217 Vrouenraets BC, Klaase JM, Nieweg OE et al (1998) Toxicity and morbidity of isolated limb perfusion. Semin Surg Oncol 14:224–231 Weber J, Mandala M, Del Vecchio M et al (2017) Adjuvant nivolumab versus ipilimumab in resected stage III or IV melanoma. N Engl J Med 377:1824–1835 Wieberdink J, Benckhuysen C, Braat RP et al (1982) Dosimetry in isolation perfusion of the limbs by assessment of perfused tissue volume and grading of toxic tissue reactions. Eur J Cancer Clin Oncol 18:905–910 Wong J, Chen YA, Fisher KJ et al (2013) Isolated limb infusion in a series of over 100 infusions: a singlecenter experience. Ann Surg Oncol 20:1121–1127 Wong J, Chen YA, Fisher KJ et al (2014) Resection of residual disease after isolated limb infusion (ILI) is equivalent to a complete response after ILI-alone in advanced extremity melanoma. Ann Surg Oncol 21:650–655 World Health Organization (1979) WHO handbook for reporting results of cancer treatments. WHO offset publication no. 48. World Health Organization, Geneva Wu ZY, Thompson MJ, Roberts MS et al (1995) Highperformance liquid chromatographic assay for the measurement of melphalan and its hydrolysis products in perfusate and plasma and melphalan in tissues from human and rat isolated limb perfusions. J Chromatogr B Biomed Appl 673:267–279 Wu ZY, Smithers BM, Anderson C et al (2000) Can tissue drug concentrations be monitored by microdialysis during or after isolated limb perfusion for melanoma treatment? Melanoma Res 10:47–54

Surveillance and Follow-Up of Melanoma Patients Rachael L. Morton, Anne Brecht Francken, and Mbathio Dieng

Contents Goals of Surveillance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 852 Patterns of Melanoma Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Risk of Local and Regional Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Regional Relapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Time to Recurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

852 852 853 854

Strategies for Active Follow-Up of Melanoma Patients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 854 Detection of Recurrences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Role of Physical Examination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Patient Education . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Patient Well-Being and Follow-Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Follow-Up Schedules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Radiologic Studies and Laboratory Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

855 855 856 856 857 857

Screening for Risk of New Primary Melanomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 860 Screening for Other Primary Cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861 Current Recommendations for Surveillance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 862

Abstract

The patient treated for melanoma lives with some risk of recurrence and needs a rational

R. L. Morton (*) · M. Dieng NHMRC Clinical Trials Centre, The University of Sydney, Camperdown, NSW, Australia e-mail: [email protected]; Rachael. [email protected]; [email protected]. au A. B. Francken Centre of Oncology, Isala, Zwolle, Netherlands e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_28

plan for posttreatment follow-up. Appropriate follow-up for melanoma survivors must balance the benefits and harms of repeated surveillance with the needs and goals of the patient and within the capacity and constraints of the healthcare system in which the followup occurs. Routine surveillance by a clinician is an important aspect of follow-up for the detection of melanoma recurrence. For decades, attempts have been made to introduce a follow-up schedule that would find international consensus, but to date, there is no consensus on the optimal 851

852

frequency, intensity, and duration of follow-up in melanoma patients. Dermatologic surveillance is important for detecting new primary and other nonmelanoma skin cancers in this group of patients, a group that frequently has a history of sun damage. A whole-body examination by an experienced dermatologist using technical instruments such as dermoscopy remains the gold standard. During routine follow-up by a healthcare professional, in addition to dermatological examination, a range of strategies including radiologic and laboratory tests are often used. However, there is little evidence to support the value of routine radiologic and laboratory testing particularly in the follow-up of stage I/II melanoma patients. In addition to visits with a healthcare provider, skin self-examination is recommended to identify primary cutaneous lesions but also for detection of recurrences in the locoregional area. Follow-up clinics and guidelines in melanoma care exist around the world. However, the evidence to date is insufficient to strongly support one specific set of follow-up guidelines. This chapter outlines the principles and components of follow-up care.

Goals of Surveillance Melanoma is a significant public health problem worldwide. The incidence of melanoma has increased rapidly in the last two decades, while the death rate has stabilized or fallen in many countries resulting in large numbers of patients surviving melanoma. In the United States in 2018, there will be an estimated 91,270 new cases of invasive melanoma and 9320 deaths attributable to melanoma (The National Cancer Institute 2018). The overall 5-year survival is 90–92% (Australian Institute for Health and Welfare 2018; The National Cancer Institute 2018), thereby dramatically increasing the prevalence of melanoma survivors in need of follow-up. Individuals with regional or distant disease account for approximately 13–15% of patients with newly diagnosed melanoma and are generally thought to warrant close surveillance. Fortunately, today most new

R. L. Morton et al.

cases of invasive melanoma are diagnosed at an early stage, approximately 85% (The National Cancer Institute 2018) of cases are either stage I or stage II, and most of those are expected to have a good prognosis after surgery alone (Jemal et al. 2008; Balch et al. 2009). Nonetheless, every patient treated for melanoma lives with some risk of recurrence and needs a rational plan for follow-up. Appropriate follow-up for melanoma survivors must balance the benefits and harms of repeated surveillance and the needs and goals of the patient, within the capacity and constraints of the healthcare system in which the follow-up occurs. After the melanoma treatment is completed, active surveillance can address many goals. The goal most often discussed and pursued is surveillance for the risk of disease recurrence, whether local, regional, or distant. Follow-up is most valuable when it leads to detection of recurrent disease that is potentially curable. Given the success of systemic therapies for treating active stage III and IV melanoma, early detection and treatment of recurrence may further improve survival. Another goal of surveillance is the detection and management of long-term complications of treatment, including lymphedema related to surgical management or side effects from systemic drug therapy. Other appropriate goals for melanoma patients who have completed treatment include assessing quality of life and patient satisfaction with care; providing psychosocial support, providing education regarding melanoma prevention and skin self-examination, and ongoing screening for the risk of new primary skin cancers and screening for the risk of other primary cancers (Morton et al. 2013; Read et al. 2018; Rychetnik et al. 2013). Melanoma clinical quality registries may play a role in long-term data collection for follow-up programs (Jochems et al. 2017).

Patterns of Melanoma Recurrence Risk of Local and Regional Recurrence The overall risk of melanoma recurrence is most strongly correlated with stage at presentation (Balch et al. 2009). Between 20% and 25% of

Surveillance and Follow-Up of Melanoma Patients

patients with (American Joint Committee on Cancer) stage I and II melanoma in the sentinel lymph node biopsy era will have a recurrence (Turner et al. 2011). The pattern of recurrence, that is, whether a first relapse is most likely to be local, regional, or distant, is related to both the stage of the original disease and the type of treatment given. A number of randomized clinical trials comparing wider versus narrower excision margins have established the current recommendations for excision margins and have also provided documentation of local recurrence rates after such (Balch et al. 2001; Cohn-Cedermark et al. 2000; Khayat et al. 2003; Thomas et al. 2004; Veronesi et al. 1991). The World Health Organization trial of 1 cm versus 3 cm margins for melanomas less than 2 mm reported local recurrence rates of less than 1% regardless of excision margin (Veronesi et al. 1991). A Swedish trial of 2 cm versus 5 cm margins for melanomas 0.8–2.0 mm reported local recurrence rates of less than 1% with no difference between the two groups; a similar European study of 2 cm versus 5 cm margins for melanomas less than 2.1 mm reported local recurrence rates of 1.5% with no difference between groups (CohnCedermark et al. 2000; Khayat et al. 2003; Veronesi et al. 1991). In the Intergroup Melanoma Surgical Trial comparing 2 cm versus 4 cm margins for intermediate-thickness (I to 4 mm) melanomas, local recurrence rates were 2.1–2.6% at a median 10year follow-up but were not correlated with excision margin. Higher local recurrence rates were seen in patients with ulcerated tumors and primary tumors of the head and neck (Balch et al. 2001). In a study from the United Kingdom of patients with melanomas greater than 2 mm, who were randomized to groups with 1 cm versus 3 cm margins, local recurrence as a first event was seen in 3.1% of patients, and the rates did not differ significantly between the two groups (Thomas et al. 2004). Factors in many of these studies that appear to be associated with higher risk of local recurrence were increased tumor thickness and tumor ulceration. More recently the MelMarT trial (Moncrieff et al. 2018) was designed to ascertain the impact

853

of 1 cm- vs 2 cm-wide excision for melanomas >1 mm Breslow thickness, on local recurrence, and melanoma-specific survival.

Regional Relapse Prior to the widespread implementation of sentinel node biopsy, regional lymph nodes were the most common initial site of recurrent disease. In the abovementioned UK study, the nodal relapse rate for stage II patients with melanomas greater than 2 mm was 29.2% at 5-year follow-up (Thomas et al. 2004). The risk of regional nodal recurrence in patients managed with wide local excision and nodal observation increases with increasing tumor thickness. The risk of nodal relapse in patients randomized to the observation arm of the Multicenter Selective Lymphadenectomy Trial I (with thicknesses ranging from 1.2 to 3.5 mm) was 15.6% at 5 years (Morton 2006). Since the adoption of sentinel node biopsy for patients who are newly diagnosed with melanoma, the regional node basin is no longer the most common initial site of clinical relapse. Inbasin recurrence among patients who have had a negative sentinel node biopsy is uncommon, with rates of 1.5–4.1% reported (Chao et al. 2002; Dalal et al. 2007; Gershenwald et al. 1998; Morton et al. 2006; Vuylsteke et al. 2003). The Multicenter Selective Lymphadenectomy Trial II (Faries et al. 2017) that randomized sentinel node biopsy-positive patients to complete lymph node dissection or nodal observation reported 92% disease control in regional lymph nodes at 3 years versus 77% control for the observed patients. In a series of patients who have undergone sentinel node biopsy as a staging procedure with complete nodal dissection for positive sentinel nodes, the regional node basin was a subsequent site of disease recurrence in 0–10% of patients (Chao et al. 2002; Gershenwald et al. 2000; Vuylsteke et al. 2003). Recurrence patterns for patients who have had a sentinel node biopsy (with selective node dissection in the event that the sentinel node is positive) now show a predominance of distant and local recurrences, with far fewer regional

854

recurrences (Chao et al. 2002; Gershenwald et al. 2000; Vuylsteke et al. 2003). In-transit metastases appear as the initial site of relapse in 2–8% of patients after wide local excision of their melanomas (Borgstein et al. 1999; Dicker et al. 1999; Fusi et al. 1993). In a large Australian study of 3642 patients undergoing sentinel node biopsy (Read et al. 2015), the rate of in-transit metastases was 7.2%. The rate was 4.7% in SN-negative patients and 21.6% in SN-positive patients. Factors that predispose patients to in-transit recurrence include increasing tumor thickness of the primary lesion, tumor ulceration, and sentinel node positivity (read et al. 2015).

Time to Recurrence In patients who have recurrences, 55–67% of these recurrences will become apparent by 2 years and 65–81% by 3 years after initial treatment of the primary tumor (Balch et al. 2009; Francken et al. 2008a). Tumor ulceration, older patient age, and site of primary melanoma are also associated with earlier recurrence. The time to recurrence varies in inverse order to the thickness of the primary tumor, that is, thicker tumors have a recurrence pattern that favors early recurrence, whereas thinner tumors may recur many years after diagnosis (Lo et al. 2018). In patients treated for melanoma that has metastasized to regional nodes, the time to distant recurrence is usually earlier than in patients without regional nodal metastases. The heterogeneity of prognosis in patients with metastatic regional nodes is striking. The three most important factors contributing to prognosis of node-positive patients are the number of positive nodes, whether the nodes were clinically palpable and whether the primary tumor showed ulceration on histologic examination. Five-year survival rates range from 69% in patients with a single microscopically positive node and a non-ulcerated primary lesion to 13% in patients with four or more clinically positive nodes and an ulcerated primary lesion. The 15year survival curves for patients with positive nodes are notable for two features. Patients with

R. L. Morton et al.

a greater number of positive nodes have distinctly poorer survival rates than those with fewer positive nodes. In addition, most deaths due to melanoma occur in the first 3 years, with the survival curves all gradually reaching a plateau (Balch et al. 2009). At 15 years, a persistent minority of patients are alive, even with a history of four or more positive nodes (Balch et al. 2009). The possibility of late recurrence (more than 10 years after initial treatment) in patients with melanoma has been well described (Faries et al. 2013; Green et al. 2012). In general, patients with late recurrence tended to be younger and have thinner primary lesions than patients with an earlier recurrence, and they almost always initially have clinically negative nodes. In general, the prognosis after late recurrence is similar to that after early recurrence (Faries et al. 2013).

Strategies for Active Follow-Up of Melanoma Patients There is no general consensus on the optimal frequency, intensity, and duration of follow-up in melanoma patients. Most recurrences are detected within the first 2–3 years after diagnosis, with decreasing likelihood of recurrence over time. The probability that a recurrence will develop more than 5 years after diagnosis is low but is clearly not zero. Surveillance for locoregional and distant recurrence seems most likely to be productive in the first few years after diagnosis, with diminishing yield expected with longer-term follow-up. It is difficult to define the relationship between intensity of follow-up and outcome in patients with melanoma. The assumption motivating the followup of any cancer patient is that early detection of recurrence will lead to early treatment, which might have an impact on long-term outcomes (Rueth et al. 2015). However, recurrences, when detected, are not always treatable and, if they are treatable, are not always curable. To date, there are no studies of follow-up of melanoma patients that have demonstrated any survival advantage with intensive monitoring for

Surveillance and Follow-Up of Melanoma Patients

recurrence (Nieweg and Kroon 2006). It is unclear whether surveillance of treated melanoma patients results in earlier curable detection and superior outcomes.

Detection of Recurrences Patient history and physical examination remain the cornerstone of recurrence detection in patients in all stages of melanoma, since local, in-transit, and regional recurrences are most frequently detected by physical examination. Distribution of recurrences was described as 3–5% local or in transit, 5–13% regional, and 3–10% distant (Francken et al. 2005). Among patients whose melanoma recurred, 20–28% first presented with local or in transit, 26–60% with regional nodal, and 15–50% with distant recurrence. A number of retrospective studies have assessed the detection of first recurrences. Patient-detected recurrences were reported in 44–90% of cases, but a variety of methodologies were used (questionnaires, interviews, and including symptomatic versus asymptomatic patients). Almost all studies described a prescribed follow-up schedule that patients adhered to. The minimum frequency of followup was twice a year for the first 3 years in all studies. Patient detection of a first recurrence has consistently been the most commonly reported type of recurrent disease detection (Baughan et al. 1993; Francken et al. 2007; Meyers et al. 2009; Mooney et al. 1997; Romano et al. 2010). To date, only one prospective study has assessed follow-up of patients with melanoma, with a focus on how recurrences were detected at follow-up visits (Garbe et al. 2003). Only 17% of the 233 recurrences in 112 patients were detected by the patient. Although methodological differences could explain this difference with retrospective studies, a large international cohort study from 8178 patients found skin self-examination to be higher in Australia, the United States, and Southern Europe than in Central Europe (Kasparian et al. 2012). This was confirmed by Livingstone et al. in a German study where 33.3% of recurrences were patient detected, even though 69.4% of recurrences were visible or palpable

855

(Livingstone et al. 2015). We can conclude that there may be large topographical and cultural differences and that there is an urgent need for better patient education in some areas, to guarantee that confirmation of metastasis occurs as early as possible. In this same study, only 12.9% of the patients reported to have received patient education about self-surveillance, 68.1% oral information, and 51.7% written information, respectively. A few studies analyzed survival according to the person who detected the first recurrence, but most did not find a significant difference in survival regardless of whether the recurrence was detected by the patient or a doctor (Baughan et al. 1993; Francken et al. 2007; Hoffman 2002). Garbe et al. however did find a survival benefit in early detected metastases (Garbe et al. 2003).

Role of Physical Examination As described above, comprehensive locoregional examination is the most important facet of melanoma follow-up. Most recurrences after treatment of stage I and II melanoma are locoregional and thus detectable by physical examination. Early detection of locoregional recurrence along with surgical intervention may provide better control of disease. A careful examination for local recurrence or visible or palpable in-transit disease, as well as examination of regional nodes, is essential. Up to 30% of patients with locoregional recurrences can be salvaged with surgical treatment. Basseres et al. reported that of 115 recurrences detected in 528 patients with stage I melanoma, 87% were detected by clinical examination (Basseres et al. 1995). However, even in a study where stage III patients were followed after primary diagnosis and received CT scans every 3 months the first 2 years and every 6 months up to 5-year follow-up, 47% of recurrences were detected by the patients, 21% by the physician, and only 32% by imaging methods (Romano et al. 2010). Physical examination of patients with resected regional disease is useful as well. A study from Gadd et al. reported a series of 1019 patients undergoing lymphadenectomy for

856

melanoma, 403 of whom had a recurrence. Of those 403 patients, 291 (72%) had a recurrence at a single site. Of these 291 single-site recurrences, 190 (65%) were non-visceral and potentially detectable by physical examination (Gadd and Coit 1992). With regard to the role of physical examination in detection of distant metastatic disease, in patients who have a recurrence with remote nodes or soft tissue metastases, physical examination represents an efficient, low-cost method of detection. It is particularly important to detect metastases, because there is a 20–25% survival rate after surgical resection alone in this group of patients (Faries and Morton 2006). Improved disease-free survival in patients with extensive unresectable soft tissue metastases may be achieved with the use of systemic therapies. Visceral metastases are not frequently detected by physical examination. Therefore for high-risk patients, screening with imaging techniques may be required.

R. L. Morton et al.

recurrences, while in other countries this can be as low as 20% (Francken et al. 2008a, b; Garbe et al. 2003; Poo-Hwu et al. 1999; Hofmann et al. 2002). These data highlight that even with education, there are great differences in patients’ individual ability to detect recurrences. A history of melanoma diagnosis by itself does not seem to improve patients’ ability to detect new lesions by themselves (Francken et al. 2008b). Self-examination is recommended following definitive local treatment for melanoma patients of any stage. The Australian clinical practice guidelines (Cancer Council Australia 2018) further recommend that high-risk patients be educated to recognize and document lesions suggestive of melanoma and to perform the skin self-examination. This surveillance for at high-risk people aims to detect recurrence or progression at an early stage, identify treatment-related morbidity, identify new melanomas or non-melanoma skin cancers, and provide support.

Patient Education Patient Well-Being and Follow-Up As a corollary to the preceding observations, an integral part of the follow-up of patients with melanoma could be the teaching of skin selfexamination, not only for detection of new primary cutaneous lesions but also for detection of recurrences in the locoregional area. Successfully implementing skin self-examination requires patient education on whole-body skin examination with particular attention given to melanoma surgical scars and the corresponding lymphatic drainage areas for in-transit and lymph node recurrence. Patients should also be given education regarding persistent symptoms that may warrant further investigation. In addition, the use of brochures or videos and the engagement of relatives in the education process may be helpful (Mills et al. 2017; Murchie et al. 2007, 2015; Francken et al. 2005, 2008b; Dancey et al. 2005; Poo-Hwu et al. 1999). Randomized controlled trials are underway (ACTRN12616001716459: https://www.anzctr.org. au/Trial/Registration/TrialReview.aspx?id=371865). In Australia, patients themselves detect up to 75% of

Improved methods for melanoma detection and treatment have led to increasing numbers of people surviving melanoma for many years. Thus, cancer survivorship should be a primary focus, and there is a recognized need for more research to increase the well-being of cancer survivors. For decades, physical outcomes such as recurrence and survival have been used to evaluate treatment effects in people with melanoma. Researchers have neglected to examine how routine follow-up can impact patients’ well-being. Few studies have investigated patient well-being in patients with melanoma and hardly any in relation to follow-up surveillance. Most studies report that follow-up is much appreciated by patients; although it evokes anxiety, consultation with doctors, psychologists and nurses can help patients to cope with uncertainties (Bell et al. 2017; Lim et al. 2018; Morton et al. 2013; Brandberg et al. 1995; Wheeler 2006). Studies of psychological and educational interventions have been shown to result in lower levels

Surveillance and Follow-Up of Melanoma Patients

of distress, at least in short-term follow-up (Dieng et al. 2016; Boesen et al. 2005; Trask et al. 2003).

Follow-Up Schedules For decades attempts have been made to introduce a follow-up schedule that would find international consensus. Most follow-up schedules are designed in relation to the expected risk of recurrence, with more intensive surveillance in the first years after diagnosis and treatment, and many are based on the expected pattern of recurrences (Baughan et al. 1993; Dicker et al. 1999; Garbe et al. 2003; Hofmann et al. 2002; Poo-Hwu et al. 1999). All recommend more follow-up visits for thicker tumors and for melanomas of more advanced stage, as well as a reduction in the frequency of visits over time. Trotter et al. summarized recommendations and guidelines on the follow-up care of several organizations over the world. This overview shows that there is still limited consensus. Most differences are seen in frequency and duration of follow-up, but also in the recommendations regarding imaging and laboratory tests. Table 1 gives an overview of followup recommendations of leading organizations from different countries in the world. Cromwell et al. studied the variability of follow-up care as well and found most disagreement in recommendations for stage I melanoma, and the duration of follow-up spreads out between one single visit to lifelong (Cromwell et al. 2012). Although most guidelines recommend frequent follow-up, overall adherence seems to be poor. An Austrian study of patients with melanomas less than 1.5 mm thickness found a mean annual dropout rate of 11.2%, with only 55.3% of patients still in follow-up at 5 years. The dropout rate was not related to sex, age, or tumor thickness (Kittler et al. 2001). Similar dropout was seen in Germany (Livingstone et al. 2015), and Australian figures were even worse: 43.2% stage I and 28.7% stage II had a maximum of 1 post-up visit in their specialized clinic. Only 13.2% stage I and 4.1% stage II had received follow-up according to the guideline over 5 years. Nonetheless patients might have found

857

their consultation for follow-up at their GP or elsewhere (Memari et al. 2015). Huibertse et al. studied patients’ preferences in the Netherlands for the healthcare professional providing follow-up care (Huibertse et al. 2017). Most patients preferred a medical specialist, but the authors concluded this might be due to unfamiliarity with other healthcare providers such as specialist nurses or nurse practitioners. In the United Kingdom, a randomized controlled trial was performed on GP-led follow-up. The authors found the strategy was safe and feasible and engendered high patient satisfaction (Murchie et al. 2010). Few studies are done on reducing follow-up schedules. The MELFO study in the Netherlands examined the well-being of patients with stages Ib–IIc melanoma in a randomized controlled trial in which patients received reduced, stage-adjusted follow-up. At 1-year follow-up, no difference in well-being was found nor any adverse results (Damude et al. 2016). The same (MELFO) follow-up schedule was studied in Australia by modelling the delay in diagnosis. Turner et al. (2011) did find a potential delay in diagnosis of about 2 months in only 44.9 per 100 patients. Fields and Coit concluded in their review on the follow-up of melanoma that many authors conclude their work with recommendations to reduce follow-up. In the absence of a randomized trial evidence, patient-related factors such as physical and mental health, travel distance, personal preferences, and available resources determine follow-up schedules, and most follow-up guidelines have remained unchanged (Fields and Coit 2011).

Radiologic Studies and Laboratory Tests There is little evidence to support the value of routine radiologic and laboratory testing in the follow-up of stage I/II melanoma patients. The use of routine chest X-ray exams for the detection of small pulmonary metastases has been investigated; however, false-positive and false-negative findings are frequent. The sensitivity of chest Xray is poor with reports varying from 7.7% to

858

R. L. Morton et al.

Table 1 Melanoma follow-up guidelines

Source

NCCN

Guidelines for Management of Melanoma in Australia and New Zealand

German Cancer Society and German Dermatological Society

Comprehensive Cancer Network of the Netherlands

Year

2013

2018

2013

2016

Basis of follow-up guideline

Stage specific

Stage specific

Stage specific

Stage specific

Follow-up guidelines Stage/ History and Physical Breslow Examination thickness Stage 0 H&P, CXR, CT, PET, MRI, S100 annually for life Stages H&P every 3–12 months for IA–IIA 5 years and then annually as clinically indicated Stages H&P every 6–12 months for IIB–IV 2 years and then 3–12 months for 3 years and then annually as clinically indicated I H&P annually for 10 years IIA History and Physical Examination every 6 months for 2 years, annually for 8 years IIB–IIC H&P every 3 months for 2 years, every 6 months for 2 years, annually for 5 years IIA–IIIC Every 3 months for 2 years, every 6 months for 1 year IA H&P every 6 months for 3 years, every 12 months for 7 years Stages H&P every 3 months for IB–IIB 3 years, every 6 months for 2 years, every 6–12 months for 5 years H&P every 3 months for Stages 5 years and then 6 months IIC–IV, not for 5 years resected

Stage 0 Stage 1A Stages IB–IV

Once at 1 month after surgery Once at 1 month after surgery H&P every 3 months 1st year, every 6 months 2nd year, every yearly 3–5 years

Imaging/laboratory evaluation None Not recommended

Consider CXR, CT+/-PET every 3–12 months and annually MRI of the brain. No imaging in asymptomatic patients after 5 years None None

None

Considered for 3 years, but no survival advantage None

Every 6 months ultrasonography. Every 3 months S100B Every 3 months ultrasonography and S100B for 3 years, every 6 months for 2 years, every 3 months S100B. Cross-sectional imaging every 6 months for 3 years None None No specific recommendations

H&P = history and Physical Examination; CXR = chest radiograph; CT = computed tomography; PET = positron emission tomography; MRI = magnetic resonance imaging; S100 (blood test).

48%. A large study of 1969 patients with stages I–III melanoma undergoing routine follow-up found only 10/204 relapses were discovered by

chest X-ray: the majority (7/10) of which were observed in patients with stage III disease (Leiter et al. 2009).

Surveillance and Follow-Up of Melanoma Patients

A large prospective study of 1235 patients found only 0.9% of chest X-rays identified pulmonary metastases, less than 10% of which were amenable to resection, with a false-positive rate of 3.1% (Brown et al. 2010). A cost-effectiveness analysis using data from the Roswell Park Cancer Institute and the National Cancer Institute’s Surveillance, Epidemiology, and End Results (SEER) program found CXR screening was not cost-effective at $165,000 per quality-adjusted life year gained (Mooney et al. 1997). A recent systematic review of PET imaging studies (Danielsen et al. 2013) was undertaken to assess the diagnostic value of PET as a tool for surveillance in the regular follow-up program of asymptomatic cutaneous malignant melanoma patients. The majority of the 739 patients in the studies were stage IIB and III. The authors repeated the mean sensitivity of PET was 96% (95% CI: 92–98), and the specificity was 92% (95% CI: 87–95). Overall, PET had a high diagnostic value; however, there were no data available to demonstrate better survival outcomes for patients as a result of routine PET surveillance. In addition, PET produces false-positive findings leading to subsequent unnecessary procedures. The usefulness of ultrasonography for followup of patients without a sentinel node biopsy treated for stage I/II melanoma depends entirely on the technical skill and experience of the personnel involved. There is a consensus of opinion that ultrasound is superior to clinical examination of regional lymph nodes, although its survival advantage is unproven (Bafounta et al. 2004). A prospective cohort study of 373 patients with a primary tumor Breslow thickness of 1.5 mm (Machet et al. 2005) reported a sensitivity of 93% for ultrasound compared with only 71% for the clinical examination of regional lymph nodes. The specificity was equally high for both procedures (>98%). Despite the superiority of ultrasound, very few patients actually benefited from the addition of ultrasound to clinical examination, due to deleterious effects such as unnecessary stress caused by repetition of ultrasounds for benign lymph nodes or the useless removal of benign lymph nodes. In summary, ultrasound

859

was advantageous in only 1% of patients. Fine needle aspiration is the current standard method to confirm the presence of suspected nodal metastases for lymphadenopathy identified after definitive local treatment of cutaneous melanoma (Dalle et al. 2006). Ultrasound may be used to identify the extent of in-transit and nodal disease and also to diagnose liver metastases. The current German guidelines, for example, include ultrasound examinations of regional lymph nodes and intransit areas at regular, risk-adapted time intervals (Garbe et al. 2007). Furthermore, reduced intensity and frequency of follow-up recommendations are known to be more cost-effective (Hengge et al. 2007). Some centers have adopted a risk-stratified approach using CT scans of the chest and abdomen exclusively for stage III patients with a high risk of distant metastases and MRI of the brain as cerebral metastases are more readily detected by MRI than by CT. Routine radiological investigations every 3–12 months may be considered for the first 3 years of follow-up after definitive local treatment of stage IIC and III melanoma where detection of recurrence would allow early commencement of systemic therapy. However, there are currently no high-quality data that early detection and treatment of recurrence improves survival. Positron emission tomography (PET) with or without concurrent CT scans (PET-CT) has been met with enthusiasm as promising imaging techniques for high-risk melanoma patients. The specificity of CT scans alone may be improved by adding a concurrent PET scan and by clinical correlation. PET-CT scans are at present best employed in the diagnostic evaluation of patients with established regional or distant disease. The routine use of PET-CT is troubled by possible false-positive results, leading to anxiety, additional investigations, and expense. Even when true distant disease is found, it is unclear whether such detection will be associated with improved survival. However, PET-CT imaging is becoming a standard of care in many melanoma treatment centers (Pflugfelder et al. 2013; The National Cancer Institute 2018), particularly in the era of adjuvant immunotherapies for stage III

860

melanoma. Early detection of advanced metastatic disease and early treatment with systemic therapies may lead to an improvement in overall survival (Rueth et al. 2015). Laboratory evaluations are still performed in melanoma treatment centers and practices specializing in melanoma, although they rarely lead to the detection of metastatic disease. As a tumor marker, S100B displays a sensitivity of 86–91% and specificity of 76–91% (Deichmann et al. 1999; Krahn et al. 2001) and may portend recurrence; however there are no data demonstrating superior survival outcomes for patients undergoing routine S100B testing in follow-up. The use of serum LDH or melanoma inhibitory activity (MIA) protein in follow-up for the detection of asymptomatic melanoma recurrence has been reviewed (Schultz et al. 1990). Abnormal blood tests were rarely the first sign of metastases. Low sensitivity, specificity, and accuracy for general laboratory profiles make them ineffective in the detection of subclinical recurrence, and their roles are yet to be defined. Other investigations during follow-up include skin self-examination, physician-led medical history, and clinical examination. A review of nine clinical practice guidelines (2014) (Marciano et al. 2014) reveals consensus that patients should be taught skin self-examination, as most recurrences are first detected by patients (Francken et al. 2007). In a large prospective study, approximately 50% of recurrences were identified by history taking/physical examination, 80% of which were local recurrences, in-transit metastases, and regional lymph node metastases (Garbe et al. 2003).

Screening for Risk of New Primary Melanomas Patients with a history of melanoma have a higher risk for the development of new primary melanomas compared with the general population. The reported incidence of new primaries ranges from 2% to 8% (Ferrone et al. 2005; Francken et al. 2005; Nieweg and Kroon 2006) and the rate remaining relatively constant over 20 years of follow-up at 6.01 per 1000 person-years

R. L. Morton et al.

indicating a high lifetime risk of second primary invasive melanomas (McCaul et al. 2008). This risk does not diminish over time and does not differ significantly between patients first diagnosed with lentigo maligna, in situ melanoma, or invasive melanoma. The risk is even higher for patients with a parental history of melanoma (Zhang et al. 2008). A second invasive melanoma is most commonly thinner than the initial primary melanoma and has a more favorable prognosis (Jones et al. 2016). However, a large population-based study of 32,238 patients reported a hazard ratio of death within 10 years from melanoma two times higher for those with two melanomas and nearly three times higher when three melanomas were diagnosed, compared with people with a single melanoma (Youlden et al. 2016). In general, subsequent primary invasive melanomas are more likely to occur at the same body site as the initial invasive or in situ melanoma (Youlden et al. 2014). The subset of melanoma patients at very high risk for developing a second primary skin cancer include people with a hereditary predisposition for melanoma (i.e., germline CDKN2A mutations or the CDK4 gene or xeroderma pigmentosum) and those with multiple atypical nevi or a strong family history of melanoma (Ferrone et al. 2005; Hansson et al. 2007). Whole genome sequencing has identified common variations of single-nucleotide polymorphisms (SNPs) in at least 20 genes that influence melanoma risk in the population, accounting for about 20% of the excess risk to relatives of melanoma cases (Cancer Council Australia 2018). Recent clinical practice guidelines suggest clinical genetic testing for CDKN2A mutations, and genetic counselling should be considered in individuals with a strong family history of melanoma (three or more cases related in the first or second degree) where predictive features are present, such as multiple primary melanoma, early age of onset, or pancreatic cancer (Cancer Council Australia 2018). While follow-up guidelines for this population are variable (Watts et al. 2015), clinical surveillance

Surveillance and Follow-Up of Melanoma Patients

in specialized clinics for people at high or very high risk of melanoma has been shown to be effective and cost-effective (Moloney et al. 2014). Dermatologic surveillance is important in detecting new primary and other nonmelanoma skin cancers in this group of patients, a group that frequently has a history of sun damage. Screening for these second melanomas is therefore essential, because early detection of primary melanoma is associated with improved survival. A wholebody examination by an experienced dermatologist in a structured followup is still the gold standard. Apart from the ABCD rule to differentiate between benign nevi and melanomas, patient concerns and symptoms are important in early detection. Other skin screening approaches include total body photography repeated at regular intervals as an adjunct to skin examination. Total body photography is commonly used to monitor patients at increased risk for melanoma, particularly those with high nevus counts and dysplastic nevi (Cancer Council Australia 2018). Comparison with prior photos may allow identification of subtle changes that may lead to biopsy or, conversely, may provide reassurance that no change has occurred and biopsy can be avoided. Recent studies have reported total body photography reduces the biopsy rate of benign nevi and improves diagnostic accuracy of melanoma in high-risk patients (Truong et al. 2016; Moloney et al. 2014). Currently the most useful technical instrument in the follow-up of melanoma is the dermatoscope. Dermoscopy is based on the ability of light to penetrate the epidermis and illuminate the superficial dermis. Many features of the appearance of the dermis under dermoscopy have been described that may help distinguish between benign and malignant lesions. Dermoscopy, with or without computer imaging, is widely used in Europe and Australia particularly in highrisk patients. The proper use of dermoscopy requires a certain amount of focused training and experience (Binder et al. 1997; Argenziano 2005). It can be a useful tool not only in detecting early melanomas but also in preventing the unnecessary removal of benign melanocytic lesions. Clinical guidelines recommend the use of short-

861

term sequential digital dermoscopy imaging (dermoscopy monitoring) to detect melanomas that lack dermoscopic features of melanoma and long-term sequential digital dermoscopy imaging (dermoscopy monitoring) in patients who are at high risk of a new primary melanoma (Cancer Council Australia 2018). Teledermoscopy is a form of teledermatology that specifically involves the store and forwarding of digital dermoscopic images. When compared with other imaging techniques, teledermoscopy improves diagnostic accuracy and has recently been shown to be cost-effective (Snoswell et al. 2018) as a referral mechanism for specialist dermatological review of skin cancers versus a standard referral letter without images. Monitoring of suspicious lesions particularly amelanotic tumors (Guitera et al. 2016) and lesions typified by regression (Borsari et al. 2016) may be best performed using reflective confocal microscopy used by a skilled dermatologist.

Screening for Other Primary Cancers For patients with a history of melanoma, Bradford et al. found significantly elevated risks for specific subsequent primary cancers other than melanoma (Bradford et al. 2010). The most common cancers with elevated risks after an initial melanoma were prostate cancer, female breast cancer, and non-Hodgkin lymphoma. Risks were also found to be significantly increased for cancers of the salivary gland, small intestine, kidney, ocular melanoma, and thyroid as well as soft tissue sarcomas and chronic lymphocytic leukemia (Bradford et al. 2010). In the absence of other medical indications, routine age-appropriate surveillance guidelines should be followed.

Current Recommendations for Surveillance Given the lack of evidence of any improvement in outcomes as a result of intensive surveillance examination after definitive treatment for

862

melanoma, one may wonder whether follow-up guidelines should exist at all (Nieweg and Kroon 2006; Fields and Coit 2011). Regardless, follow-up clinics and guidelines in melanoma care exist around the world. Among other places, follow-up guidelines have been developed in the United Kingdom, Germany, Switzerland, Scotland, the Netherlands, the United States, and Australia. The content of these guidelines varies widely. The duration of follow-up varies from 3 to 10 years total, often depending on the perceived level of risk. Many guidelines do not recommend any routine radiologic or laboratory investigations, except when specific symptoms are present, as seems to be concordant with the evidence. However, some do consider routine or optional chest radiographs, lymph node ultrasound imaging, blood tests, or PET-CT imaging for higher-risk patients. The optimal duration of follow-up is unknown; late recurrences more than 10 years after treatment of localized melanoma are uncommon but well recognized, and the risk of developing a second primary melanoma is spread out over a lifetime. Current recommendations for follow-up meet the goals of surveillance for recurrence of treated melanomas, screening for new primary skin cancers, and monitoring for complications of treatment. The current guidelines in the United States, Australia, Germany, and the Netherlands in patients with stages 0–IV melanoma are outlined in Table 1 (Saiag et al. 2007; Garbe et al. 2007). An optimal follow-up program for patients with melanoma should be defined by the results of a large-scale randomized trial. However in the absence of this evidence, and given the increasing incidence of melanoma, the development of rational, productive, and cost-effective followup plans is of great importance. The evidence in the existing literature is insufficient to strongly support one specific set of follow-up guidelines. The wide variance in guidelines currently in use demonstrates that historical policy, cultural experience, and personal views still play an important role in both the guidelines and the practice of follow-up care in patients with melanoma.

R. L. Morton et al.

References Argenziano G (2005) The impact of dermoscopy on the management of pigmented skin lesions: the role of follow-up. J Am Acad Dermatol 52:178; author reply 179–180 Australian Institute for Health and Welfare (AIHW) (2018) Melanoma skin cancer in Australia [Online]. https:// melanoma.canceraustralia.gov.au/statistics. Accessed Aug 2018 Bafounta ML, Beauchet A, Chagnon S, Saiag P (2004) Ultrasonography or palpation for detection of melanoma nodal invasion: a meta-analysis. Lancet Oncol 5:673–680 Balch CM, Soong SJ, Smith T, Ross MI, Urist MM, Karakousis CP, Temple WJ, Mihm MC, Barnhill RL, Jewell WR, Wanebo HJ, Desmond R, Investigators from the Intergroup Melanoma Surgical, T (2001) Long-term results of a prospective surgical trial comparing 2 cm vs. 4 cm excision margins for 740 patients with 1–4 mm melanomas. Ann Surg Oncol 8:101–108 Balch CM, Gershenwald JE, Soong SJ, Thompson JF, Atkins MB, Byrd DR, Buzaid AC, Cochran AJ, Coit DG, Ding S, Eggermont AM, Flaherty KT, Gimotty PA, Kirkwood JM, Mcmasters KM, Mihm MC Jr, Morton DL, Ross MI, Sober AJ, Sondak VK (2009) Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol 27:6199–6206 Basseres N, Grob JJ, Richard MA, Thirion X, Zarour H, Noe C, Collet-Vilette AM, Lota I, Bonerandi JJ (1995) Cost-effectiveness of surveillance of stage I melanoma. A retrospective appraisal based on a 10-year experience in a dermatology department in France. Dermatology 191:199–203 Baughan CA, Hall VL, Leppard BJ, Perkins PJ (1993) Follow-up in stage I cutaneous malignant melanoma: an audit. Clin Oncol (R Coll Radiol) 5:174–180 Bell KJL, Mehta Y, Turner RM, Morton RL, Dieng M, Saw R, Guitera P, Mccaffery K, Low D, Low C, Jenkins M, Irwig L, Webster AC (2017) Fear of new or recurrent melanoma after treatment for localised melanoma. Psychooncology 26:1784–1791 Binder M, Puespoeck-Schwarz M, Steiner A, Kittler H, Muellner M, Wolff K, Pehamberger H (1997) Epiluminescence microscopy of small pigmented skin lesions: short-term formal training improves the diagnostic performance of dermatologists. J Am Acad Dermatol 36:197–202 Boesen EH, Ross L, Frederiksen K, Thomsen BL, Dahlstrom K, Schmidt G, Naested J, Krag C, Johansen C (2005) Psychoeducational intervention for patients with cutaneous malignant melanoma: a replication study. J Clin Oncol 23:1270–1277 Borgstein PJ, Meijer S, Van Diest PJ (1999) Are locoregional cutaneous metastases in melanoma predictable? Ann Surg Oncol 6:315–321 Borsari S, Pampena R, Lallas A, Kyrgidis A, Moscarella E, Benati E, Raucci M, Pellacani G, Zalaudek I, Argenziano G, Longo C (2016) Clinical indications

Surveillance and Follow-Up of Melanoma Patients for use of reflectance confocal microscopy for skin cancer diagnosis. JAMA Dermatol 152:1093–1098 Bradford PT, Freedman DM, Goldstein AM, Tucker MA (2010) Increased risk of second primary cancers after a diagnosis of melanoma. Arch Dermatol 146: 265–272 Brandberg Y, Mansson-Brahme E, Ringborg U, Sjoden PO (1995) Psychological reactions in patients with malignant melanoma. Eur J Cancer 31A:157–162 Brown RE, Stromberg AJ, Hagendoorn LJ, Hulsewede DY, Ross MI, Noyes RD, Goydos JS, Urist MM, Edwards MJ, Scoggins CR, Mcmasters KM, Martin RC 2nd (2010) Surveillance after surgical treatment of melanoma: futility of routine chest radiography. Surgery 148:711–716; discussion 716–717 Cancer Council Australia (2018) Clinical practice guidelines for the diagnosis and management of melanoma [Online]. https://wiki.cancer.org.au/australia/Guidelines: Melanoma Chao C, Wong SL, Ross MI, Reintgen DS, Noyes RD, Cerrito PB, Edwards MJ, Mcmasters KM, Sunbelt Melanoma Trial, G (2002) Patterns of early recurrence after sentinel lymph node biopsy for melanoma. Am J Surg 184:520–524; discussion 525 Cohn-Cedermark G, Rutqvist LE, Andersson R, Breivald M, Ingvar C, Johansson H, Jonsson PE, Krysander L, Lindholm C, Ringborg U (2000) Long term results of a randomized study by the Swedish Melanoma Study Group on 2-cm versus 5-cm resection margins for patients with cutaneous melanoma with a tumor thickness of 0.8–2.0 mm. Cancer 89:1495–1501 Cromwell KD, Ross MI, Xing Y, Gershenwald JE, Royal RE, Lucci A, Lee JE, Cormier JN (2012) Variability in melanoma post-treatment surveillance practices by country and physician specialty: a systematic review. Melanoma Res 22:376–385 Dalal KM, Patel A, Brady MS, Jaques DP, Coit DG (2007) Patterns of first-recurrence and post-recurrence survival in patients with primary cutaneous melanoma after sentinel lymph node biopsy. Ann Surg Oncol 14:1934–1942 Dalle S, Paulin C, Lapras V, Balme B, Ronger-Savle S, Thomas L (2006) Fine-needle aspiration biopsy with ultrasound guidance in patients with malignant melanoma and palpable lymph nodes. Br J Dermatol 155:552–556 Damude S, Hoekstra-Weebers JE, Francken AB, Ter Meulen S, Bastiaannet E, Hoekstra HJ (2016) The MELFO-study: prospective, randomized, clinical trial for the evaluation of a stage-adjusted reduced followup schedule in cutaneous melanoma patients-results after 1 year. Ann Surg Oncol 23:2762–2771 Dancey A, Rayatt S, Courthold J, Roberts J (2005) Views of UK melanoma patients on routine follow-up care. Br J Plast Surg 58:245–250 Danielsen M, Hojgaard L, Kjaer A, Fischer BM (2013) Positron emission tomography in the follow-up of cutaneous malignant melanoma patients: a systematic review. Am J Nucl Med Mol Imaging 4:17–28

863 Deichmann M, Benner A, Bock M, Jackel A, Uhl K, Waldmann V, Naher H (1999) S100-Beta, melanomainhibiting activity, and lactate dehydrogenase discriminate progressive from nonprogressive American Joint Committee on Cancer stage IV melanoma. J Clin Oncol 17:1891–1896 Dicker TJ, Kavanagh GM, Herd RM, Ahmad T, Mclaren KM, Chetty U, Hunter JA (1999) A rational approach to melanoma follow-up in patients with primary cutaneous melanoma. Scottish Melanoma Group. Br J Dermatol 140:249–254 Dieng M, Butow PN, Costa DS, Morton RL, Menzies SW, Mireskandari S, Tesson S, Mann GJ, Cust AE, Kasparian NA (2016) Psychoeducational intervention to reduce fear of cancer recurrence in people at high risk of developing another primary melanoma: results of a randomized controlled trial. J Clin Oncol 34:4405–4414 Faries MB, Morton DL (2006) The promise of metastasectomy in melanoma. Ann Surg Oncol 13:607–609 Faries MB, Steen S, Ye X, Sim M, Morton DL (2013) Late recurrence in melanoma: clinical implications of lost dormancy. J Am Coll Surg 217:27–34; discussion 34–36 Faries MB, Thompson JF, Cochran AJ, Andtbacka RH, Mozzillo N, Zager JS, Jahkola T, Bowles TL, Testori A, Beitsch PD, Hoekstra HJ, Moncrieff M, Ingvar C, Wouters M, Sabel MS, Levine EA, Agnese D, Henderson M, Dummer R, Rossi CR, Neves RI, Trocha SD, Wright F, Byrd DR, Matter M, Hsueh E, MackenzieRoss A, Johnson DB, Terheyden P, Berger AC, Huston TL, Wayne JD, Smithers BM, Neuman HB, Schneebaum S, Gershenwald JE, Ariyan CE, Desai DC, Jacobs L, Mcmasters KM, Gesierich A, Hersey P, Bines SD, Kane JM, Barth RJ, Mckinnon G, Farma JM, Schultz E, Vidal-Sicart S, Hoefer RA, Lewis JM, Scheri R, Kelley MC, Nieweg OE, Noyes RD, Hoon DSB, Wang HJ, Elashoff DA, Elashoff RM (2017) Completion dissection or observation for sentinel-node metastasis in melanoma. N Engl J Med 376:2211–2222 Ferrone CR, Ben Porat L, Panageas KS, Berwick M, Halpern AC, Patel A, Coit DG (2005) Clinicopathological features of and risk factors for multiple primary melanomas. JAMA 294:1647–1654 Fields RC, Coit DG (2011) Evidence-based follow-up for the patient with melanoma. Surg Oncol Clin N Am 20:181–200 Francken AB, Bastiaannet E, Hoekstra HJ (2005) Followup in patients with localised primary cutaneous melanoma. Lancet Oncol 6:608–621 Francken AB, Shaw HM, Accortt NA, Soong SJ, Hoekstra HJ, Thompson JF (2007) Detection of first relapse in cutaneous melanoma patients: implications for the formulation of evidence-based follow-up guidelines. Ann Surg Oncol 14:1924–1933 Francken AB, Accortt NA, Shaw HM, Colman MH, Wiener M, Soong SJ, Hoekstra HJ, Thompson JF (2008a) Follow-up schedules after treatment for malignant melanoma. Br J Surg 95:1401–1407

864 Francken AB, Shaw HM, Thompson JF (2008b) Detection of second primary cutaneous melanomas. Eur J Surg Oncol 34:587–592 Fusi S, Ariyan S, Sternlicht A (1993) Data on first recurrence after treatment for malignant melanoma in a large patient population. Plast Reconstr Surg 91:94–98 Gadd MA, Coit DG (1992) Recurrence patterns and outcome in 1019 patients undergoing axillary or inguinal lymphadenectomy for melanoma. Arch Surg 127:1412–1416 Garbe C, Paul A, Kohler-Spath H, Ellwanger U, Stroebel W, Schwarz M, Schlagenhauff B, Meier F, Schittek B, Blaheta HJ, Blum A, Rassner G (2003) Prospective evaluation of a follow-up schedule in cutaneous melanoma patients: recommendations for an effective follow-up strategy. J Clin Oncol 21:520–529 Garbe C, Hauschild A, Volkenandt M, Schadendorf D, Stolz W, Reinhold U, Kortmann RD, Kettelhack C, Frerich B, Keilholz U, Dummer R, Sebastian G, Tilgen W, Schuler G, Mackensen A, Kaufmann R (2007) Evidence and interdisciplinary consensus-based German guidelines: diagnosis and surveillance of melanoma. Melanoma Res 17:393–399 Gershenwald JE, Colome MI, Lee JE, Mansfield PF, Tseng C, Lee JJ, Balch CM, Ross MI (1998) Patterns of recurrence following a negative sentinel lymph node biopsy in 243 patients with stage I or II melanoma. J Clin Oncol 16:2253–2260 Gershenwald JE, Berman RS, Porter G, Mansfield PF, Lee JE, Ross MI (2000) Regional nodal basin control is not compromised by previous sentinel lymph node biopsy in patients with melanoma. Ann Surg Oncol 7:226–231 Green AC, Baade P, Coory M, Aitken JF, Smithers M (2012) Population-based 20-year survival among people diagnosed with thin melanomas in Queensland, Australia. J Clin Oncol 30:1462–1467 Guitera P, Menzies SW, Argenziano G, Longo C, Losi A, Drummond M, Scolyer RA, Pellacani G (2016) Dermoscopy and in vivo confocal microscopy are complementary techniques for diagnosis of difficult amelanotic and light-coloured skin lesions. Br J Dermatol 175:1311–1319 Hansson J, Bergenmar M, Hofer PA, Lundell G, ManssonBrahme E, Ringborg U, Synnerstad I, Bratel AT, Wennberg AM, Rosdahl I (2007) Monitoring of kindreds with hereditary predisposition for cutaneous melanoma and dysplastic nevus syndrome: results of a Swedish preventive program. J Clin Oncol 25:2819–2824 Hengge UR, Wallerand A, Stutzki A, Kockel N (2007) Cost-effectiveness of reduced follow-up in malignant melanoma. J Dtsch Dermatol Ges 5:898–907 Hoffman BR, Katsaros D, Scorilas A, Diamandis P, Fracchioli S, Rigault De La Longrais IA, Colgan T, Puopolo M, Giardina G, Massobrio M, Diamandis EP (2002) Immunofluorometric quantitation and histochemical localisation of kallikrein 6 protein in ovarian cancer tissue: a new independent unfavourable prognostic biomarker. Br J Cancer 87:763–771

R. L. Morton et al. Hofmann U, Szedlak M, Rittgen W, Jung EG, Schadendorf D (2002) Primary staging and follow-up in melanoma patients – monocenter evaluation of methods, costs and patient survival. Br J Cancer 87:151–157 Huibertse LJ, Van Eenbergen M, De Rooij BH, Bastiaens MT, Fossion LM, De La Fuente RB, Kil PJ, Koldewijn EL, Meier AH, Mommers RJ, Niemer AQ, Oddens JR, Oomens EH, Prins M, De Roos KP, Thissen MR, Timmermans MW, Wijsman BP, Van De Poll-Franse LV, Ezendam NP (2017) Cancer survivors’ preference for follow-up care providers: a cross-sectional study from the population-based PROFILES-registry. Acta Oncol 56:278–287 Jemal A, Siegel R, Ward E, Hao Y, Xu J, Murray T, Thun MJ (2008) Cancer statistics, 2008. CA Cancer J Clin 58:71–96 Jochems A, Schouwenburg MG, Leeneman B, Franken MG, Van Den Eertwegh AJ, Haanen JB, Gelderblom H, Uyl-De Groot CA, Aarts MJ, Van Den Berkmortel FW, Blokx WA, Cardous-Ubbink MC, Groenewegen G, De Groot JW, Hospers GA, Kapiteijn E, Koornstra RH, Kruit WH, Louwman MW, Piersma D, Van Rijn RS, Ten Tije AJ, Vreugdenhil G, Wouters MW, Van Der Hoeven JJ (2017) Dutch Melanoma Treatment Registry: quality assurance in the care of patients with metastatic melanoma in the Netherlands. Eur J Cancer 72:156–165 Jones MS, Torisu-Itakura H, Flaherty DC, Schoellhammer HF, Lee J, Sim MS, Faries MB (2016) Second primary melanoma: risk factors, histopathologic features, survival, and implications for follow-up. Am Surg 82:1009–1013 Kasparian NA, Branstrom R, Chang YM, Affleck P, Aspinwall LG, Tibben A, Azizi E, Baron-Epel O, Battistuzzi L, Bruno W, Chan M, Cuellar F, Debniak T, Pjanova D, Ertmanski S, Figl A, Gonzalez M, Hayward NK, Hocevar M, Kanetsky PA, Leachman S, Bergman W, Heisele O, Palmer J, Peric B, Puig S, Schadendorf D, Gruis NA, Newton-Bishop J, Brandberg Y, Melanoma Genetics, C (2012) Skin examination behavior: the role of melanoma history, skin type, psychosocial factors, and region of residence in determining clinical and selfconducted skin examination. Arch Dermatol 148:1142–1151 Khayat D, Rixe O, Martin G, Soubrane C, Banzet M, Bazex JA, Lauret P, Verola O, Auclerc G, Harper P, Banzet P, French Group of Research on Malignant, M (2003) Surgical margins in cutaneous melanoma (2 cm versus 5 cm for lesions measuring less than 2.1-mm thick). Cancer 97:1941–1946 Kittler H, Weitzdorfer R, Pehamberger H, Wolff K, Binder M (2001) Compliance with follow-up and prognosis among patients with thin melanomas. Eur J Cancer 37:1504–1509 Krahn G, Kaskel P, Sander S, Waizenhofer PJ, Wortmann S, Leiter U, Peter RU (2001) S100 beta is a more reliable tumor marker in peripheral blood for patients with newly occurred melanoma metastases compared with MIA, albumin and lactate-dehydrogenase. Anticancer Res 21:1311–1316

Surveillance and Follow-Up of Melanoma Patients Leiter U, Marghoob AA, Lasithiotakis K, Eigentler TK, Meier F, Meisner C, Garbe C (2009) Costs of the detection of metastases and follow-up examinations in cutaneous melanoma. Melanoma Res 19:50–57 Lim WY, Morton RL, Turner RM, Jenkins MC, Guitera P, Irwig L, Webster AC, Dieng M, Saw RPM, Low D, Low C, Bell KJL (2018) Patient preferences for followup after recent excision of a localized melanoma. JAMA Dermatol 154:420–427 Livingstone E, Krajewski C, Eigentler TK, WindemuthKieselbach C, Benson S, Elsenbruch S, Hauschild A, Rompel R, Meiss F, Mauerer A, Kahler KC, Dippel E, Mollenhoff K, Kilian K, Mohr P, Utikal J, Schadendorf D (2015) Prospective evaluation of follow-up in melanoma patients in Germany – results of a multicentre and longitudinal study. Eur J Cancer 51:653–667 Lo SN, Scolyer RA, Thompson JF (2018) Long-term survival of patients with thin (T1) cutaneous melanomas: a Breslow thickness cut point of 0.8 mm separates higher-risk and lower-risk tumors. Ann Surg Oncol 25:894–902 Machet L, Nemeth-Normand F, Giraudeau B, Perrinaud A, Tiguemounine J, Ayoub J, Alison D, Vaillant L, Lorette G (2005) Is ultrasound lymph node examination superior to clinical examination in melanoma follow-up? A monocentre cohort study of 373 patients. Br J Dermatol 152:66–70 Marciano NJ, Merlin TL, Bessen T, Street JM (2014) To what extent are current guidelines for cutaneous melanoma follow up based on scientific evidence? Int J Clin Pract 68:761–770 Mccaul KA, Fritschi L, Baade P, Coory M (2008) The incidence of second primary invasive melanoma in Queensland, 1982–2003. Cancer Causes Control 19:451–458 Memari N, Hayen A, Bell KJ, Rychetnik L, Morton RL, Mccaffery K, Thompson JF, Irwig L, Turner RM (2015) How often do patients with localized melanoma attend follow-up at a specialist center? Ann Surg Oncol 22(Suppl 3):S1164–S1171 Meyers MO, Yeh JJ, Frank J, Long P, Deal AM, Amos KD, Ollila DW (2009) Method of detection of initial recurrence of stage II/III cutaneous melanoma: analysis of the utility of follow-up staging. Ann Surg Oncol 16:941–947 Mills K, Emery J, Lantaff R, Radford M, Pannebakker M, Hall P, Burrows N, Williams K, Saunders CL, Murchie P, Walter FM (2017) Protocol for the melatools skin self-monitoring trial: a phase II randomised controlled trial of an intervention for primary care patients at higher risk of melanoma. BMJ Open 7:e017934 Moloney FJ, Guitera P, Coates E, Haass NK, Ho K, Khoury R, O’connell RL, Raudonikis L, Schmid H, Mann GJ, Menzies SW (2014) Detection of primary melanoma in individuals at extreme high risk: a prospective 5-year follow-up study. JAMA Dermatol 150:819–827 Moncrieff MD, Gyorki D, Saw R, Spillane AJ, Peach H, Oudit D, Geh J, Dziewulski P, Wilson E, Matteucci P, Pritchard-Jones R, Olofsson Bagge R, Wright FC,

865 Crampton N, Cassell O, Jallali N, Berger A, Kelly J, Hamilton S, Durrani A, Lo S, Paton E, Henderson MA (2018) 1 versus 2-cm excision margins for pT2-pT4 primary cutaneous melanoma (MelMarT): a feasibility study. Ann Surg Oncol 25:2541–2549 Mooney MM, Mettlin C, Michalek AM, Petrelli NJ, Kraybill WG (1997) Life-long screening of patients with intermediate-thickness cutaneous melanoma for asymptomatic pulmonary recurrences: a cost-effectiveness analysis. Cancer 80:1052–1064 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Elashoff R, Essner R, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Reintgen DS, Coventry BJ, Glass EC, Wang HJ, Group M (2006) Sentinel-node biopsy or nodal observation in melanoma. N Engl J Med 355:1307–1317 Morton RL, Rychetnik L, Mccaffery K, Thompson JF, Irwig L (2013) Patients’ perspectives of long-term follow-up for localised cutaneous melanoma. Eur J Surg Oncol 39:297–303 Murchie P, Hannaford PC, Wyke S, Nicolson MC, Campbell NC (2007) Designing an integrated follow-up programme for people treated for cutaneous malignant melanoma: a practical application of the MRC framework for the design and evaluation of complex interventions to improve health. Fam Pract 24:283–292 Murchie P, Nicolson MC, Hannaford PC, Raja EA, Lee AJ, Campbell NC (2010) Patient satisfaction with GP-led melanoma follow-up: a randomised controlled trial. Br J Cancer 102:1447–1455 Murchie P, Allan JL, Brant W, Dennis M, Hall S, Masthoff J, Walter FM, Johnston M (2015) Total skin self-examination at home for people treated for cutaneous melanoma: development and pilot of a digital intervention. BMJ Open 5:e007993 Nieweg OE, Kroon BB (2006) The conundrum of followup: should it be abandoned? Surg Oncol Clin N Am 15:319–330 Pflugfelder A, Kochs C, Blum A, Capellaro M, Czeschik C, Dettenborn T, Dill D, Dippel E, Eigentler T, Feyer P, Follmann M, Frerich B, Ganten MK, Gartner J, Gutzmer R, Hassel J, Hauschild A, Hohenberger P, Hubner J, Kaatz M, Kleeberg UR, Kolbl O, Kortmann RD, Krause-Bergmann A, Kurschat P, Leiter U, Link H, Loquai C, Loser C, Mackensen A, Meier F, Mohr P, Mohrle M, Nashan D, Reske S, Rose C, Sander C, Satzger I, Schiller M, Schlemmer HP, Strittmatter G, Sunderkotter C, Swoboda L, Trefzer U, Voltz R, Vordermark D, Weichenthal M, Werner A, Wesselmann S, Weyergraf AJ, Wick W, Garbe C, Schadendorf D, German Dermatological S, Dermatologiccooperative Oncology, G (2013) Malignant melanoma S3-guideline “diagnosis, therapy and follow-up of melanoma”. J Dtsch Dermatol Ges 11(Suppl 6):1–116, 1–126 Poo-Hwu WJ, Ariyan S, Lamb L, Papac R, Zelterman D, Hu GL, Brown J, Fischer D, Bolognia J, Buzaid AC (1999) Follow-up recommendations for patients with American Joint Committee on Cancer Stages I–III malignant melanoma. Cancer 86:2252–2258

866 Read RL, Haydu L, Saw RP, Quinn MJ, Shannon K, Spillane AJ, Stretch JR, Scolyer RA, Thompson JF (2015) In-transit melanoma metastases: incidence, prognosis, and the role of lymphadenectomy. Ann Surg Oncol 22:475–481 Read RL, Madronio CM, Cust AE, Goumas C, Watts CG, Menzies S, Curtin AM, Mann G, Thompson JF, Morton RL (2018) Follow-up recommendations after diagnosis of primary cutaneous melanoma: a population-based study in New South Wales, Australia. Ann Surg Oncol 25:617–625 Romano E, Scordo M, Dusza SW, Coit DG, Chapman PB (2010) Site and timing of first relapse in stage III melanoma patients: implications for follow-up guidelines. J Clin Oncol 28:3042–3047 Rueth NM, Cromwell KD, Cormier JN (2015) Long-term follow-up for melanoma patients: is there any evidence of a benefit? Surg Oncol Clin N Am 24:359–377 Rychetnik L, Mccaffery K, Morton RL, Thompson JF, Menzies SW, Irwig L (2013) Follow-up of early stage melanoma: specialist clinician perspectives on the functions of follow-up and implications for extending follow-up intervals. J Surg Oncol 107:463–468 Saiag P, Bosquet L, Guillot B, Verola O, Avril MF, Bailly C, Cupissol D, Dalac S, Danino A, Dreno B, Grob JJ, Leccia MT, Renaud-Vilmer C, Negrier S, Societe Francaise De D (2007) Management of adult patients with cutaneous melanoma without distant metastasis. 2005 update of the French Standards, Options and Recommendations guidelines. Summary report. Eur J Dermatol 17:325–331 Schultz S, Kane M, Roush R, Miller V, Berd D, Goldman L, Mastrangelo M (1990) Time to recurrence varies inversely with thickness in clinical stage 1 cutaneous melanoma. Surg Gynecol Obstet 171:393–397 Snoswell CL, Caffery LJ, Whitty JA, Soyer HP, Gordon LG (2018) Cost-effectiveness of skin cancer referral and consultation using teledermoscopy in Australia. JAMA Dermatol 154:694–700 The National Cancer Institute, S., Epidemiology, and End Results Program (2018) Cancer stat facts: melanoma of the skin [Online]. https://seer.cancer.gov/statfacts/html/ melan.html. Accessed 31 Aug 2018 Thomas JM, Newton-Bishop J, A’Hern R, Coombes G, Timmons M, Evans J, Cook M, Theaker J, Fallowfield M, O’Neill T, Ruka W, Bliss JM, United Kingdom Melanoma Study, G, British Association of Plastic, S, Scottish Cancer Therapy, N (2004) Excision margins in

R. L. Morton et al. high-risk malignant melanoma. N Engl J Med 350:757–766 Trask PC, Paterson AG, Griffith KA, Riba MB, Schwartz JL (2003) Cognitive-behavioral intervention for distress in patients with melanoma: comparison with standard medical care and impact on quality of life. Cancer 98:854–864 Truong A, Strazzulla L, March J, Boucher KM, Nelson KC, Kim CC, Grossman D (2016) Reduction in nevus biopsies in patients monitored by total body photography. J Am Acad Dermatol 75:135–143 e5 Turner RM, Bell KJ, Morton RL, Hayen A, Francken AB, Howard K, Armstrong B, Thompson JF, Irwig L (2011) Optimizing the frequency of follow-up visits for patients treated for localized primary cutaneous melanoma. J Clin Oncol 29:4641–4646 Veronesi U, Farante G, Galimberti V, Greco M, Luini A, Sacchini V, Andreola S, Leoni F, Menard S, Ronco M et al (1991) Evaluation of resection margins after breast conservative surgery with monoclonal antibodies. Eur J Surg Oncol 17:338–341 Vuylsteke RJ, Van Leeuwen PA, Statius Muller MG, Gietema HA, Kragt DR, Meijer S (2003) Clinical outcome of stage I/II melanoma patients after selective sentinel lymph node dissection: long-term follow-up results. J Clin Oncol 21:1057–1065 Watts CG, Dieng M, Morton RL, Mann GJ, Menzies SW, Cust AE (2015) Clinical practice guidelines for identification, screening and follow-up of individuals at high risk of primary cutaneous melanoma: a systematic review. Br J Dermatol 172:33–47 Wheeler T (2006) Psychological consequences of malignant melanoma: patients’ experiences and preferences. Nurs Stand 21:42–46 Youlden DR, Youl PH, Soyer HP, Aitken JF, Baade PD (2014) Distribution of subsequent primary invasive melanomas following a first primary invasive or in situ melanoma Queensland, Australia, 1982–2010. JAMA Dermatol 150:526–534 Youlden DR, Baade PD, Soyer HP, Youl PH, Kimlin MG, Aitken JF, Green AC, Khosrotehrani K (2016) Ten-year survival after multiple invasive melanomas is worse than after a single melanoma: a population-based study. J Invest Dermatol 136:2270–2276 Zhang H, Bermejo JL, Sundquist J, Hemminki K (2008) Modification of second cancer risk after malignant melanoma by parental history of cancer. Br J Cancer 99:536–538

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes, and Selection of Treatment Options John F. Thompson, Nicola Mozzillo, and Merrick I. Ross

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 868 Etiology of Local and In-transit Metastases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869 Incidence of Locoregional Metastasis and Survival . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 870 Treatment Options for Local and In-transit Metastases: Overview . . . . . . . . . . . . . . . 874 Surgical Excision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874 Cryotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875 Topical Treatments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875 Diphencyprone Cream . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875 Imiquimod Cream . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876 Intralesional Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . BCG . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . DNCB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interferon Alpha . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Allovectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coxsackie Virus A-21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Interleukin-2 (IL-2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

877 877 877 877 878 878 878

J. F. Thompson (*) Melanoma Institute Australia, Faculty of Medicine and Health, The University of Sydney, Sydney, NSW, Australia Department of Melanoma and Surgical Oncology, Royal Prince Alfred Hospital, Sydney, NSW, Australia e-mail: [email protected] N. Mozzillo Department Melanoma and Soft Tissue, Istituto Nazionale dei Tumori, Napoli, Italy e-mail: [email protected] M. I. Ross Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_32

867

868

J. F. Thompson et al. Rose Bengal (PV-10) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Talimogene Laherparepvec (T-VEC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Daromun (Combined IL-2 and TNF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Intralesional Oncolytic Immunotherapy and the Abscopal Effect . . . . . . . . . . . . . . . . . . . . . .

878 878 879 879

Laser and Light-Based Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ablative Laser Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Non-ablative Laser Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Radio-Frequency Ablation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

880 880 880 881

Photodynamic Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 881 Electrochemotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 882 Reported Results of ECT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 883 ECT in Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 884 Regional Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885 Isolated Limb Perfusion and Infusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885 Radiation Therapy for Locally Recurrent Melanoma and In-transit Metastases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886 Amputation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886 Systemic Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 886 Adjuvant and Neoadjuvant Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887 Adjuvant Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887 Neoadjuvant Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 888

Abstract

The most frequent site of first melanoma recurrence is locoregional, and this is frequently a harbinger of systemic recurrence. Local recurrence, satellitosis, and in-transit metastasis are all now understood to be forms of “intralymphatic” metastasis, with similar prognostic implications, a concept incorporated in the most recent 8th Edition (2017) AJCC Melanoma Staging System. In some patients with locoregional melanoma recurrence, however, systemic metastasis never occurs. In these patients and in those with troublesome symptoms from their locoregionally recurrent disease, effective treatment is clearly important. If it is easily possible, complete surgical excision is the best option, but if it is not, there is a range of other options, which include local, regional, and systemic therapies. The choice of treatment is guided by the number of lesions, their anatomic location, whether they are cutaneous or subcutaneous, their size, and the presence or absence of metastatic disease in regional lymph nodes or at

distant sites. Local therapies include topical therapies, cryotherapy, diathermy-fulguration, laser ablation, radio-frequency ablation, intralesional injection of vaccines or cytotoxic agents, and electrochemotherapy. Regional therapies include isolated limb perfusion, isolated limb infusion and radiation therapy. Even amputation may occasionally be considered. An understanding of the options, and the benefits and disadvantages of each, will allow selection of the most appropriate treatment strategy for each patient with local melanoma recurrence, satellitosis, or intransit metastasis.

Introduction When melanoma recurs after definitive surgical treatment of the primary tumor, it most commonly first does so locally or regionally. The term “locoregional” recurrence includes true local recurrence and satellite metastases (within or close to the primary melanoma excision site), in-

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

transit recurrence (between the primary site and the nearest regional lymph node basin), and metastases in the regional lymph nodes. In the literature, the definitions of local and in-transit recurrence have varied considerably, with limits of 5 cm for in-transit lesions. The American Joint Committee of Cancer (AJCC) Melanoma Staging System uses a distance of >2 cm from the primary lesion to define in-transit metastases. The 8th Edition of the AJCC Melanoma Staging System, published in 2017, introduced the term “intralymphatic metastases” for the combined group of local and intransit lesions, based on the observation that survival outcomes for the two entities are very similar. Within the concept of local recurrence, the terms satellite and microsatellite metastases are used. Satellite metastases are usually regarded as clinically apparent lesions within 2 cm of the primary melanoma site (Balch 2009) and microsatellites as those immediately adjacent to or deep to a primary melanoma, identified histologically (Gershenwald et al. 2017). In three large studies, each including over 1000 patients with recurrent melanoma, similar rates of locoregional recurrence were reported (71–77%), the remaining 23–29% being distant recurrences (Fusi et al. 1993; Soong et al. 1998; Reintgen et al. 1992) This chapter will consider the selection of appropriate treatment options for locally recurrent and in-transit metastatic melanoma. The management of regional lymph node metastases is discussed separately (see chapters ▶ “Local and Recurrent Regional Metastases of Melanoma” and ▶ “Radiotherapy for Primary and Regional Melanoma”). Complete surgical excision of low-volume recurrent disease is usually the best and simplest form of therapy, but this is not always possible due to the size, location, or number of recurrences; thus other treatment options must be considered. These include localized therapies such as topical therapies, cryotherapy, diathermy-fulguration, laser ablation, radio-frequency ablation, intralesional injection of vaccines or cytotoxic agents, and electrochemotherapy and regional treatments such as isolated limb perfusion, isolated limb infusion, and regional radiotherapy. The rare option of amputation must also be borne in mind.

869

Some patients who develop locoregional recurrences already have or rapidly develop systemic disease progression; today this is usually an indication for systemic therapy. In other patients, the localized metastases can remain regionally confined for many years or even decades, occasionally indefinitely, and treatments to control this disease and to prevent or reduce troublesome symptoms are required. A wide variety of locoregional treatment options is available. The observation that so many options are being used at specialist melanoma treatment centers around the world reflects the reality that no single treatment modality is suitable for every patient or available in every institution. This chapter aims to explain and discuss all the treatment options and assist the decision-making process for patients with locoregionally recurrent melanoma when surgical excision is not easily possible and systemic therapy is not considered appropriate or has been ineffective.

Etiology of Local and In-transit Metastases Local and in-transit melanoma metastases are usually considered to be due to entrapment of tumor cells in dermal and subdermal lymphatics between the site of the primary melanoma and the regional lymph node basin (or basins) to which lymphatics from the primary tumor drain (Gershenwald and Fidler 2002). However, other theories have been suggested. These include tumor cell dispersion through tissue fluid (McCarthy 2002), spread of tumor cells around the abluminal surface of lymphatics and blood vessels in a pericytic location (termed angiotropism) (Wilmott et al. 2012), and implantation of tumor cells after spread via the bloodstream (Heenan and Ghaznawie 1999). When present, in-transit metastases are usually multiple, evolve over time, and are often a harbinger of subsequent systemic disease; however, they may also be simply a nuisance and/or a concern for the patient, with no significant impact on daily activities or survival. Several studies have shown that classification of local and in-transit relapses based on distance from the primary tumor site is of little prognostic value (Cascinelli et al. 1986; Singletary et al. 1988), and while earlier staging classifications distinguished between local

870

recurrence/satellitosis and in-transit metastases (Fleming et al. 1997), more recent staging systems have recognized the pathophysiologic and prognostic similarities between these entities and classified them together as N3 disease (Balch et al. 2001) (see chapter ▶ “Melanoma Prognosis and Staging”). The 2017 update of the AJCC Melanoma Staging System considers any number of intransit, satellite, and/or microsatellite metastases in conjunction with the number of tumor-involved nodes to classify patients as N1 (no involved nodes), N2 (1 involved node), or N3 (2 involved nodes or any number of matted nodes) (Gershenwald et al. 2017).

Incidence of Locoregional Metastasis and Survival The incidence rates of local, in-transit, and regional node recurrence of melanoma have been detailed in a number of studies and are summarized in Table 1. Across 29 studies reporting local, in-transit, and/or regional lymph node recurrence as a first recurrence, individual study estimates ranged widely for the different types of recurrence: 0.7–12.8% for local recurrence, 0.3–11% for in-transit metastasis, and 0.5–19.1% for regional node metastasis. Twelve studies grouped all locoregional recurrences together, and these gave a median incidence of 10.2%. A systematic review of 36 studies assessing the association of primary melanoma treatment and intralymphatic metastases (i.e., local and in-transit lesions) reported a pooled estimate of 3.4% but with enormous variability across the individual studies (from 0.62% to 15%) (Sloot et al. 2016). Ideally, incidence should be estimated from a large, unselected population of melanoma patients, whereas many of the individual studies and those in the systematic review included a select group of patients such as stage II only or all with primary tumors >1 mm. Variability in estimates probably also arises through differing definitions of local, in-transit, and locoregional recurrences, differing time points for measuring recurrence, and differing methods of diagnosing nodal disease.

J. F. Thompson et al.

Estimates of survival after local, in-transit, and regional node recurrence were not clearly reported in several of the studies; however, three of them reported 5-year survival rates after local recurrence that were quite similar: 41.9%, 45%, and 57% (Soong et al. 1998; Reintgen et al. 1992; Dalal et al. 2007). A large study of 505 patients with in-transit recurrence within a population of 11,614 melanoma patients found a median survival time of 19.9 months and a 5-year survival rate of 32.8% (Read et al. 2015a), an estimate quite similar to the rate of 33.3% reported in a previous, smaller study (Dalal et al. 2007). Factors associated with recurrence of melanoma have been identified in many studies. These include tumor features such as primary tumor thickness and site, ulceration, mitotic rate, sentinel lymph node positivity, and patient characteristics such as age and gender (Cascinelli et al. 1986; Lyth et al. 2017; Matheson et al. 2017; Berger et al. 2017). Fewer studies have explored features that specifically predict local, in-transit, or regional node recurrence; however, several have reported that increased primary tumor thickness and primary tumor location on an extremity are associated with an increased likelihood of local recurrence (Griffiths and Briggs 1986; Reintgen et al. 1992; Wagner et al. 2003). The presence of three or more positive lymph nodes was associated with a higher likelihood of intransit recurrence in one study (Cascinelli et al. 1986), and primary lesions on lower extremities compared to upper extremities or the trunk, head, or neck were associated with in-transit recurrences in another report (Calabro et al. 1989). A further study added age >50 years to the previously listed features associated with in-transit metastases (Pawlik et al. 2005). The study of 505 patients with in-transit metastases by Read et al. found that those with primary tumors >1 mm in thickness more frequently developed in-transit metastases than those with primary tumors 1.0 mm thick, non-palpable nodes, underwent lymphadenectomy

1998 4568

2000 200

Jansen

48

26

1990 1965 1992 4185 1993 1090

Wong Reintgen Fusi

Stage I

Stage I, = 1.0 mm, non-palpable nodes Stages I and II, undergoing SLN mapping Stage I, non-palpable nodes, undergoing SLNB Undergoing lymphatic mapping, non-palpable nodes >1 mm thick primary lesion, no regional therapy, failed SLNB, 0.75 mm thick primary lesion, non-palpable nodes, no distant disease Stages I and II, undergoing SLNB Stages I and II First cutaneous malignant melanoma, Clarke level II, no distant metastases, curative surgery Melanoma patients No distant metastases

Single primary lesion 2017 581 Stage II, >12 mo follow-up 2017 1029 Stages I and II, 2016 27974 Studies reporting (in 36 numbers of patients studies) with intralymphatic metastases

2015 11614

2012 2829 2012 404

Francken 2008 4704 Hohnheiser 2011 2487

2006 1187

Rukowski

123 NS

171

NS

NS 93

873 523

203

NS

4.8

25.6 12–132

8

0.8

5.2

30

>12

2.1 6.9

190 1.6

58 28

86 1.8 NS

50

NS

40.6

36 18

63.6 156

36

37.5

1

19

26

NS

1.8

4.5

41 1.4 NS

95 2 NS

10

NS

3.8

36

52 3.5

9

213 1.8

NS 73 18

300 6.4 NS

40

NS

9.6

3.5

3.8

5.7

7.7

NS 1148 3.4

56

NS

99 NS

181 NS

60

91

NS

108

NS

NS NS

481 362

100

NS

18.6

10.2 14.6

9.6

NS

NS

NS

NS NS

74 NS

57.1

NS

NS

NS

61.1

NS NS

44 NS

33.3

NS

NS

NS

NS

45 NS

NS NS

NS

3 year; 29

NS

NS

NS

NS NS

50 NS

21.8

NS

NS

NS

NS

NS NS

NS 34.4

NS

NS

NS

17.9 (in-transit) NS

13 mo 9 (in-transit)

NS NS

NS

9.9–13.5 mo (in-transit and local)

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . . 873

874

performance estimates were similar to results reported in another study, in which there was a positive predictive value of 18% (46/259) and a sensitivity of 53% (46/86) (Pawlik et al. 2005). Whether having a sentinel lymph node biopsy increased a patient’s risk of in-transit metastasis became a matter of concern after it was suggested that there was a higher rate of subsequent intransit metastasis after sentinel lymph node biopsy, possibly due to an iatrogenic risk of increased in-transit metastasis if a sentinel lymph node biopsy was performed (Estourgie et al. 2004; Thomas and Clark 2004). However, the first Multicenter Selective Lymphadenectomy Trial-I (MSLT-1) clearly demonstrated no difference in the rate of in-transit metastasis between patients in the wide local excision alone group and those in the wide local excision plus sentinel lymph node biopsy group (Morton et al. 2006), and this convincing evidence from a large, multicenter randomized controlled trial was supported by the results of large non-randomized studies of patients who had undergone sentinel lymph node biopsy (Pawlik et al. 2005; Wong et al. 2006; Read et al. 2015a). Few studies have reported median times to different types of recurrence; however, in one a mean interval between excision of the primary melanoma and appearance of a local metastasis was 18.4 months (Griffiths and Briggs 1986). Across four studies the median time to the first appearance of in-transit melanoma metastasis ranged from 9 to 24 months (Dalal et al. 2007; Wong et al. 1990; Clemente-Ruiz de Almiron and Serrano-Ortega 2012), with the largest study reporting a median time to first in-transit recurrence of 17.9 months (Read et al. 2015a). Melanoma recurrence in regional lymph nodes was reported after a median interval from primary diagnosis of 55 months in one small study (Zapas et al. 2003) and after 19 months in another (Mansson-Brahme et al. 1994). Patients who are sentinel lymph node-positive appear to have earlier locoregional recurrence, after a median of 13 months, compared with 24 months in patients who are sentinel node-negative (Dalal et al. 2007).

J. F. Thompson et al.

Approximately 70–80% of all locoregional relapses occur within 3 years of definitive management of the primary melanoma (Reintgen et al. 1992; Fusi et al. 1993; Morton et al. 2014; Faries et al. 2010), but some occur much later.

Treatment Options for Local and In-transit Metastases: Overview The choice of therapy for locally recurrent and/or in-transit melanoma is generally guided by the number of lesions, their anatomic location, whether they are cutaneous or subcutaneous, their size, and the presence or absence of metastatic disease in regional nodes and/or at distant sites. Treatment options may be classified as local, regional, or systemic. Local therapies (apart from surgical resection) include topical therapies, diathermy-fulguration, cryotherapy, laser ablation, radio-frequency ablation, intralesional injection of cytotoxic or oncolytic agents, and electrochemotherapy. Regional therapies include isolated limb perfusion and isolated limb infusion and regional radiation therapy. Amputation is yet another option but very rarely considered necessary. Systemic therapies proven to be effective in patients with metastatic disease are now being used in patients with multiple and or unresectable disease either alone or in combination with local and/or regional therapies.

Surgical Excision Surgical excision is appropriate as the initial treatment for patients who have a single or a small number of local or in-transit metastases, particularly those with a solitary subcutaneous metastasis. Unfortunately, further locoregional recurrence often occurs following surgical excision. In a study of 648 patients with a local recurrence as the first melanoma recurrence that was treated with wide local excision, 196 (30%) developed further local recurrences, while 7 (1%) developed in-transit metastases, and 171 (26%) developed nodal

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

metastases during a median follow-up period of 39 months (Dong et al. 2000). In the group with further local recurrence, 107/196 (55%) died of melanoma within 39 months, while 4/7 (57%) patients with in-transit metastases and 113/171 (66%) patients with regional node recurrence died. This contrasted sharply with a much lower death rate in patients who had a single local recurrence excised and who developed no further lesions, with only 28/124 patient deaths (23%) during follow-up.

Cryotherapy Cryotherapy for cutaneous melanoma recurrences and metastases is based on the principle that there is complete destruction of melanoma cells at 4 oC to 7 oC, whereas keratinocytes and connective tissue cells are not destroyed until the temperature drops to 20 oC (Breitbart 1990). In addition, studies have shown that cryotherapy can stimulate an immune response by the release of tumor antigens as a result of tumor tissue necrosis (Tanaka 1982). Cryotherapy is usually delivered using a liquid nitrogen spray gun which freezes the lesion. Spraying leads to formation of a white ice field that increases steadily in diameter, and treatment is continued until at least 5 mm of the surrounding skin is involved. The area is then allowed to thaw, and the process is repeated, ideally twice more. The rapid heat transfer from the skin by the liquid nitrogen causes intra- and extracellular ice formation. This ice causes rapid cytolysis, with vasoconstriction and ischemic necrosis occurring over the following few hours. Afterward an inflammatory response commences that may give rise to blistering and edema of surrounding tissue (Thai and Sinclair 1999). One report of cryotherapy stated that 67 patients with advanced cutaneous metastatic melanoma were treated with cryotherapy and 3–10 years later no tumor progression was seen. No further details were provided (Breitbart 1990). Another study reported 30 patients with metastatic melanoma treated with cryotherapy with good responses and an average time to

875

remission of 36 months (Grotmann et al. 1991). More recent studies have reported combinations of treatment that include cryotherapy. In one of these, cryotherapy was used in combination with topical imiquimod in 20 patients with locoregional cutaneous melanoma metastases. After a mean of five treatment sessions, eight had a complete response, and five had a partial response, while seven had locoregional progression (Rivas-Tolosa et al. 2016). The treatment appeared to have no impact on disease progression, with 16 going on to develop systemic disease. However, no serious side effects were noted, and only slight erythema occurred after imiquimod. In a case report of the combination of cryotherapy with intralesional injection of interleukin-2 in one patient, full remission was achieved and persisted for 5 years (Wulfken et al. 2015).

Topical Treatments Diphencyprone Cream Diphencyprone (DPCP) (2,3-Diphenylcycloprop2-en-1-one) is a contact sensitizer that induces an allergic response following contact with the skin. It is inexpensive and has been used topically as single-agent therapy for melanoma and also in combination with other treatments. In a study of 50 patients sensitized to DPCP and subsequently treated with 0.01–0.1% DPCP in aqueous cream for at least 1 month, 23 (46%) achieved complete clearance of their cutaneous melanoma metastases, while an additional 19 (38%) showed a partial response (Damian et al. 2014). Six complete responders later developed cutaneous recurrences that were cleared completely with intensive repeat DPCP treatment. This finding was slightly better than that seen in a subsequent series of 54 patients treated with DPCP, in which a complete response occurred in 22% and a partial response in 39% (Read et al. 2017a). It is noteworthy, however, that in both studies nearby but untreated cutaneous and subcutaneous metastases regressed and very occasionally nodal and systemic metastases

876

J. F. Thompson et al.

Fig. 1 In-transit metastases on the scalp 6 weeks (panel A) and 6 months (panel B) after commencement of topical diphencyprone therapy. Nine years later there is no

evidence of further recurrence on the scalp. (Images courtesy of Professor Diona Damian)

underwent involution, apparently an abscopal effect of the DPCP. The effects of DPCP therapy are illustrated in Fig. 1.

describe complete clearance with imiquimod treatment (Grady and Spencer 2016) (Moon and Spencer 2013), and a third report of two patients describes a response to imiquimod after failure of intralesional bacillus Calmette-Guerin (BCG) (Kibbi et al. 2015). Several studies of combination local therapies that include topical imiquimod treatment have been reported. In one study a combination of topical 5% imiquimod with fluorouracil 5 days a week until lesion response was used to treat 5 patients with 3–17 primarily cutaneous lesions (Florin et al. 2012). A complete response was seen in 19 lesions and a partial response in 25. A study of the combination of imiquimod with carbon dioxide laser was used to treat two patients, while imiquimod with electrocoagulation was used in a further two patients (Elfatoiki et al. 2014). Two patients experienced complete regression of lesions after treatment with imiquimod+laser and imiquimod +electrocoagulation, while two progressed. Neither responder remained free of disease, with both developing further lesions and distant metastases during follow-up. In a study of 11 patients treated with intralesional interleukin-2 and topical imiquimod and retinoid cream, a complete response was reported in all patients within 1–3 months, and 9 remained alive at 2 years, 7 of whom had no further recurrence (Shi et al. 2015). Six patients experienced rigors as a side effect, and

Imiquimod Cream Imiquimod [1-(2-methylpropyl)imidazo[4,5-c] quinolin4-amine] is a small, synthetic molecule that binds to and activates Toll-like receptor 7 (TLR7). Its mechanism of action is complex and involves activation of immune cells, secretion of cytokines, and activation of Langerhans cells (Miller et al. 1999; Tyring 2001). No randomized controlled trials of topical imiquimod treatment for melanoma have been carried out, and evidence consists of case series. A review of 11 case studies involving 17 melanoma patients treated with 5% imiquimod cream applied once daily (for 6–8 h) for 5 days per week, three times a week, or 7 days a week, under occlusive conditions, for varying time periods of 8–72 weeks reported complete regression of metastases in 12/17 patients and a partial response in 2/17, with only mild side effects of short duration (Sisti et al. 2014). Only two studies reported recurrence and survival, with one documenting no recurrence in their single patient at 15 months (Wolf et al. 2003), and in the other, the patient died of disease progression at 2 years (Hesling et al. 2004). Two additional case reports

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

one required surgical debridement of necrotic tissue resulting from the intralesional therapy. A report of the combination of local application of imiquimod and injection of dye with laser irradiation in 11 patients described a complete response in 6, no grade 4 toxicities, and a 12-month survival of 70% (Li et al. 2010). The combination of topical imiquimod with a systemic therapy, ipilimumab, has been reported in two patients who had intransit metastases as well as distant organ and lymph node metastases (Joseph et al. 2016). Both patients experienced a complete response in distant and in-transit lesions, with continued remission 6 months after treatment. A trial that compares imiquimod with DPCP is currently recruiting but is designed as a pilot study and with a target accrual of only 30 patients is not powered to provide precise estimates of relative efficacy (Read et al. 2017b). The commencement of this trial does, however, demonstrate that there is a perceived equivalence of the two treatments at least among some clinicians and researchers. As simple, noninvasive, and relatively inexpensive treatments, topical therapy with DPCP or imiquimod may be beneficial in treating local or in-transit lesions in melanoma patients with cutaneous/dermal metastases in whom other treatments are not considered appropriate. Ideally a large-scale, randomized trial with welldefined outcomes measured at specified time points would provide greater certainty in relation to the relative efficacy of the two treatment modalities.

877

BCG One of the earliest agents to be used for intralesional treatment of solid tumors was bacillus CalmetteGuerin (BCG) (Mastrangelo et al. 1975; Morton et al. 1974). It consists of an attenuated strain of Mycobacterium bovis which can induce a non-specific host immune response leading to tumor regression. In a study of 45 patients treated with intralesional BCG, an overall response rate of 96% was reported, with a complete response in 68% (Bauer et al. 1990). An earlier study had reported a somewhat lower response rate (74%) in 20 patients treated with intralesional BCG (Storm et al. 1979). BCG injections are associated with a low but finite incidence of systemic complications, including allergic reactions and systemic “BCG-osis.” These diverse and occasionally lethal adverse effects have limited the use of BCG in clinical practice (Karakousis et al. 1976; Robinson 1977).

DNCB Dinitrochlorobenzene (DNCB) is an organic compound that induces a hypersensitivity reaction in most people. It was evaluated in a small prospective randomized trial of 18 melanoma patients, 9 of whom received intralesional DNCB, while 9 received intralesional BCG. Results showed that DNCB appeared to be equally effective as BCG but with significantly fewer systemic side effects (Cohen et al. 1978). Until recently, however, pharmaceutical grade of DNCB has not been available commercially, which has limited the ability to evaluate its efficacy or consider it as a feasible treatment option.

Intralesional Therapies Intralesional injection can be used in patients who have a limited number of dermal or subdermal intransit metastases. The agents that have been injected include BCG, interferon, interleukin-2, Rose Bengal (PV-10), talimogene laherparepvec (T-VEC), and tumor necrosis factor (TNF). Particularly promising results have been seen recently with several agents, including Rose Bengal (PV10), talimogene laherparepvec (T-VEC), Daromun (L19-IL2+L19-TNF), and oncolytic viruses.

Interferon Alpha Interferon alpha (IFNα) has been used for intralesional injection, and a study of 51 patients with histologically proven metastatic melanoma and at least 1 skin metastasis was treated with either purified natural IFN alpha or recombinant IFN alpha 2b (Wussow et al. 1988). A complete or partial response was seen in 24 of the 51 patients (47%). Also noted in this study were nine

878

transient responses in non-injected metastases. Little additional evidence on the efficacy of intralesional IFN is available, and its clinical use is currently uncommon.

Allovectin Intralesional injection of a DNA carrier system encoding genes for a major histocompatibility complex class I heavy and light chains of HLAB7 and β2microglobulin, called Allovectin (velimogene aliplasmid), was designed to stimulate the immune response and was trialed in 390 melanoma patients (Vical 2013). However it failed to meet the sponsoring company’s specified minimum efficacy for further development, and it has been abandoned.

Coxsackie Virus A-21 Another novel system uses the Coxsackie virus A21 which targets and destroys cells expressing intercellular adhesion molecule (ICAM-1) and decay accelerating factor (DAF), both of which are upregulated on melanoma cells. A study of late stage melanoma patients receiving intratumoral injection of Coxsackie virus A21 (called CALM) reported an overall response rate of 28% (16/57 patients) (Andtbacka et al. 2015a) (https://viralytics.com/wp-content/uploads/2016/ 10/IOVMC-2016-Andtbacka-CALM-extension. pdf – accessed 15 February 2019). A further study (called MITCI) by the same group is combining Coxsackie intralesional treatment with systemic ipilimumab in 26 melanoma patients with stages III and IV disease. Preliminary data suggest that the overall response rate may be around 80%, higher than the response rates for either agent used alone. However full publications providing details of these trials are not yet available, and considerable uncertainty about their efficacy and cost-effectiveness remains.

Interleukin-2 (IL-2) Interleukin-2 (IL-2) is highly toxic when administered systemically (Eklund and Kuzel 2004; Tarhini and Agarwala 2005) but has been explored

J. F. Thompson et al.

as an intralesional therapy in melanoma patients. A systematic review of six studies of intralesional injection of IL-2 in melanoma patients with intransit lesions showed a complete response in 78% of lesions and 50% of patients (Byers et al. 2014). Side effects tended to be minor, with localized pain, swelling, and flu-like symptoms the most frequent. The high response rate and durability of intralesional IL-2 treatment have made it the first-line treatment for in-transit disease in some centers (Temple-Oberle et al. 2014).

Rose Bengal (PV-10) Rose Bengal is a small molecule in the xanthene family that has long been used in eye-drop form to stain damaged conjunctival and corneal cells and also by systemic administration for liver function testing. Its antitumor activity was recognized more recently, and a sterile 10% Rose Bengal solution (PV-10) has been developed for intralesional application in patients with melanoma metastases and also patients with metastatic breast and liver cancer. Its mechanism of action is by selective uptake through the cell membrane of cancer cells and accumulation in lysosomes. This accumulation leads to rapid lysosomal rupture and destruction of the cell by its own enzymes. A trial involving 80 melanoma patients who received intralesional injections of PV-10 showed an overall response rate of 51%, with 26% having a complete response (Thompson et al. 2015). A multicenter international trial is under way comparing intralesional PV-10 with systemic therapy or intralesional oncolytic viral therapy (NCT02288897).

Talimogene Laherparepvec (T-VEC) Talimogene laherparepvec (T-VEC) is a type1 herpes simplex virus that preferentially infects and kills cancer cells. Modifications to this virus have been designed to enhance specificity for tumor cells and increase the immune response (Liu et al. 2003). A randomized controlled trial of 436 patients compared intralesional T-VEC with subcutaneous injections of granulocyte

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

macrophage colony-stimulating factor (GMCSF); a durable response (6 months) was reported in 16.3% of T-VEC treated patients compared with 2.1% of GM-CSF treated patients (Andtbacka et al. 2015b). There was a median increase in survival of 4 months (23.3 months TVEC versus 18.9 months GM-CSF). Combinations of T-VEC and the newer systemic treatments are currently being trialed after favorable safety outcomes in phase 1 studies (Puzanov et al. 2016; Long et al. 2016). A phase 2 trial of 198 patients with stage III and stage IV unresectable melanoma treated with either T-VEC plus ipilimumab or ipilimumab alone reported a response in 39/100 patients treated with the combination therapies compared to 18/98 treated with ipilimumab alone, giving a relative risk of 2.2, meaning that the combination therapy was more than twice as likely to elicit a positive response than the ipilimumab alone (Chesney et al. 2017). A phase 3 randomized trial comparing the response rate, progression-free survival, and overall survival of single-agent pembrolizumab to pembrolizumab plus T-VEC as frontline treatment for patients with unresectable stage III disease (and therefore including patients with multiple in-transit lesions) and stage IV disease with injectable lesions has recently completed accrual. The initial phase 1-2 combination lead-in arm to the randomized portion of the trial showed the combination to be well tolerated and demonstrated an encouragingly high response rate.

Daromun (Combined IL-2 and TNF) Daromun is the combination of two immunocytokines L19-IL2 (Darleukin) and L19-TNF (Fibromun) that individually boost immune responses in melanoma patients (Johannsen et al. 2010; Weide et al. 2014; Papadia et al. 2013). Data from a mouse study have shown that a combination of L19-IL2 and L19-TNF given systemically induced complete tumor remission. The authors hypothesized that used alone these immunocytokines induce tumor retardation but in combination may result in more complete tumor eradication (Pretto et al. 2014). A treatment series of the combination of L19-IL2

879

and L19-TNF as intralesional injection therapy in 22 unresectable melanoma patients reported a partial response in 10 patients at week 12. Also of interest was some response in 9 of 13 noninjected lesions, suggesting a systemic bystander effect of treatment (Danielli et al. 2015). A larger randomized trial is currently under way (NCT02938299).

Intralesional Oncolytic Immunotherapy and the Abscopal Effect Many of the local ablative therapies mentioned above, as well as electrochemotherapy described below, have been shown to induce responses in non-injected regional and distant metastatic sites, a so-called bystander or “abscopal” effect. These systemic responses are likely immunologically based, with evidence demonstrating that the oncolysis produced at the injection sites elicits a robust cytokine and cellular inflammatory response. The inflammatory response in the context of now-exposed, non-denatured tumor antigens can promote the priming, maturation, and expansion of tumor-specific CD8-T cells. The circulation of these T cells and their subsequent migration into metastatic tumor deposits are at least in part responsible for the observed clinical responses in un-injected metastases; this is the basis for classifying these agents and strategies as “oncolytic immunotherapy.” The multiple steps that lead to this systemic response have been described as part of a hypothetical “cancer immunity cycle.” In this hypothetical model (see Fig. 2), checkpoint blocking agents (anti-CTLA-4 and anti PD(L)-1 in particular) may augment the immunological response produced by the oncolytic therapy. This potential synergism has provided the basis for testing systemically administered checkpoint blockade in combination with locally administered intralesional therapy in ongoing clinical trials. Such a strategy is particularly attractive in the management of patients with stage III unresectable in-transit disease as it may not only effectively control the clinically apparent regional disease but also the systemic micrometastatic disease that is the source of subsequent distant failure, commonplace in this patient population.

880

J. F. Thompson et al.

Oncolytic Immunotherapy Designed to Produce Local and Systemic Effects STEP 2 Uptake, process, and presentation of tumor antigens by APCs

STEP 1 Tumor cell lysis and release of tumor-derived antigens

Anti-CTLA-4

STEP 3 • T-cell priming and activation • Generation of memory T cells

The Cancer–Immunity Cycle

STEP 7 • Killing of tumor cells • Memory-mediated control of tumor cell recurrence

Anti-PD-1

STEP 6 T-cell recognition of tumor cells

STEP 4 Travel of activated T cells to tumors

STEP 5 T-cell infiltration into tumors

APC, antigen-presenting cell Chen DS et al. Immunity. 2013;39:1-10.

Fig. 2 Oncolytic immunotherapy figure

Laser and Light-Based Therapies Laser and light-based therapies have occasionally been used for patients with metastatic melanoma, primarily as palliative care options. A systematic review of these therapies identified 27 articles in which ablative laser therapy (n = 10), non-ablative laser therapy (n = 9), and photodynamic therapy (n = 8) were used (Austin et al. 2017).

Ablative Laser Therapy Ablative lasers include non-fractionated and fractionated CO2 lasers. Non-fractionated, 10,600 nm CO2 laser destroys melanoma tissue through water vaporization. Nine case series of 225 patients in total showed highly variable response rates, and there were insufficient details to be confident about efficacy in treated lesions or effects on mortality (Austin et al. 2017). Fractionated CO2 laser divides the beam of energy into many microbeams; this preserves intervening tissue so it can act as a healing reservoir. However, melanoma cells are left intact in these intervening regions. One case series of three patients reported

a good response in all three, although two of the patients had further recurrences that were again treated with laser therapy and all three were alive at follow-up after 6–9 years (Chan and Quaba 2012). Additional studies are required if the effects of this treatment are to be better understood.

Non-ablative Laser Therapy Non-ablative lasers include ruby laser, pulsed dye laser, neodymium laser, and near-infrared diode laser. Ruby lasers generate 694 nm wavelength light that is absorbed by melanin and leads to destruction of melanoma cells. In two case reports, each of three patients, one patient had a complete response (McGuff et al. 1966), while two had partial responses but with regrowth of the lesions after treatment (Helsper et al. 1964). The ruby laser was developed in the 1960s and has been largely replaced with other types of laser. Pulsed dye laser, which generates 585–595 nm wavelength light, is sometimes used to treat vascular lesions. In a case series of ten melanoma patients, nine had a complete or partial response in their treated skin lesions, although eight went on

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

to develop disease in lymph nodes or at distant sites (Kottschade et al. 2010). A case report of pulsed dye laser in combination with topical imiquimod cream reported a complete clearance of the melanoma lesions at 28 weeks of treatment and no recurrence 6 months after completion of treatment (Zeitouni et al. 2005). Pulsed dye laser may be helpful in reducing morbidity from recurrent lesions, but there is insufficient evidence to determine its true efficacy. Neodymium laser, at 1060 nm wavelength, creates an area of coagulation necrosis which is thought to prevent melanoma cells from migrating into blood vessels. In two case series of 58 patients (Kozlov and Moskalik 1980) and 18 patients (Wagner et al. 1975), good responses were reported, with no recurrences of the treated skin lesions during follow-up of 3–78 months and 2–30 months, respectively. Given the limited evidence available, none of it recent, more data are required to establish the efficacy of this treatment. Near-infrared diode laser generates a wavelength of 805 nm which is absorbed by locally injected indocyanine green dye, generating heat and destroying melanoma tissue. Three clinical series involving 14 patients all reported a good response, with resolution of skin lesions although with frequent adverse events. Studies included patients receiving concurrent systemic therapies, so assessing the contributions of the laser treatment and the systemic therapies to the resolution and the adverse effects was not possible (Austin et al. 2017).

Radio-Frequency Ablation Radio-frequency ablation is a treatment that uses a high-frequency electrical current to selectively heat and destroy tissue. The process involves insertion of a partly insulated active electrode into the lesion, while a second larger area electrode is placed in a body location with good thermal and electrical conductivity, usually the back or thigh. A high-frequency current, but normally below 1 MHz, is established between the two electrodes causing frictional heating. The heating is greatest in the smaller electrode due to the

881

current density caused by the difference in surface area between the two electrodes. The effective temperature is between 50  C and 100  C (Shashank et al. 2014). Limitations to the use of radio-frequency ablation include proximity of the lesion of concern to a large blood vessel, which tends to dissipate the heat, and when the tumor is large requiring overlapping areas of ablation, which may result in some areas being missed. The use of radio-frequency ablation has also been explored in laparoscopic and open surgical procedures, including melanoma lung metastases (Schumacher et al. 2007; Eisele et al. 2009; Siperstein et al. 2000). The results of the very few studies of the use of radio-frequency ablation in melanoma patients suggest that the procedure is a potentially useful treatment for melanoma metastases (Shashank et al. 2014).

Photodynamic Therapy Photodynamic therapy uses chemicals that are preferentially absorbed by actively metabolizing cells and are photosensitive, so that exposure to light induces cell death by the release of reactive oxygen molecules. Photosensitive chemicals used for this purpose include aminolevulinic acid, hematoporphyrin derivative, Chlorin e6 (phytochlorine), and methylene blue. Eight clinical studies of 29 patients have been reported (Austin et al. 2017). Five studies involving 13 patients used hematoporphyrin derivative, 1 used Chlorin e6 (in 14 patients), 1 used methylene blue (1 patient), and 1 used aminolevulinic acid (1 patient). In the studies using hematoporphyrin derivative, 11 of 13 patients showed some response to treatment, but side effects were frequent and included pain, photosensitivity, leukocytosis, elevated transaminases, tachyarrhythmia, and wound infection (Austin et al. 2017). In the single study evaluating Chlorin e6, all 14 patients had a complete response, with no recurrences during 6–24 months of follow-up (Sheleg et al. 2004). The treatment did not appear to impact on progression of disease, however, and 11 of the patients died from distant, non-cutaneous metastases. Mild pain appeared to be the only adverse

882

effect. The case report using aminolevulinic acid in one patient reported no response at 3 months in any of the eight lesions treated (Wolf et al. 1993). The single case report of treatment with methylene blue reported that five of six lesions in a single patient demonstrated a complete response a few months after treatment (Tardivo et al. 2004). From these studies it appears that photodynamic therapy with Chlorin e6 and methylene blue may be worth exploring further in well-designed randomized trials to establish efficacy for treating recurrent cutaneous melanoma metastases. Overall, laser and light-based therapies for treating recurrent localized skin lesions in melanoma patients show considerable promise in reducing the size of these lesions, with few adverse events, but further studies, ideally randomized trials of alternative therapies with welldefined outcomes and time frames, are needed before there is sufficient certainty for them to be included in standard clinical practice.

Electrochemotherapy Electrochemotherapy (ECT) is a tumor ablation modality involving the local application of short duration, high-voltage pulses that leads to destabilization of the cell membrane, transiently increasing the permeability to hydrophilic cytotoxic chemotherapeutic drugs such as bleomycin or cisplatin (Mir et al. 1991; Sersa et al. 2009). Secondary effects of ECT include a reduction of tumor blood flow, inducing drug entrapment and localized vascular disruption, resulting in dramatically improved effectiveness of the drug (Gehl et al. 2002; Sersa et al. 1999). The permeabilizing effect is restricted to an area encompassed by the electrodes. To date, regular use of this treatment modality has largely been limited to continental Europe. The European Standard Operating Procedure for ECT (ESOPE) outlines the role of this treatment in various clinical scenarios, specifies indications and contraindications, and describes the technical details (Marty et al. 2006; Mir et al. 2006). ECT equipment includes an electric pulse generator and disposable electrodes with different types of

J. F. Thompson et al.

needles or plaques, suitable for various applications. Figure 3 shows a clinical electroporation device and electrodes (Cliniporator™ Igea SpA, Italy). The procedure involves the insertion of electrodes into the tumor and surrounding tissue so the entire tumor lies within the electrical field generated. Alternatively, noninvasive plate electrodes are placed on the skin surface to encompass the entire tumor mass. Tumors on or in the skin are generally easily accessible for the application of electric pulses. ECT can be a safe and well-tolerated treatment for superficial metastases and may be particularly useful when patients are too frail for other treatments due to their age and comorbidities (Campana et al. 2014). It can also be an efficient palliative treatment for bleeding metastases (Gehl and Geertsen 2000; Jarm et al. 2010). ECT can be performed under local, spinal, regional, or general anesthesia depending on the number and size of lesions to be treated. Because of the muscular contraction caused by the electric pulses, compliance of the patient under local anesthesia is limited to no more than 2–4 pulse sequences, thus limiting ECT applicability in this setting. In the majority of cases, post-procedure pain is mild, but a few patients can experience severe pain. Risk factors for post-electrochemotherapy pain are large tumors, moderate or severe pain before treatment, previous irradiation, and application of high electrical currents (>5 amps) (Quaglino et al. 2015). ECT can readily be repeated if progression occurs or if tumor remains (Campana et al. 2012, 2009; Testori et al. 2010). Both bleomycin and cisplatin have been used for ECT. Bleomycin can be administered either intravenously or intralesionally for ECT but cisplatin only intralesionally. The cytotoxic effect of bleomycin is potentiated in vitro approximately 8000 times and that of cisplatin approximately 80 times (Kis et al. 2011). Bleomycin is today the most commonly used agent for ECT. It is a very potent cytotoxic molecule when introduced into a cell. A few hundred molecules are sufficient for the drug to be cytotoxic, inducing single- and double-strand DNA breaks; however it does not freely diffuse through

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

883

Fig. 3 Clinical electroporation device (Cliniporator™ Igea SpA, Italy) and disposable electrodes with various types of needles or plaques. An image of ECT treatment is also shown. (Use of figure courtesy of IGEA)

the plasma membrane (Rols et al. 2000). It is therefore an excellent candidate for combining with electric pulses because it is non-permeant and at the same time highly cytotoxic inside the cell. When intravenous bleomycin is used, a bolus is administered at a dose of 15,000 IU/m2 8 min before the electroporation procedure to allow tissue distribution of the drug. A previous recommendation to end treatment at 20 min (28 min after completion of bleomycin administration) is now expanded to a treatment window of 40 min, based on clinical experience (Gehl et al. 2018). When intratumoral bleomycin injection is used, the electric pulses must be applied within 10 min. Prospective randomized trials in melanoma patients have shown that ECT is significantly

more effective than the drug alone in the treatment of metastatic tumor nodules, more than doubling the OR rate (Table 1) (Sersa et al. 2000; Byrne et al. 2005; Heller et al. 1998; Glass et al. 1996).

Reported Results of ECT Sixteen studies have reported results for 106 melanoma patients treated with ECT; 11 of these used bleomycin, while 5 used cisplatin. The average overall response rate in the bleomycin-treated patients was 89% and for the cisplatin-treated patients 83% (Campana et al. 2014). A larger study of 61 patients, 32 of whom had melanoma, was treated with electrochemotherapy with

884

bleomycin or cisplatin. The overall response rate in melanoma tumors was 80.6% and the complete response rate 66.3% (Marty et al. 2006). This study served to standardize the technique of ECT, and many subsequent studies have reported findings using this procedure for melanoma patients (Quaglino et al. 2008; Campana et al. 2012; 2009; 2014; 2012; Kis et al. 2011; Mozzillo et al. 2012). Pooled results from 5 of these studies, that included 161 melanoma patients, gave an average overall response rate of 89% with an average complete response rate of 56%. The largest of these studies (Campana et al. 2012) explored predictors of treatment response and reported that larger tumors were less responsive, while fewer electrode application cycles and lesions located on a limb as opposed to trunk were associated with a greater response to treatment. A more recent multicenter cohort study of 114 melanoma patients treated with ECT using bleomycin and followed for at least 60 days reported that 84 (74%) experienced a complete or partial response (Kunte et al. 2017). Multivariate analysis for predictors of overall response identified coverage of deep margins, treatment of nonirradiated areas, presence of lymphedema, and absence of visceral metastases as being associated with response to treatment. No serious adverse events were reported, and treatment was well tolerated. Caracò et al. (2015) reported that in 60 patients with in-transit disease or distant cutaneous

Fig. 4 Panel A: Metastatic tumor appeared in the ilioinguinal node dissection scar of a patient previously treated for cutaneous melanoma of the lower limb. The metastasis appeared during follow-up and was

J. F. Thompson et al.

metastases treated with ECT, only 13.3% had no change or progressive disease, and 44.8% of those who achieved a CR experienced a long-lasting response after one ECT session and were disease free after a mean follow-up of 27.5 months (Caracò et al. 2013). Figure 4 shows a typical patient treated with ECT and the clinical outcome. After the initial report by Rudolf et al. (1995) of an overall response rate of 92% in 24 metastatic lesions treated with intralesional drug, Glass et al. demonstrated similar results using intravenous bleomycin injection (Glass et al. 1996). Heller et al. (1998) in a larger cohort of patients with 84 melanoma metastases showed an OR of 99% with a CR of 89%; Mir et al. (1998) in 20 patients with 142 metastatic lesions reported an OR rate of 92% with a CR rate of 53% and with a similar range of response rates reported in other studies (Möller et al. 2009; Snoj et al. 2009; Kaehler et al. 2010; Quaglino et al. 2011; Mali et al. 2013; Queirolo et al. 2014; Spratt et al. 2014; Mir-Bonafé et al. 2015; Aguado-Romeo et al. 2017; Testori et al. 2017; Al-Hadithy et al. 2018).

ECT in Perspective Based on a large clinical experience, with consistently high response rates across a wide range of studies, ECT can be considered an effective alternative to other local and locoregional treatments

immediately superficial to the femoral vessels. Panel B: Complete resolution of the tumor 6 weeks after ECT. (Use of figure courtesy of IGEA)

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

for the local control of unresectable cutaneous and subcutaneous metastases in melanoma patients (Mali et al. 2013). This is reflected in several melanoma management guidelines: ESMO (European Society of Medical Oncology) (Dummer et al. 2015), Age.Na.S. (Agenzia Nazionale per i Servizi Sanitari Regionali) (http://www.salute.gov.it/ imgs/C_17_downloadMiaPelle_3_allegati_item Allegati_0_allegato.pdf), NICE (National Institute for Health and Care Excellence) (http:// www.nice.org.uk/Search.do?x=0&y=0&search Text=electrochemotherapy&newsearch=true#/ search/?reload, 2013, http://www.nice.org.uk/ guidance/ng14, 2015), ECCO (European CanCer Organization) (Beets et al. 2017), AIOM (Associazione Italiana di Oncologia Medica) (http://www.aiom.it/professionisti/documenti-sci entifici/linee-guida/1,413,1,#TopList), EDF, EADO, EORTC European consensus-based interdiscipli nary guideline (Garbe et al. 2016), and (DDG) Deutsche Dermatologischen Gesellschaft (http:// www.awmf.org/uploads/tx_szleitlinien/032-02 4l_S3_Melanom_Diagnostik_Therapie_Nachsor ge_2013-02.pdf).

Regional Therapies Patients treated with the local therapies discussed above have a high rate of subsequent local recurrence. The optimal therapy for these patients would be one that resulted in durable long-term remission of not only clinical but also subclinical regional disease. Regional therapies that aim to achieve this goal are outlined below but are discussed in more detail in other chapters (see chapters ▶ “Local and Recurrent Regional Metastases of Melanoma,” ▶ “Radiotherapy for Distant Melanoma Metastases,” ▶ “Hyperthermic Regional Perfusion for Melanoma of the Limbs,” and ▶ “Isolated Limb Infusion for Melanoma”).

Isolated Limb Perfusion and Infusion Isolated limb perfusion (ILP) is a technique that was introduced by Creech et al. in the 1950s (Creech et al. 1961). It involves delivery of high-

885

dose chemotherapy drugs at elevated temperature (up to 42 oC) directly to a limb without causing serious systemic side effects. The flow of blood to and from the tumor-bearing limb is temporarily interrupted, and an extracorporeal circuit is established via large bore catheters inserted surgically into the axial artery and vein of the limb. Drugs are then introduced into the extracorporeal circuit and circulated through the limb. Isolated limb infusion (ILI) is an alternative technique developed by Thompson et al. in the 1980s (Thompson et al. 1994). It is less complex and uses a minimally invasive percutaneous approach to insert small caliber catheters into the limb vessels prior to isolation of the limb from the body with a pneumatic tourniquet. This procedure is well tolerated, even by frail and elderly patients, allowing effective regional treatment to be offered even to individuals who might not be considered fit for treatment by ILP. ILP differs from ILI in that the isolated limb circulation is connected to an extracorporeal heart-lung machine, so the chemotherapy is administered to the limb while maintaining physiologic oxygenation and pH. During an ILI there is no oxygenator in the circuit, so progressive hypoxia and acidosis develop. However, it has been shown that the cytotoxic effect of melphalan is potentiated under these nonphysiological conditions. The ILP and ILI techniques are useful for patients with multiple local and in-transit metastases too numerous or too large to excise. The drug L-phenylalanine mustard (melphalan, LPAM) is most commonly used, and dosage is based on the volume of the extremity, usually calculated by displacement from a water reservoir prior to the procedure. A standard dose is 10 mg/L of perfused tissue for the lower limb and 13 mg/L for an upper limb. Other single agents or combinations of drugs have been used, including cisplatin, vindesine, dacarbazine, fotemustine, interleukin-2, and lymphokine-activated killer cells. A systematic review of trials and observational studies in which patients are treated with isolated limb perfusion reported a complete response rate of 58.2% and a median overall response rate of 90% (Moreno-Ramirez et al. 2010). A summary

886

of seven observational studies of ILI reported an overall response rate of 73% (Kroon et al. 2014). However, results from the two procedures are not directly comparable because the patient populations treated using the two procedures have tended to differ quite markedly (Kroon and Thompson 2018). No randomized trials directly comparing the efficacy of ILP and ILI have been conducted to date. Unfortunately, the results achieved by ILP and ILI are not durable in many cases. Analysis of ten ILP studies gave a median recurrence rate of 40.5% with a median time to recurrence after treatment of 10.5 months (Moreno-Ramirez et al. 2010). However, these treatments may improve quality of life for a considerable time in treated patients, many of them elderly, with multiple painful lesions confined to a single limb who may not be considered suitable for treatment with new systemic treatments. Further details on isolated limb perfusion and infusion are provided in other chapters (see chapters. ▶ “Hyperthermic Regional Perfusion for Melanoma of the Limbs,” ▶ “Isolated Limb Infusion for Melanoma,” and ▶ “Local and Recurrent Regional Metastases of Melanoma”).

Radiation Therapy for Locally Recurrent Melanoma and In-transit Metastases Radiation therapy kills tumor cells by causing double-stranded breaks in the DNA within them. Normal cells can repair DNA breaks, but tumor cells usually cannot do so and therefore are unable to survive and divide. Radiation therapy is useful in situations where normal tissue conservation is important. Involution of nonirradiated tumors after localized radiation therapy, an “abscopal” effect, is sometimes observed in patients with metastatic melanoma (Postow et al. 2012; Stamell et al. 2013). Techniques for the delivery of radiation have become increasingly sophisticated in recent times, meaning that therapy can be delivered more precisely, maximizing the dose to specific lesions and minimizing the impact on surrounding normal tissue. Indications for the adjuvant use of radiation therapy for melanoma include close or positive surgical margins where

J. F. Thompson et al.

further surgery is not possible, when there is lymphatic system involvement, and for tumors likely to have multiple, frequent local recurrences such as desmoplastic melanomas, particularly those with neurotropism (Hong and Fogarty 2012). It is unclear whether radiation therapy is beneficial after wide excision of neurotropic melanomas of the head and neck or a prospective randomized trial is currently under way, with expected completion of accrual in late 2019 (NCT00975520).

Amputation When all other treatment options have failed to control seriously symptomatic locoregionally recurrent melanoma, the only remaining option may be amputation. Although it is usually regarded as a purely palliative procedure offered as a last resort, long-term disease-free survival is achieved in up to one third of patients. In a review of 9 studies including 325 patients treated by major amputation for uncontrollable locoregional melanoma recurrence, 5-year survival ranged from 16% to 35% (Table 2). Only three studies reported median survival times postamputation, and these ranged from 5 to 13 months. Eight postoperative deaths were reported (3.5%) (Jaques et al. 1989; Ebskov 1996; Kroon et al. 2009; Turnbull et al. 1973; Cox 1974; Kourtesis et al. 1983; McPeak et al. 1963; Read et al. 2015b). In most cases amputation followed extensive prior treatments including radiotherapy, ILP and ILI, and local and systemic chemotherapy, yet the disease was persistent, painful, and uncontrollable. In these patients amputation provided pain relief and in some patients durable disease control, with prolonged disease-free survival sometimes achieved, as mentioned above.

Systemic Therapy With the recent introduction of frequently effective targeted and immune therapies to treat patients with systemic melanoma metastases, patients with difficult to manage, recurrent locoregional disease may be considered for systemic therapy. Options for systemic therapy are reviewed in detail in other chapters (see chapters

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . .

887

Table 2 Prospective randomized trials of ECT in melanoma patients. Effectiveness of ECT Studies Gaudy et al. 2006 Byrne et al. 2005 Sersa et al. 1999 Heller et al. 1998 Glass et al. 1996

Study type Lesion randomized controlled study Phase 2 trial: lesion randomized Phase 2 trial: lesion randomized Lesion randomized controlled study Phase 2 trial: lesion randomized

Drug i.t. BLM i.t. BLM i.t. CDDP i.t. BLM i.t. BLM

▶ “Evolving Role of Chemotherapy-Based Treatment of Metastatic Melanoma,” ▶ “Targeted Therapies for BRAF-Mutant Metastatic Melanoma” ▶ “Molecularly Targeted Therapy for Patients with BRAF Wild-Type Melanoma,” ▶ “Checkpoint Inhibitors in the Treatment of Metastatic Melanoma,” and ▶ “Novel Immunotherapies and Novel Combinations of Immunotherapy for Metastatic Melanoma”).

Adjuvant and Neoadjuvant Therapies Adjuvant Therapies For patients who develop locally recurrent melanoma or in-transit metastases and who are subsequently rendered free of disease by any of the local or regionally administered treatment modalities that have been discussed, it may be reasonable to consider the administration of adjuvant targeted or immune systemic therapies in an attempt to diminish the risk of subsequent regional and/or systemic relapse. A detailed discussion of the risks and benefits of the available adjuvant therapies can be found in the chapter ▶ “Adjuvant Systemic Therapy for High-Risk Melanoma Patients.”

Neoadjuvant Therapies Alternatively, neoadjuvant strategies are currently being studied as a compelling approach to patients with locoregional soft tissue metastases without demonstrable evidence of distant disease. For

Number of patients/ histology 12/skin metastasis on melanoma 19/skin metastasis on melanoma 10/skin metastasis on melanoma 34/melanoma, Kaposi, BCC, SCC 5/skin metastasis on melanoma

ORR lesions: drug alone vs. ECT 54% vs. 82% 32% vs. 78% 38% vs. 78%

0% vs. 95%

patients with a limited number of metastatic sites at any anatomic location that can be easily resected, a classic neoadjuvant approach using effective systemic therapies upfront, either alone or in combination with intralesional oncolytic agents, is followed by surgical resection. Patients with multiple and/or bulky unresectable extremity disease present a challenging clinical problem but at the same time offer a unique opportunity to integrate the most effective locoregional and systemic therapies available within a modified neoadjuvant approach, where isolated limb infusion is the consolidating treatment modality instead of surgical resection. A multicenter trial is under development that will include upfront treatment with a combination of intralesional T-VEC in combination with systemic pembrolizumab for 12 weeks. The patients who do not achieve a complete clinical response will then undergo isolated limb infusion followed by adjuvant single-agent pembrolizumab. This strategy was designed with the aim of improving distant disease-free survival outcomes as well as enhancing the response to and durability of isolated limb infusion.

Conclusions The most frequent site of initial failure after definitive primary treatment of melanoma is locoregional, and it is often the harbinger of subsequent systemic disease. This probably reflects the inherent biological aggressiveness of the initial primary tumor rather than being a direct sequel of the locoregional recurrence itself. Despite the

888

multitude of both local and regional treatment options available, many of them remarkably effective, all locoregional treatment strategies in these patients should be tempered by the understanding that this is a group of patients at high risk of subsequent locoregional recurrence and of systemic metastasis. The long-term results of adjuvant and neoadjuvant systemic treatment protocols in these patients are therefore awaited with considerable interest, while early results of combined local and systemic therapies are encouraging.

References Aguado-Romeo MJ, Benot-López S, Romero-Tabares A (2017) Electrochemotherapy for the treatment of unresectable locoregionally advanced cutaneous melanoma: a systematic review. Actas Dermosifiliogr (English Edition) 108(2):91–97. https://doi.org/ 10.1016/j.adengl.2016.12.002 Al-Hadithy N, Dehnel A, George A, Kisiel R, Lunt C, Stone C (2018) Patient reported outcomes in prospective cohort study of electrochemotherapy. Int J Surg 52:110–119. https://doi.org/10.1016/j.ijsu.2018.02.040 Andtbacka R, Curti BD, Hallmeyer S, Feng Z, Paustian C, Bifulco C, Fox B, Grose M, Shafren D (2015a) Phase II calm extension study; Coxsackie virus A21 delivered intratumorally to patients with advanced melanoma induces immune-cell infiltration in the tumor microenvironment. J Immunother Cancer 3(Suppl 2):P343 Andtbacka R, Kaufman HL, Collichio F, Amatruda T, Senzer N, Chesney J, Delman KA, Spitler LE, Puzanov I, Agarwala SS (2015b) Talimogene laherparepvec improves durable response rate in patients with advanced melanoma. J Clin Oncol 33(25):2780–2788 Austin E, Mamalis A, Ho D, Jagdeo J (2017) Laser and light-based therapy for cutaneous and soft-tissue metastases of malignant melanoma: a systematic review. Arch Dermatol Res 309(4):229–242. https://doi.org/ 10.1007/s00403-017-1720-9 Balch CM (2009) Microscopic satellites around a primary melanoma: another piece of the puzzle in melanoma staging. Ann Surg Oncol 16(5):1092–1094 Balch CM, Buzaid AC, Soong S-J, Atkins MB, Cascinelli N, Coit DG, Fleming ID, Gershenwald JE, Houghton A Jr, Kirkwood JM (2001) Final version of the American Joint Committee on Cancer staging system for cutaneous melanoma. J Clin Oncol 19(16):3635–3648 Bauer R, Kopald K, Lee J, Wong J, Morton D (1990) Longterm results of intralesional BCG for locally advanced recurrent melanoma. Proc Am Soc Clin Oncol 9:276 Beets G, Sebag-Montefiore D, Andritsch E, Arnold D, Beishon M, Crul M, Dekker JW, Delgado-Bolton R, Fléjou J-F, Grisold W, Henning G, Laghi A, Lovey J, Negrouk A, Pereira P, Roca P, Saarto T, Seufferlein T,

J. F. Thompson et al. Taylor C, Ugolini G, Cvd V, Bv H, Yared W, Costa A, Naredi P (2017) ECCO essential requirements for quality cancer care: colorectal cancer. A critical review. Crit Rev Oncol Hematol 110:81–93. https://doi.org/ 10.1016/j.critrevonc.2016.12.001 Berger AC, Ollila DW, Christopher A, Kairys JC, Mastrangelo MJ, Feeney K, Dabbish N, Leiby B, Frank JA, Stitzenberg KB, Meyers MO (2017) Patient symptoms are the most frequent indicators of recurrence in patients with American Joint Committee on Cancer Stage II Melanoma. J Am Coll Surg 224(4):652-659. https://doi.org/10.1016/j.jamcollsurg.2016.12.038 Breitbart EW (1990) Cryosurgery in the treatment of cutaneous malignant melanoma. Clin Dermatol 8 (1):96–100 Byers BA, Temple-Oberle CF, Hurdle V, McKinnon JG (2014) Treatment of in-transit melanoma with intralesional interleukin-2: a systematic review. J Surg Oncol 110(6):770–775. https://doi.org/10.1002/jso.23702 Byrne CM, Thompson JF, Johnston H, Hersey P, Quinn MJ, Michael Hughes T, McCarthy WH (2005) Treatment of metastatic melanoma using electroporation therapy with bleomycin (electrochemotherapy). Melanoma Res 15(1):45–51 Calabro A, Singletary SE, Balch CM (1989) Patterns of relapse in 1001 consecutive patients with melanoma nodal metastases. Arch Surg 124(9):1051–1055 Campana LG, Mocellin S, Basso M, Puccetti O, De Salvo GL, Chiarion-Sileni V, Vecchiato A, Corti L, Rossi CR, Nitti D (2009) Bleomycin-based electrochemotherapy: clinical outcome from a single institution’s experience with 52 patients. Ann Surg Oncol 16(1):191–199. https://doi.org/10.1245/s10434-008-0204-8 Campana L, Valpione S, Mocellin S, Sundararajan R, Granziera E, Sartore L, Chiarion-Sileni V, Rossi C (2012) Electrochemotherapy for disseminated superficial metastases from malignant melanoma. Br J Surg 99(6):821–830 Campana LG, Testori A, Mozzillo N, Rossi CR (2014) Treatment of metastatic melanoma with electrochemotherapy. J Surg Oncol 109(4):301–307 Caracò C, Mozzillo N, Marone U, Simeone E, Benedetto L, Di Monta G, Di Cecilia ML, Botti G, Ascierto PA (2013) Long-lasting response to electrochemotherapy in melanoma patients with cutaneous metastasis. BMC Cancer 13(1). https://doi.org/10.1186/1471-2407-13-564 Caracò C, Marone U, Simeone E, Grimaldi AM, Botti G, Del Giudice M, Ascierto PA, Mozzillo N (2015) Electrochemotherapy in melanoma patients: a single institution experience. Melanoma Manag 2 (2):127–132. https://doi.org/10.2217/mmt.15.9 Cascinelli N, Bufalino R, Marolda R, Belli F, Nava M, Galluzzo D, Santinami M, Levene A (1986) Regional non-nodal metastases of cutaneous melanoma. Eur J Surg Oncol 12(2):175–180 Chan LK, Quaba AA (2012) Improving the quality of life in melanoma–the role of the CO2 laser. J Cosmet Laser Ther 14(1):43–47 Chesney J, Puzanov I, Collichio F, Singh P, Milhem MM, Glaspy J, Hamid O, Ross M, Friedlander P, Garbe C (2017) Randomized, open-label phase II study

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . . evaluating the efficacy and safety of talimogene laherparepvec in combination with ipilimumab versus ipilimumab alone in patients with advanced, unresectable melanoma. J Clin Oncol 36 (17):1658–1667 Clemente-Ruiz de Almiron A, Serrano-Ortega S (2012) Risk factors for in-transit metastasis in patients with cutaneous melanoma. Actas Dermosifiliogr 103 (3):207-213. https://doi.org/10.1016/j.ad.2011.06.002 Cohen MH, Jessup JM, Felix EL, Weese JL, Herberman RB (1978) Intralesional treatment of recurrent metastatic cutaneous malignant melanoma. A randomized prospective study of intralesional bacillus calmetteguerin versus intralesional dinitrochlorobenzene. Cancer 41(6):2456–2463 Cox K (1974) Survival after amputation for recurrent melanoma. Surg Gynecol Obstet 139(5):720 Creech O Jr, Ryan RF, Krementz ET (1961) Regional chemotherapy by isolated perfusion in the treatment of melanoma of the extremities. Plast Reconstr Surg Transplant Bull 28:333–346 Dalal KM, Patel A, Brady MS, Jaques DP, Coit DG (2007) Patterns of first-recurrence and post-recurrence survival in patients with primary cutaneous melanoma after sentinel lymph node biopsy. Ann Surg Oncol 14 (6):1934–1942 Damian DL, Saw RP, Thompson JF (2014) Topical immunotherapy with diphencyprone for in transit and cutaneously metastatic melanoma. J Surg Oncol 109 (4):308–313 Danielli R, Patuzzo R, Di Giacomo AM, Gallino G, Maurichi A, Di Florio A, Cutaia O, Lazzeri A, Fazio C, Miracco C (2015) Intralesional administration of L19-IL2/L19-TNF in stage III or stage IVM1a melanoma patients: results of a phase II study. Cancer Immunol Immunother 64(8):999–1009 Dong XD, Tyler D, Johnson JL, DeMatos P, Seigler HF (2000) Analysis of prognosis and disease progression after local recurrence of melanoma. Cancer 88 (5):1063–1071 Dummer R, Hauschild A, Lindenblatt N, Pentheroudakis G, Keilholz U (2015) Cutaneous melanoma: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann Oncol 26(suppl 5):v126–v132. https://doi.org/10.1093/annonc/mdv297 Ebskov L (1996) Relative mortality in lower limb amputees with diabetes mellitus. Prosthetics Orthot Int 20 (3):147–152 Eisele RM, Neumann U, Neuhaus P, Schumacher G (2009) Open surgical is superior to percutaneous access for radiofrequency ablation of hepatic metastases. World J Surg 33(4):804–811 Eklund JW, Kuzel TM (2004) A review of recent findings involving interleukin-2-based cancer therapy. Curr Opin Oncol 16(6):542–546 Elfatoiki FZ, Longvert C, Clerici T, Bourgault-Villada I, Roudier-Pujol C, Vasseur E, Saiag P (2014) In-transit metastasis in melanoma: efficacy of topical imiquimod combined with carbon dioxide laser or with electrocautery. Ann Dermatol Venereol 141(2):106–110. https:// doi.org/10.1016/j.annder.2013.10.058

889

Estourgie S, Nieweg O, Kroon B (2004) High incidence of in-transit metastases after sentinel node biopsy in patients with melanoma. Br J Surg 91(10):1370–1371 Faries MB, Thompson JF, Cochran A, Elashoff R, Glass EC, Mozzillo N, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Reintgen DS, Coventry BJ, Wang HJ, Morton DL, Group MC (2010) The impact on morbidity and length of stay of early versus delayed complete lymphadenectomy in melanoma: results of the Multicenter Selective Lymphadenectomy Trial (I). Ann Surg Oncol 17(12):3324–3329. https://doi.org/10.1245/ s10434-010-1203-0 Fleming I, Cooper J, Henson D, Hutter R, Kennedy B, Murphy G (1997) American Joint Committee on Cancer manual for staging of cancer. JB Lippincott, Philadelphia Florin V, Desmedt E, Vercambre-Darras S, Mortier L (2012) Topical treatment of cutaneous metastases of malignant melanoma using combined imiquimod and 5-fluorouracil. Investig New Drugs 30(4):1641–1645 Fusi S, Ariyan S, Sternlicht A (1993) Data on first recurrence after treatment for malignant melanoma in a large patient population. Plast Reconstr Surg 91(1):94–98 Gaudy C, Richard M-A, Folchetti G, Bonerandi JJ, Grob JJ (2006) Randomized controlled study of electrochemotherapy in the local treatment of skin metastases of melanoma. J Cutan Med Surg 10(3):115–121 Gehl J, Geertsen PF (2000) Efficient palliation of haemorrhaging malignant melanoma skin metastases by electrochemotherapy. Melanoma Res 10(6):585–589. https://doi.org/10.1097/00008390-200012000-00011 Gehl J, Skovsgaard T, Mir LM (2002) Vascular reactions to in vivo electroporation: characterization and consequences for drug and gene delivery. Biochim Biophys Acta Gen Subj 1569(1–3):51–58. https://doi.org/ 10.1016/s0304-4165(01)00233-1 Gehl J, Sersa G, Matthiessen LW, Muir T, Soden D, Occhini A, Quaglino P, Curatolo P, Campana LG, Kunte C, Clover AJP, Bertino G, Farricha V, Odili J, Dahlstrom K, Benazzo M, Mir LM (2018) Updated standard operating procedures for electrochemotherapy of cutaneous tumours and skin metastases. Acta Oncol 57(7):874–882. https://doi.org/10.1080/0284186x. 2018.1454602 Gershenwald JE, Fidler IJ (2002) Targeting lymphatic metastasis. Science 296(5574):1811–1812 Gershenwald JE, Scolyer RA, Hess KR, Sondak VK, Long GV, Ross MI, Lazar AJ, Faries MB, Kirkwood JM, McArthur GA, Haydu LE, Eggermont AMM, Flaherty KT, Balch CM, Thompson JF, For Members of the American Joint Committee on Cancer Melanoma Expert P, The International Melanoma D, Discovery P (2017) Melanoma staging: evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J Clin 67(6):472–492. https://doi.org/10.3322/caac.21409 Glass LF, Pepine ML, Genske NA, Jaroszeski MJ, Reintgen DS, Heller R (1996) Bleomycin-mediated electrochemotherapy of metastatic melanoma. Arch Dermatol 132(11):1353–1357. https://doi.org/ 10.1001/archderm.132.11.1353

890 Grady BE, Spencer JM (2016) Topical imiquimod clears invasive melanoma. Cutis 98(5):E28–E30 Griffiths RW, Briggs JC (1986) Incidence of locally metastatic (‘recurrent’) cutaneous malignant melanoma following conventional wide margin excisional surgery for invasive clinical stage I tumours: importance of maximal primary tumour thickness. Br J Surg 73 (5):349–353 Grotmann P, Ernst K, Hundeiker M (1991) Kryochirurgie bei multiplen kutanen Melanommetastasen. Z Hautkr 66(5):385–389 Heenan P, Ghaznawie M (1999) The pathogenesis of local recurrence of melanoma at the primary excision site. Br J Plast Surg 52(3):209–213 Heller R, Jaroszeski MJ, Reintgen DS, Puleo CA, DeConti RC, Gilbert RA, Glass LF (1998) Treatment of cutaneous and subcutaneous tumors with electrochemotherapy using intralesional bleomycin. Cancer 83(1):148–157. https://doi.org/10.1002/(sici)1097-0142(19980701)83:1 3.0.co;2-w Helsper JT, Sharp GS, Williams HF, Fister HW (1964) The biological effect of laser energy on human melanoma. Cancer 17(10):1299–1304 Hesling C, D'incan M, Mansard S, Franck F, Corbin-Duval A, Chèvenet C, Déchelotte P, Madelmont JC, Veyre A, Souteyrand P (2004) In vivo and in situ modulation of the expression of genes involved in metastasis and angiogenesis in a patient treated with topical imiquimod for melanoma skin metastases. Br J Dermatol 150(4):761–767 Hong A, Fogarty G (2012) Role of radiation therapy in cutaneous melanoma. Cancer J 18(2):203–207 Jaques D, Coit D, Brennan M (1989) Major amputation for advanced malignant melanoma. Surg Gynecol Obstet 169(1):1–6 Jarm T, Cemazar M, Miklavcic D, Sersa G (2010) Antivascular effects of electrochemotherapy: implications in treatment of bleeding metastases. Expert Rev Anticancer Ther 10(5):729–746. https://doi.org/10.1586/ era.10.43 Johannsen M, Spitaleri G, Curigliano G, Roigas J, Weikert S, Kempkensteffen C, Roemer A, Kloeters C, Rogalla P, Pecher G (2010) The tumour-targeting human L19IL2 immunocytokine: preclinical safety studies, phase I clinical trial in patients with solid tumours and expansion into patients with advanced renal cell carcinoma. Eur J Cancer 46(16):2926–2935 Joseph RW, Cappel M, Tzou K, Bagaria S, Gilstrap C, Swaika A, Jambusaria-Pahlajani A (2016) Treatment of in-transit and metastatic melanoma in two patients treated with ipilimumab and topical imiquimod. Melanoma Res 26(4):409–412. https://doi.org/10.1097/ CMR.0000000000000247 Kaehler KC, Egberts F, Hauschild A (2010) Electrochemotherapy in symptomatic melanoma skin metastases: intraindividual comparison with conventional surgery. Dermatol Surg 36(7):1200–1202. https://doi. org/10.1111/j.1524-4725.2010.01608.x Karakousis CP, Douglass HO, Yeracaris PM, Holyoke ED (1976) BCG immunotherapy in patients with malignant melanoma. Arch Surg 111(6):716–718

J. F. Thompson et al. Kibbi N, Ariyan S, Faries M, Choi JN (2015) Treatment of in-transit melanoma with intralesional bacillus CalmetteGuerin (BCG) and topical imiquimod 5% cream: a report of 3 cases. J Immunother 38(9):371–375. https:// doi.org/10.1097/CJI.0000000000000098 Kis E, Oláh J, Ócsai H, Baltás E, Gyulai R, Kemény L, Horvath AR (2011) Electrochemotherapy of cutaneous metastases of melanoma – a case series study and systematic review of the evidence. Dermatol Surg 37(6):816–824 Kottschade LA, Weenig RH, Otley CC, McWilliams RR, Markovic SN (2010) The use of pulsed dye laser in the treatment of melanoma metastatic to the skin: a Mayo Clinic case series. J Am Acad Dermatol 62(6):e22–e25 Kourtesis G, McCarthy W, Milton G (1983) Major amputations for melanoma. ANZ J Surg 53(3):241–244 Kozlov A, Moskalik K (1980) Pulsed laser radiation therapy of skin tumors. Cancer 46(10):2172–2178 Kroon HM, Thompson JF (2018) Isolated limb infusion and isolated limb perfusion for melanoma: can the outcomes of these procedures be compared? Ann Surg Oncol 26 (1):8–9. https://doi.org/10.1245/s10434-018-7067-4 Kroon HM, Lin DY, Kam PC, Thompson JF (2009) Major amputation for irresectable extremity melanoma after failure of isolated limb infusion. Ann Surg Oncol 16(6):1543–1547. https://doi.org/10.1245/s10434-0090394-8 Kroon HM, Huismans AM, Kam PC, Thompson JF (2014) Isolated limb infusion with melphalan and actinomycin D for melanoma: a systematic review. J Surg Oncol 109(4):348–351. https://doi.org/10.1002/jso.23553 Kunte C, Letulé V, Gehl J, Dahlstroem K, Curatolo P, Rotunno R, Muir T, Occhini A, Bertino G, Powell B, Saxinger W, Lechner G, Liew SH, Pritchard-Jones R, Rutkowski P, Zdzienicki M, Mowatt D, Sykes AJ, Orlando A, Mitsala G, Rossi CR, Campana L, Brizio M, de Terlizzi F, Quaglino P, Odili J (2017) Electrochemotherapy in the treatment of metastatic malignant melanoma: a prospective cohort study by InspECT. Br J Dermatol 176(6):1475–1485. https://doi.org/10.1111/ bjd.15340 Li X, Naylor MF, Le H, Nordquist RE, Teague TK, Howard CA, Murray C, Chen WR (2010) Clinical effects of in situ photoimmunotherapy on late-stage melanoma patients: a preliminary study. Cancer Biol Ther 10 (11):1081–1087 Liu B, Robinson M, Han Z, Branston R, English C, Reay P, McGrath Y, Thomas S, Thornton M, Bullock P (2003) ICP34. 5 deleted herpes simplex virus with enhanced oncolytic, immune stimulating, and anti-tumour properties. Gene Ther 10(4):292 Long GV, Dummer R, Ribas A, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, Olszanski AJ, Malvehy J, Cebon JS (2016) Efficacy analysis of MASTERKEY-265 phase 1b study of talimogene laherparepvec (T-VEC) and pembrolizumab (pembro) for unresectable stage IIIB-IV melanoma. American Society of Clinical Oncology New York. J Clin Oncol 34(15_suppl):9568–9568. Lyth J, Falk M, Maroti M, Eriksson H, Ingvar C (2017) Prognostic risk factors of first recurrence in patients

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . . with primary stages I-II cutaneous malignant melanoma – from the population-based Swedish melanoma register. J Eur Acad Dermatol Venereol 31 (9):1468–1474. https://doi.org/10.1111/jdv.14280 Mali B, Jarm T, Snoj M, Sersa G, Miklavcic D (2013) Antitumor effectiveness of electrochemotherapy: A systematic review and meta-analysis. Eur J Surg Oncol 39(1):4–16. https://doi.org/10.1016/j.ejso.2012. 08.016 Mansson-Brahme E, Carstensen J, Erhardt K, Lagerlof B, Ringborg U, Rutqvist LE (1994) Prognostic factors in thin cutaneous malignant melanoma. Cancer 73 (9):2324–2332 Marty M, Sersa G, Garbay JR, Gehl J, Collins CG, Snoj M, Billard V, Geertsen PF, Larkin JO, Miklavcic D (2006) Electrochemotherapy–An easy, highly effective and safe treatment of cutaneous and subcutaneous metastases: results of ESOPE (European Standard Operating Procedures of Electrochemotherapy) study. Eur J Cancer Suppl 4(11):3–13 Mastrangelo MJ, Bellet RE, Berkelhammer J, Clark WH Jr (1975) Regression of pulmonary metastatic disease associated with intralesional BCG therapy of intracutaneous melanoma metastases. Cancer 36(4):1305–1308 Matheson JA, Te Marvelde L, Mailer S, Speakman D, Spillane J, Henderson MA, Gyorki DE (2017) Prospective evaluation of prognostic indicators for early recurrence of cutaneous melanoma. Melanoma Res 27(1):43–49. https://doi. org/10.1097/CMR.0000000000000302 McCarthy WH (2002) Melanoma: margins for error–another view. ANZ J Surg 72(4):304–306 McGuff PE, Deterling RA, Gottlieb LS (1966) Laser radiation for metastatic malignant melanoma. JAMA 195(5):393–394 McPeak CJ, McNeer GP, Whiteley HW, Booher RJ (1963) Amputation for melanoma of the extremity. Surgery 54(3):426–431 Miller R, Gerster J, Owens M, Slade H, Tomai M (1999) Review article imiquimod applied topically: a novel immune response modifier and new class of drug. Int J Immunopharmacol 21(1):1–14 Mir LM, Orlowski S, Belehradek J, Paoletti C (1991) Electrochemotherapy potentiation of antitumour effect of bleomycin by local electric pulses. Eur J Cancer Clin Oncol 27(1):68–72. https://doi.org/10.1016/0277-5379 (91)90064-k Mir LM, Glass LF, Serša G, Teissié J, Domenge C, Miklavčič D, Jaroszeski MJ, Orlowski S, Reintgen DS, Rudolf Z, Belehradek M, Gilbert R, Rols MP, Belehradek J, Bachaud JM, DeConti R, Štabuc B, Čemažar M, Coninx P, Heller R (1998) Effective treatment of cutaneous and subcutaneous malignant tumours by electrochemotherapy. Br J Cancer 77 (12):2336–2342. https://doi.org/10.1038/bjc.1998.388 Mir LM, Gehl J, Sersa G, Collins CG, Garbay J-R, Billard V, Geertsen PF, Rudolf Z, O’Sullivan GC, Marty M (2006) Standard operating procedures of the electrochemotherapy: Instructions for the use of bleomycin or cisplatin administered either systemically or locally and electric pulses delivered by the CliniporatorTM by means of invasive or non-invasive electrodes. Eur J

891

Cancer Suppl 4(11):14–25. https://doi.org/10.1016/j. ejcsup.2006.08.003 Mir-Bonafé JM, Vilalta A, Alarcón I, Carrera C, Puig S, Malvehy J, Rull R, Bennàssar A (2015) Electrochemotherapy in the treatment of melanoma skin metastases: a report on 31 cases. Actas Dermosifiliogr (English Edition) 106(4):285–291. https://doi.org/ 10.1016/j.adengl.2015.03.004 Möller MG, Salwa S, Soden DM, O’Sullivan GC (2009) Electrochemotherapy as an adjunct or alternative to other treatments for unresectable or in-transit melanoma. Expert Rev Anticancer Ther 9(11):1611–1630. https://doi.org/10.1586/era.09.129 Moon SD, Spencer JM (2013) Clearance of invasive melanoma with topical imiquimod. J Drugs Dermatol 12(1):107–108 Moreno-Ramirez D, de la Cruz-Merino L, Ferrandiz L, Villegas-Portero R, Nieto-Garcia A (2010) Isolated limb perfusion for malignant melanoma: systematic review on effectiveness and safety. Oncologist 15(4):416–427. https://doi.org/10.1634/theoncologist. 2009-0325 Morton DL, Eilber FR, Holmes EC, Hunt JS, Ketcham AS, Silverstein MJ, Sparks FC (1974) BCG immunotherapy of malignant melanoma: summary of a seven-year experience. Ann Surg 180(4):635–643 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Elashoff R, Essner R, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP (2006) Sentinel-node biopsy or nodal observation in melanoma. N Engl J Med 355(13):1307–1317 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Puleo CA, Coventry BJ, Kashani-Sabet M, Smithers BM, Paul E, Kraybill WG, JG MK, Wang HJ, Elashoff R, Faries MB, Group M (2014) Final trial report of sentinel-node biopsy versus nodal observation in melanoma. N Engl J Med 370(7):599–609. https://doi.org/ 10.1056/NEJMoa1310460 Mozzillo N, Caracò C, Mori S, Di Monta G, Botti G, Ascierto PA, . . . Aloj L (2012) Use of neoadjuvant electrochemotherapy to treat a large metastatic lesion of the cheek in a patient with melanoma. J Transl Med, 10:131. https://doi.org/10.1186/1479-5876-10-131 Papadia F, Basso V, Patuzzo R, Maurichi A, Di Florio A, Zardi L, Ventura E, González-Iglesias R, Lovato V, Giovannoni L (2013) Isolated limb perfusion with the tumor-targeting human monoclonal antibody–cytokine fusion protein L19-TNF plus melphalan and mild hyperthermia in patients with locally advanced extremity melanoma. J Surg Oncol 107(2):173–179 Pawlik TM, Ross MI, Johnson MM, Schacherer CW, McClain DM, Mansfield PF, Lee JE, Cormier JN, Gershenwald JE (2005) Predictors and natural history of in-transit melanoma after sentinel lymphadenectomy. Ann Surg Oncol 12(8):587–596 Postow MA, Callahan MK, Barker CA, Yamada Y, Yuan J, Kitano S, Mu Z, Rasalan T, Adamow M, Ritter E (2012) Immunologic correlates of the abscopal effect in a patient with melanoma. N Engl J Med 366 (10):925–931

892 Pretto F, Elia G, Castioni N, Neri D (2014) Preclinical evaluation of IL2-based immunocytokines supports their use in combination with dacarbazine, paclitaxel and TNF-based immunotherapy. Cancer Immunol Immunother 63(9):901–910 Puzanov I, Milhem MM, Minor D, Hamid O, Li A, Chen L, Chastain M, Gorski KS, Anderson A, Chou J (2016) Talimogene laherparepvec in combination with ipilimumab in previously untreated, unresectable stage IIIB-IV melanoma. J Clin Oncol 34(22):2619–2626 Quaglino P, Mortera C, Osella-Abate S, Barberis M, Illengo M, Rissone M, Savoia P, Bernengo MG (2008) Electrochemotherapy with intravenous bleomycin in the local treatment of skin melanoma metastases. Ann Surg Oncol 15(8):2215–22. https://doi.org/10.1245/s10434008-9976-0 Quaglino P, Osella-Abate S, Marenco F, Nardò T, Gado C, Novelli M, Savoia P, Bernengo MG (2011) FoxP3 expression on melanoma cells is related to early visceral spreading in melanoma patients treated by electrochemotherapy. Pigment Cell Melanoma Res 24 (4):734–736. https://doi.org/10.1111/j.1755-148x. 2011.00879.x Quaglino P, Matthiessen LW, Curatolo P, Muir T, Bertino G, Kunte C, Odili J, Rotunno R, Humphreys AC, Letulé V, Marenco F, Cuthbert C, Albret R, Benazzo M, De Terlizzi F, Gehl J (2015) Predicting patients at risk for pain associated with electrochemotherapy. Acta Oncol 54(3):298–306. https://doi.org/10.3109/ 0284186x.2014.992546 Queirolo P, Marincola F, Spagnolo F (2014) Electrochemotherapy for the management of melanoma skin metastasis: a review of the literature and possible combinations with immunotherapy. Arch Dermatol Res 306(6):521–526. https://doi.org/10.1007/s00403-014-1462-x Read RL, Haydu L, Saw RP, Quinn MJ, Shannon K, Spillane AJ, Stretch JR, Scolyer RA, Thompson JF (2015a) In-transit melanoma metastases: incidence, prognosis, and the role of lymphadenectomy. Ann Surg Oncol 22(2):475–481. https://doi.org/10.1245/ s10434-014-4100-0 Read RL, Stalley P, Thompson JF (2015b) The contemporary role of major amputation in the management of advanced limb melanoma. Ann Surg Oncol 22 (12):4067–4072. https://doi.org/10.1245/s10434-0154416-4 Read T, Webber S, Tan J, Wagels M, Schaider H, Soyer H, Smithers B (2017a) Diphenylcyclopropenone for the treatment of cutaneous in-transit melanoma metastases–results of a prospective, non-randomized, single-centre study. J Eur Acad Dermatol Venereol 31(12):2030–2037 Read T, Webber S, Thomas J, Wagels M, Schaider H, Soyer HP, Smithers BM (2017b) Protocol for the TIDAL Melanoma Study: topical imiquimod or diphenylcyclopropenone for the management of cutaneous in-transit melanoma metastases-a phase II, single centre, randomised, pilot study. BMJ Open 7(10): e016816. https://doi.org/10.1136/bmjopen-2017-016816 Reintgen DS, Cox C, Slingluff JC, Seigler HF (1992) Recurrent malignant melanoma: the identification of

J. F. Thompson et al. prognostic factors to predict survival. Ann Plast Surg 28(1):45–49 Rivas-Tolosa N, Ortiz-Brugues A, Toledo-Pastrana T, Baradad M, Traves V, Soriano V, Sanmartin V, Requena C, Marti R, Nagore E (2016) Local cryosurgery and imiquimod: a successful combination for the treatment of locoregional cutaneous metastasis of melanoma: a case series. J Dermatol 43(5):553–556. https://doi.org/10.1111/1346-8138.13197 Robinson JC (1977) Risks of BCG intralesional therapy: an experience with melanoma. J Surg Oncol 9 (6):587–593 Rols MP, Bachaud JM, Giraud P, Chevreau C, Roché H, Teissié J (2000) Electrochemotherapy of cutaneous metastases in malignant melanoma. Melanoma Res 10(5):468–474. https://doi.org/10.1097/00008390200010000-00009 Rudolf Z, Stabuc B, Cemazar M, Miklavcic D, Vodovnik L, Sersa G (1995) Electrochemotherapy with bleomycin. The first clinical experience in malignant melanoma patients. Radiol Oncol 29:229–235 Schumacher G, Eisele R, Spinelli A, Schmidt SC, Jacob D, Pratschke J, Neuhaus P (2007) Indications for handassisted laparoscopic radiofrequency ablation for liver tumors. J Laparoendosc Adv Surg Tech A 17 (2):153–159 Sersa G, Cemazar M, Miklavcic D, Chaplin DJ (1999) Tumor blood flow modifying effect of electrochemotherapy with bleomycin. Anticancer Res 19(5B):4017–4022 Sersa G, Stabuc B, Cemazar M, Miklavcic D, Rudolf Z (2000) Electrochemotherapy with cisplatin: the systemic antitumour effectiveness of cisplatin can be potentiated locally by the application of electric pulses in the treatment of malignant melanoma skin metastases. Melanoma Res 10(4):381–5 Sersa G, Cemazar M, Snoj M (2009) Electrochemotherapy of tumours. Curr Oncol 16(2):34–35 Shashank A, Shehata M, Morris DL, Thompson JF (2014) Radiofrequency ablation in metastatic melanoma. J Surg Oncol 109(4):366–369 Sheleg SV, Zhavrid EA, Khodina TV, Kochubeev GA, Istomin YP, Chalov VN, Zhuravkin IN (2004) Photodynamic therapy with chlorin e6 for skin metastases of melanoma. Photodermatol Photoimmunol Photomed 20(1):21–26 Shi VY, Tran K, Patel F, Leventhal J, Konia T, Fung MA, Wilken R, Garcia MS, Fitzmaurice SD, Joo J, Monjazeb AM, Burrall BA, King B, Martinez S, Christensen SD, Maverakis E (2015) 100% Complete response rate in patients with cutaneous metastatic melanoma treated with intralesional interleukin (IL)-2, imiquimod, and topical retinoid combination therapy: results of a case series. J Am Acad Dermatol 73(4):645–654. https://doi. org/10.1016/j.jaad.2015.06.060 Singletary SE, Tucker SL, Boddie JRAW (1988) Multivariate analysis of prognostic factors in regional cutaneous metastases of extremity melanoma. Cancer 61 (7):1437–1440 Siperstein A, Garland A, Engle K, Rogers S, Berber E, String A, Foroutani A, Ryan T (2000) Laparoscopic

Local Melanoma Recurrence, Satellitosis, and In-transit Metastasis: Incidence, Outcomes. . . radiofrequency ablation of primary and metastatic liver tumors. Surg Endosc 14(4):400–405 Sisti A, Sisti G, Oranges CM (2014) Topical treatment of melanoma skin metastases with imiquimod: a review. Dermatol Online J 21(2):pii. 13030/qt8rj4k7r6 Sloot S, Speijers MJ, Bastiaannet E, Hoekstra HJ (2016) Is there a relation between type of primary melanoma treatment and the development of intralymphatic metastasis? A review of the literature. Cancer Treat Rev 45:120–128. https://doi.org/10.1016/j.ctrv.2016. 02.007 Snoj M, Cemazar M, Srnovrsnik T, Kosir SP, Sersa G (2009) Limb sparing treatment of bleeding melanoma recurrence by electrochemotherapy. Tumori J 95 (3):398–402. https://doi.org/10.1177/030089160909 500324 Soong SJ, Harrison RA, McCarthy WH, Urist MM, Balch CM (1998) Factors affecting survival following local, regional, or distant recurrence from localized melanoma. J Surg Oncol 67(4):228–233 Spratt DE, Gordon Spratt EA, Wu S, DeRosa A, Lee NY, Lacouture ME, Barker CA (2014) Efficacy of skindirected therapy for cutaneous metastases from advanced cancer: a meta-analysis. J Clin Oncol 32(28):3144–3155. https://doi.org/10.1200/jco.2014.55.4634 Stamell EF, Wolchok JD, Gnjatic S, Lee NY, Brownell I (2013) The abscopal effect associated with a systemic anti-melanoma immune response. Int J Radiat Oncol Biol Phys 85(2):293–295 Storm F, Sparks F, Morton D (1979) Treatment for melanoma of the lower extremity with intralesional injection of bacille Calmette Guérin and hyperthermic perfusion. Surg Gynecol Obstet 149(1):17–21 Tanaka S (1982) Immunological aspects of cryosurgery in general surgery. Cryobiology 19(3):247–262 Tardivo J, Del Giglio A, Paschoal L, Ito A, Baptista MS (2004) Treatment of melanoma lesions using methylene blue and RL50 light source. Photodiagn Photodyn Ther 1(4):345–346 Tarhini AA, Agarwala SS (2005) Interleukin-2 for the treatment of melanoma. Curr Opin Investig Drugs (London, England: 2000) 6(12):1234–1239 Temple-Oberle CF, Byers BA, Hurdle V, Fyfe A, McKinnon JG (2014) Intra-lesional interleukin-2 therapy for in transit melanoma. J Surg Oncol 109 (4):327–331. https://doi.org/10.1002/jso.23556 Testori A, Tosti G, Martinoli C, Spadola G, Cataldo F, Verrecchia F, Baldini F, Mosconi M, Soteldo J, Tedeschi I, Passoni C, Pari C, Di Pietro A, Ferrucci PF (2010) Electrochemotherapy for cutaneous and subcutaneous tumor lesions: a novel therapeutic approach. Dermatol Ther 23(6):651–661. https://doi.org/10.1111/ j.1529-8019.2010.01370.x Testori A, Ribero S, Bataille V (2017) Diagnosis and treatment of in-transit melanoma metastases. Eur J Surg Oncol 43(3):544–560. https://doi.org/10.1016/j. ejso.2016.10.005 Thai KE, Sinclair RD (1999) Cryosurgery of benign skin lesions. Australas J Dermatol 40(4):175–186 Thomas JM, Clark MA (2004) Selective lymphadenectomy in sentinel node-positive patients

893

may increase the risk of local/in-transit recurrence in malignant melanoma. Eur J Surg Oncol 30(6):686–691 Thompson JF, Waugh RC, Saw RPM, Kam PCA (1994) Isolated limb infusion with melphalan for recurrent limb melanoma: a simple alternative to isolated limb perfusion. Reg Cancer Treat 7:188–192 Thompson JF, Agarwala SS, Smithers BM, Ross MI, Scoggins CR, Coventry BJ, Neuhaus SJ, Minor DR, Singer JM, Wachter EA (2015) Phase 2 study of intralesional PV-10 in refractory metastatic melanoma. Ann Surg Oncol 22(7):2135–2142 Turnbull A, Shah J, Fortner J (1973) Recurrent melanoma of an extremity treated by major amputation. Arch Surg 106(4):496–498 Tyring S (2001) Imiquimod applied topically: a novel immune response modifier. Skin Therapy Lett 6(6):1–4 Vical (2013) A phase 3 pivotal trial comparing allovectin7® Alone vs chemotherapy alone in patients with stage 3 or stage 4 melanoma. https://clinicaltrials.gov/ct2/ show/NCT00395070. Accessed 6 Feb 2019 e2019 Wagner R, Kozlov A, Moskalik K, Khachaturyan L, Pertsov O (1975) Laser therapy of human benign and malignant neoplasms of the skin. Acta Radiol Ther Phys Biol 14(5):417–423 Wagner JD, Ranieri J, Evdokimow DZ, Logan T, Chuang T-Y, Johnson CS, Jung S-H, Wenck S (2003) Patterns of initial recurrence and prognosis after sentinel lymph node biopsy and selective lymphadenectomy for melanoma. Plast Reconstr Surg 112(2):486–497 Weide B, Eigentler TK, Pflugfelder A, Zelba H, Martens A, Pawelec G, Giovannoni L, Ruffini PA, Elia G, Neri D (2014) Intralesional treatment of stage III metastatic melanoma patients with L19–IL2 results in sustained clinical and systemic immunologic responses. Cancer Immunol Res 2(7):668–678 Wilmott J, Haydu L, Bagot M, Zhang Y, Jakrot V, McCarthy S, Lugassy C, Thompson J, Scolyer R, Barnhill R (2012) Angiotropism is an independent predictor of microscopic satellites in primary cutaneous melanoma. Histopathology 61(5):889–898 Wolf P, Rieger E, Kerl H (1993) Topical photodynamic therapy with endogenous porphyrins after application of 5-aminolevulinic acid: an alternative treatment modality for solar keratoses, superficial squamous cell carcinomas, and basal cell carcinomas? J Am Acad Dermatol 28(1):17–21 Wolf IH, Smolle J, Binder B, Cerroni L, Richtig E, Kerl H (2003) Topical imiquimod in the treatment of metastatic melanoma to skin. Arch Dermatol 139 (3):273–276 Wong JH, Cagle LA, Kopald KH, Swisher SG, Morton DL (1990) Natural history and selective management of in transit melanoma. J Surg Oncol 44(3):146–150 Wong SL, Morton DL, Thompson JF, Gershenwald JE, Leong SP, Reintgen DS, Gutman H, Sabel MS, Carlson GW, McMasters KM (2006) Melanoma patients with positive sentinel nodes who did not undergo completion lymphadenectomy: a multi-institutional study. Ann Surg Oncol 13(6):809–816 Wulfken LM, Scholl F, Pfeiff B, Dill D (2015) Kryochirurgie beim kutan-subkutan metastasierten

894 malignen Melanom. Aktuelle Derm 41(08/09):A55. https://doi.org/10.1055/s-0035-1558624 Wussow PV, Block B, Hartmann F, Deicher H (1988) Intralesional interferon-alpha therapy in advanced malignant melanoma. Cancer 61(6):1071–1074 Zapas JL, Coley HC, Beam SL, Brown SD, Jablonski KA, Elias EG (2003) The risk of regional lymph node

J. F. Thompson et al. metastases in patients with melanoma less than 1.0 mm thick: recommendations for sentinel lymph node biopsy. J Am Coll Surg 197(3):403–407 Zeitouni N, Dawson K, Cheney R (2005) Treatment of cutaneous metastatic melanoma with imiquimod 5% cream and the pulsed-dye laser. Br J Dermatol 152(2):376–377

Part VI Uncommon Presentations of Melanoma

Acral Lentiginous Melanoma Yukiko Teramoto, Hector Martinez-Said, Jun Guo, and Claus Garbe

Contents Epidemiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 898 Pathogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 899 Clinical Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900 Dermoscopic Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900 Pathological Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905 Immunohistochemistry of ALM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 906 Molecular Feature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . BRAF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NRAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . KIT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . TERT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NF1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mutation Burden . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

908 908 908 908 909 909 909

Y. Teramoto Department of Skin Oncology/Dermatology, Saitama Medical University International Medical Center, Saitama, Japan e-mail: [email protected] H. Martinez-Said Melanoma Clinic, Surgical Oncology, Instituto Nacional de Cancerología, México City, Mexico e-mail: [email protected] J. Guo Department of Renal Cancer and Melanoma, Key Laboratory of Carcinogenesis and Translational Research, Ministry of Education, Peking University Cancer Hospital and Institute, Beijing, China e-mail: [email protected] C. Garbe (*) Centre for Dermatooncology, Department of Dermatology, Eberhard Karls University, Tuebingen, Germany e-mail: [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_67

897

898

Y. Teramoto et al. Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909 Prognosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Compared with Other Subtypes of CM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Comparison in the Localization of ALM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Compared in Racial Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

910 910 910 911

Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surgery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Immunotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Targeted Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

911 911 916 917

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 918 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 918

Abstract

Acral lentiginous melanoma (ALM), described by Reed more than 30 years ago, has many unique characteristics. However, ALM is still a mysterious form of skin cancer. Of the four major subtypes of cutaneous melanoma (CM) worldwide, ALM is the rarest. Although ALM is the most common subtype of CM among Asians, Middle Easterners, and Africans, the absolute number of ALM cases is small. This is because CM is a rather rare skin cancer in Asia, Middle East, and Africa, compared to those in Western countries. ALM occurs on the palms, soles, and nails as unexposed lesions. The pathogeneses of ALM remain unknown. Furthermore, ALM has a more advanced status at diagnosis, and the prognosis of ALM is the worst among all types of CM. Several reasons for the poor prognosis are proposed from various perspectives; however, they have not been ascertained. Therefore, there is need for further studies to understand the nature of ALM, establish its original diagnosis, and identify treatment modalities for ALM in the future.

Epidemiology In 1976, Reed first described a unique variant of melanoma as lentiginous melanoma (otherwise known as lentigo maligna), that is, pigmented lesions on the palms and soles that are characterized by the lentiginous (radial) growth phase pattern, evolving over months or years to a dermal

(vertical) invasive stage (Reed 1976). Reed named this subgroup of melanoma on the foot, which had a characteristic lentiginous, radial component of melanocytic proliferation, as “plantar lentiginous melanoma (PLM).” PLM also qualified as the most common subtype of melanoma in Blacks and that those with this melanoma are a very poor prognosis group (Arrington et al. 1977). In 1986, Clark et al. classified cutaneous melanoma (CM) into four major subtypes based on its clinicopathological features: superficial spreading melanoma (SSM), nodular melanoma (NM), lentigo maligna melanoma (LMM), and acral lentiginous melanoma (ALM) (Clark et al. 1986). This classification has been most commonly used worldwide for the past three decades. Subungual melanoma is one anatomical variant of ALM, with the first clinical details described by Hutchinson in 1886 (Hutchinson 1886). Subungual melanoma develops within the nail matrix and is relatively rare, accounting for only 2–3% of all cutaneous melanomas in whiteskinned populations (Blessing et al. 1991) and approximately 10–20% in Asians (Ishihara et al. 2008; Kato et al. 1996; Takematsu et al. 1985). The incidence rates of CM show ethnic differences in some respects. Firstly, among darkerskinned populations (i.e., Asians, Blacks, and Hispanics), lower overall CM incidence has been reported compared to the white-skinned populations (i.e., Caucasians) (Bradford et al. 2009; Cronin et al. 2018). The reported estimated CM incidence in Europe and the United States ranges from 13.5 to 19.7 per 100,000 population per year (Guy et al. 2015; Svedman et al. 2016).

Acral Lentiginous Melanoma

However, in Japan, 1500–2000 cases per year were reported (Yamazaki et al. 2015). Furthermore, evidence has shown that in darker-skinned patients, ALM proportion of CM is much higher than in the white-skinned ones. In white-skinned populations, ALM accounts for approximately 1–7% of all CM (Bradford et al. 2009; Phan et al. 2006; Shaw and Koea 1988; Wang et al. 2016; Cress and Holly 1997). In contrast, the proportion of ALM in CM was reported to be 47–86.6% in East Asians (Jung et al. 2013; Chang et al. 2004; Luk et al. 2004), 36% in American black-skinned populations (Bradford et al. 2009), and 44.1% in Mexico (Lino-Silva et al. 2016). Overall, the incidence rate of ALM shows only few ethnic differences and is in the range of 0.4–0.8 cases per 100,000 inhabitants per year (Teramoto et al. 2018). Consequently, the absolute number of ALM is extremely low globally, with ALM being the rarest of the four CM subtypes (see also chapter ▶ “Clinical Epidemiology of Melanoma”). While the incidence of all CM has been increasing dramatically in most western countries, the incidence of ALM did not increase at the same rate as other CM subtypes. The proportion of ALM patients among all CM gradually decreased from 4.3% in 1983 to 3.3% in 2015, in the German Central Malignant Melanoma Registry (Teramoto et al. 2018). This might have resulted from the fact that the impact of ultraviolet exposure is not required for the development of ALM, unlike in other CM subtypes. The mean age of ALM patients at diagnosis was reported as 50–60 years in previous reports (Bradford et al. 2009; Phan et al. 2006; Jung et al. 2013; Paolino et al. 2016; Lv et al. 2016; Bello et al. 2013; Teramoto et al. 2018). ALM occurs more in older patients, compared with SSM or NM (Teramoto et al. 2018). The incidence rates of ALM between men and women are similar; however, men tend to have more advanced status of tumor at diagnosis than women (Bradford et al. 2009). Regarding clinicopathological characteristics of ALM, at diagnosis, about 36–47.9% of patients with ALM present with ulceration (Jung et al. 2013; Phan et al. 2006; Lv et al. 2016; Bello

899

et al. 2013; Kuchelmeister et al. 2000), although this proportion is lower (17.6–25.5%) in CM (Bello et al. 2013; Kuchelmeister et al. 2000). Ulcerations are detected more commonly at diagnosis in AML cases than in the other subtypes of CM. The high rate of ulceration has been explained to be a consequence of mechanical factors, similar to the presentation in acral lesions which result from frequent exposure to trauma (Durbec et al. 2012). Previous findings of researches have also shown that ALM tends to have higher tumor thickness compared to non-ALM tumors, with the median tumor thickness reported at 0.8–2.2 mm (Phan et al. 2006; Paolino et al. 2016; Bello et al. 2013). However, these numeric data have been shown to have a high variability. At diagnosis, a fifth (20%) of patients with ALM have some form of metastasis (Teramoto et al. 2018). Gumaste et al. (2014) reported that the local recurrence rate in ALM was nearly twice that of non-ALM CM at 39%. This high local recurrence could potentially worsen the survival in ALM patients. ALM occurs more frequently on the foot than on the hand, as reported in previous studies where lower incidence of ALM occurring on the hand (17.9–22%) than on the feet (78–82.1%) has been shown (Bradford et al. 2009; Teramoto et al. 2018). The most predominant site of ALM on the foot is the sole, especially on the weightbearing areas. The second most frequent site is the first toe. Meanwhile, the predominant sites of ALM on the hands are the fingernails, especially those on the thumb. The thumb and first toe are subjected to long-term physical stress and trauma and are, therefore, the most common sites of subungual melanoma (Jung et al. 2013).

Pathogenesis Many candidate factors for pathogenesis of ALM have been proposed, because of its various unique characteristics compared with the other subtypes. Firstly, almost all ALM arise on the palmoplantar or subungual areas. These sites are chiefly unexposed to the sun; therefore, it is assumed that ultraviolet radiation (UVR) does not play a

900

significant role in the occurrence of ALM. Recently, high degrees of ultraviolet signature mutations were identified in non-ALM, compared with ALM in whole genome sequence studies. These results support the hypothesis that ALM is unrelated to UVR. Secondly, the predilection sites of ALM are weight-bearing and received chronic mechanical stimulations. Thus, trauma (Zhang et al. 2014; Phan et al. 2006; Green et al. 1999) and pressure (Minagawa et al. 2016; Jung et al. 2013) are considered as two main predisposing factors of ALM. Minagawa et al. plotted the sites of ALMs on the sole and found that ALM occurred frequently on areas with plantar and shear stress (Minagawa et al. 2016). In a retrospective study, Zhang et al. indicated that the extremities recorded remarkably higher risks (adjusted odds ratio 3.968; 95% confidence interval 2.267–5.592) of post-trauma melanoma than other sites. The conclusion was that trauma was associated with ALM (Zhang et al. 2014). In addition, several distinct gene mutations (Liau et al. 2014; Fernandes et al. 2015; Furney et al. 2014; Zebary et al. 2013) have been hypothesized to promote ALM development, as described below.

Clinical Features ALM affects the soles, palms, toes, fingers, and nails. ALM, on the volar surfaces of the hands and feet, typically, first appears as a slowly enlarging, darkly pigmented macule and patch. It shows as variegated brown or black pigmented patch with irregular borders, mimicking a stain, bruise, hematoma, or hemorrhage. These appearances correspond to a lentiginous (radial) growth phase, clinically, and an in situ phase, histopathologically (Fig. 1, 1). In this phase, melanoma cells usually remain within the epidermis. This early phase can persist for months or years. Subsequently, this flat pigmented patch becomes partially elevated and ulcerated, with bleeding, when the melanoma cells invade the dermis (Fig. 1, 2). This is the onset of the vertical growth phase. Sometimes ALM can present as amelanotic or hypomelanotic lesions (Fig. 1, 3). It appears as

Y. Teramoto et al.

pinkish-red or light brownish lesions, which mimic several diseases, such as warts, calluses, non-healing ulcers, or other skin tumors (e.g., eccrine poroma, pyogenic granuloma, and so on). When subungual melanoma begins in the nail matrix, it usually presents as a black or brown longitudinal streak in the nail, most commonly found in the first toe or the thumb. The streak slowly becomes wider, with variable colors (Fig. 1, 4). A dark pigmentation appears over the hyponychium and continues onto the proximal or lateral nail folds (Hutchinson’s sign) (Fig. 1, 5). With progression of the tumor, the nail plate destruction begins, and a mass forms below the nail plate (Fig. 1, 6). When subungual melanoma shows up with a nonpigmented appearance, the correct diagnosis is very difficult. The differential diagnoses of subungual melanoma are subungual hematoma, onychomycosis, chronic paronychia, and benign melanonychia. In particular, melanonychia in children often shows worrisome clinical features suggestive of melanoma, such as pseudo-Hutchison sign, width of the pigmented band, evolution, various colors, and nail dystrophy. However, majority of cases do not show aggressive behavior (Fig. 2). Spontaneous regression is often observed during the course of several years. Therefore, watchful waiting is an accurate choice in children’s melanonychia, and biopsy is required in selected cases (Cooper et al. 2015; Ohn et al. 2016).

Dermoscopic Findings The use of dermoscopy by a well-trained dermatologist is very helpful for ALM diagnosis. The distinct dermoscopic features in acral sites are caused by the presence of skin marks. The four major dermoscopic patterns that are observed in melanocytic lesions on volar skin are the parallel furrow pattern, lattice-like pattern, fibrillar pattern, and parallel ridge pattern (Fig. 3). These dermoscopic patterns provide diagnostic information. The most specific dermoscopic feature of ALM is the parallel ridge pattern, which is a band-like pigmentation on ridges of volar skin marks. The sensitivity and

Acral Lentiginous Melanoma

Fig. 1 Clinical features of ALM. (1) A lentiginous growth phase. Irregular pigmentations with asymmetrical shape, irregular and poor defined borders, and various colors are shown on volar skin. (2) A vertical growth phase. The part preceding the flat pigment on the volar site became elevated and ulcerated, with frequent bleeding. (3) Amelanotic ALM. It appears as a pinkish-red lesion, which is mimicking several diseases. Early and correct diagnosis for these noncolored lesions is very difficult based only on the

901

appearance. (4) Subungual melanoma at early stage. The streak becomes wider slowly with variable colors for several years. (5) Hutchinson’s sign. A dark irregular pigmentation extends to the skin over the proximal or lateral nail folds. (6) Subungual melanoma at advanced stage. A mass forms below the nail plate; it then grows, with destruction of the nail. Further progressive tumor invades into the distal phalanx. Rather advanced cases often present with loss of nail and distal phalanx

902

Y. Teramoto et al.

Fig. 2 Clinical features of melanonychia in children. (1) The great toe of a 10-year-old Japanese girl. (2) The thumb of a 7-yearold Japanese girl. Most melanonychia in children shows worrisome clinical features, suggestive of melanoma, such as pseudoHutchison sign, width of the pigmented band, evolution, various colors, and nail dystrophy

Fig. 3 The four major dermoscopic patterns of melanocytic lesions on volar skin. (1) The parallel furrow pattern: in which parallel pigmented lines are seen along the furrows of the skin marks. (2) The fibrillar pattern: fibrillar pigmentation run in a slanting direction across the parallel skin marks. The endpoints of the fibrils composing

this regular fibrillar pattern are lined up on the straight lines corresponding to the surface furrows. (3) The lattice-like pattern: composed of the parallel lines on the furrows as well as the lines bridging the parallel lines, resulting in a lattice-like pigmentation. (4) The parallel ridge pattern: band-like pigmentation on the ridges of the skin marks

Acral Lentiginous Melanoma

903

specificity of the parallel ridge pattern for diagnosing early ALM are 86% and 99% (Saida et al. 2011). In contrast, the prevalent pigmentation patterns of acral melanocytic nevi are the parallel furrow pattern (with the pigmentation appearing along the skin furrows) and the lattice-like pattern (with longitudinal and thicker transversal lines surrounding the eccrine pores). Although the fibrillar pattern can be detected in both ALM and melanocytic nevus, an irregular and partial fibrillar pattern is a sign of ALM. The eccrine pores can be observed as small open circles, which help in distinguishing the ridges of the surface skin marks from the furrow (see also chapter ▶ “Dermoscopy/Confocal Microscopy for Melanoma Diagnosis”). A three-step algorithm, proposed by Saida et al. (Saida et al. 2011; Koga and Saida 2011) in 2011, provided management approaches for acquired acral melanocytic lesions (Fig. 4). The first step was to perform a biopsy of any lesion that shows the parallel ridge pattern immediately, regardless of the size. In the absence of the parallel ridge pattern, the three typical benign dermoscopic patterns,

namely, typical parallel furrow, typical lattice-like, and regular fibrillar pattern, were checked, as a second step. If these three typical benign patterns are shown, further follow-up is not needed, regardless of the size. The third step becomes necessary when the lesion does not show these typical benign dermoscopic patterns outlined in the second step. The third step requires that the maximum diameter of the lesion be measured. Any lesion with >7 mm maximum diameter requires biopsy, whereas 7 mm lesions would only need periodic clinical and dermoscopic review. The differential diagnoses of ALM, using dermoscopy, include some diseases and conditions such as Laugier-Hunziker syndrome, Peutz-Jeghers syndrome, and skin reactions due to cytotoxic chemotherapies (e.g., 5-FU, TS-1, docetaxel, and cytarabine); these also present with volar pigmentations with the parallel ridge pattern (Fig. 5). However, these conditions present with characteristics pigmented lesions that distinguish them from ALM in situ, easily. The distinguishing features are based on the unique labial and digital distribution, smaller size, and

Acquired melanocytic lesions on volar skin

1 st step

2nd step

Non-parallel ridge pattern

Typical parallel furrow, lattice-like pattern or regular fibrillar pattern

3rd step

Dermoscopic features except definitive typical parallel furrow, lattice-like pattern, or regular fibrillar pattern

Diameter ≤7 mm

No follow-up

Parallel ridge pattern

periodic follow-up

Diameter >7 mm

Biopsy for histopathological evaluation

Fig. 4 The three-step algorithm for the management of acquired volar melanocytic lesions. (Adapted from Saida et al. 2011)

904

Y. Teramoto et al.

Fig. 5 The clinical features of diseases (differential diagnosis) mimicking ALM. (1) All fingernails change to diffuse brown color by chemotherapy with docetaxel for lung

cancer. (2) Green nails mimicking subungual melanoma. (3) Dermoscopic feature of green nail. (4) Hematoma mimicking volar ALM. (5) Dermoscopic feature of hematoma

number of pigmented macules. With the development of the tumor, the observations in the dermal invasive ALM include blue-whitish veil, regression structure, and polymorphous and atypical vascular patterns, as observed in other subtypes of CM (Fig. 6, 3 and 4). In subungual melanoma, the prominent dermoscopic feature at its early stage is the

irregular longitudinal lines on a brown background. According to the progression of the condition, dermoscopy reveals a widening irregular pigmented band and a pigmentation of the cuticle (micro Hutchinson’s sign) (Fig. 6, 1 and 2). Hutchinson’s sign is often presented as the parallel ridge pattern on the digital tip, or the irregular pigment network on the proximal and lateral nail

Acral Lentiginous Melanoma

905

Fig. 6 Dermoscopic features of ALM. (1) A subungual melanoma at an early stage. Irregular longitudinal lines on a brown background. (2) Micro-Hutchinson’s sign

(upward arrowhead). (3) Nodule on developed ALM. Dermal invasive lesion shows blue-whitish veil. (4) Ulcerated nodule shows polymorphous and atypical vascular pattern

folds, on dermoscopy. As a result of the rapid enlargement of the proximal edge of the longitudinal lines, the longitudinal band shows a wider proximal end than the distal end, which presents with a triangular shape.

5. The crista intermedia and the crista limitans are situated under the ridges and furrows on the skin surface. 6. Thicker cornified layer compared with hairy skin lesion.

Pathological Features There are several unique pathological features in the normal skin of a volar lesion: 1. No hair follicles and no sebaceous glands. 2. Poor melanin. 3. The dermatoglyphic pattern, consisting of furrows and ridges, is regular. 4. Dermal eccrine ducts are connected to the crista intermedia and open to the center of the ridges.

Also important is the anatomical structure of acral skin, especially when considering the dermoscopy of acral nevus and melanoma (Fig. 7). Histopathological diagnosis of early-stage ALM is often challenging because its pathological features show under diagnostic atypia. In the pathological diagnosis of early-stage ALM, it is important to cut perpendicular to the skin marks, so that the two types of epidermal rete ridges, namely, crista limitans and crista intermedia, can be observed under the parallel furrows and

906

Y. Teramoto et al.

1

2

furrow ridge

crista limitans

crista intermedia

Fig. 7 Correlations between dermoscopic feature. (1) and pathological feature (2) of the normal skin. When volar skin is cut perpendicular to the skin marks (1, pink borderline), we can observe two types of epidermal rete ridges, crista limitans and crista intermedia. They correspond with

furrows (blue downward arrowhead) and ridges (yellow left right arrow), respectively. Dermal eccrine ducts (2, green rightward arrowhead) are connecting to the crista intermedia and open to the center of the ridges (1/2, green upward/downward arrowhead)

parallel ridges, respectively, at dermoscopy. In early-stage ALM, slightly atypical melanocytes show solitary proliferations at the basal layer of the crista intermedia, which is the epidermal rete ridge situated under the surface ridges (Fig. 8, 1). These pathological features correspond to the parallel ridge pattern at dermoscopy (Fig. 7). This finding is extremely helpful in the diagnosis of early-stage ALM. In early subungual melanoma, atypical melanocytes proliferate slightly in the epidermis of the nail matrix. The pathological diagnosis of early subungual melanoma is also difficult for the same reasons as the acral sites. In the radical growth phase, lentiginous atypical melanocytes spread along the basal cell layer. There are acanthosis and epidermal hyperplasia with elongation of the rete ridges. Thereafter some tumor cells can be also found in the upper layer of the epidermis (ascent) and show a pagetoid spread (Fig. 8, 2). In the vertical growth phase, severe cytologic atypical cells, which are often spindleshaped or rounded pagetoid tumor cells, extend into the deeper levels with lichenoid lymphocytic infiltration and fibrosis, as with other subtypes of CM. In the dermis, dermal components of ALM

do not show maturation (Fig. 8, 3 and 4) (see also chapter ▶ “Classification and Histopathology of Melanoma”).

Immunohistochemistry of ALM For ALM lesions, hematoxylin and eosin staining can be used to visualize the histopathology, for diagnosis; this is important in differentiating ALM from nevus. When the distinction is unclear, immunohistochemistry can be useful. Furthermore, it is particularly valuable for measuring Breslow thickness and detecting micrometastases in sentinel lymph nodes, because immunohistochemistry can distinguish between melanocytederived tumors and nonmelanocyte-derived tumors. Numerous immunohistochemical markers used for ALM exists as follows (Desai et al. 2018): 1. S-100 protein, which was first discovered in glial cells and is a 21 kDa acidic calciumbinding protein. It is very sensitive, with a positivity rate of >95% in ALM (Kim et al. 2003); however, it is not a specific marker.

Acral Lentiginous Melanoma

907

Fig. 8 Pathological features of ALM. (1) A lentiginous growth phase. ALM in situ prefers to grow around the openings of eccrine sweat ducts along the ridges of the dermatoglyphics (upward arrowhead). (2) The start of vertical growth phase. The tumor cells proliferate to the upper layer of the epidermis (ascent) and show a pagetoid

spread. (3) The vertical growth phase (low magnification). Tumor nests invade into the deeper levels with lymphocytic infiltration. (4) The vertical growth phase (high magnification). Tumor cells show severe cytologic atypia, which are often spindle-shaped or rounded pagetoid tumor cells

2. HMB-45, an earlier melanoma-specific marker, stains the cytoplasmic premelanosomal glycoprotein gp100. It is often shown as positive, in cases of CM (sensitivity of 69–93%), but can be negative or focally positive in 36% of amelanotic ALM (Choi et al. 2013) 3. Melanoma antigen recognized by T cells-1 (MART-1) also known as melan-A is a cytoplasmic protein of melanosomal differentiation recognized by T cells. The antibody of this antigen is similar to HMB-45 (sensitivity of 75–92% and specificity of 95%) (Ohsie et al. 2008). However, it is more diffuse and intense in staining than HMB-45 (Ohsie et al. 2008) 4. Ki67 is used to assess the mitotic activity of the dermal component. It is useful for

distinguishing benign from malignant melanocytic tumors. Careful interpretation of Ki67 staining is needed, because Ki67 stains also lymphocytes from an associated lymphocytic infiltrate (Piliang 2011) 5. Tyrosinase is an enzyme that functions in the first step of melanin synthesis to hydroxylate tyrosine. The reported sensitivity for CM is 84–94% (Ohsie et al. 2008) 6. Microphthalmia transcription factor (MITF) functions during embryogenesis as a transcription factor protein for the development of melanocytes. Its sensitivity for CM is 81–100% (Ohsie et al. 2008); however, it is relatively nonspecific. Lymphocytes, histiocytes, fibroblasts, Schwann cells, and smooth muscle cells show MITF staining (Busam et al. 2001)

908

Molecular Feature In 2005, according to Bastian et al., the four new groups into which melanoma was classified based on genetic alterations at different sites were (1) chronic sun-damaged melanoma (CSD); (2) non-CSD melanoma; (3) acral melanoma (AM), which roughly correspond to LMM, SSM, and AM, respectively; and (4) mucosal melanoma. Each of these groups has differing levels of ultraviolet light exposure. Furthermore, it was reported that 81% of non-CSD melanoma had mutations in BRAF or NRAS, while the other groups had few mutations (33% of AM) in these two genes. Furthermore, melanoma with wildtype BRAF or NRAS frequently had increased number of copies of cyclin-dependent kinase 4 (CDK4) and cyclin D1 (CCND1) genes. CDK4 and CCND1, downstream components of RAS-BRAF pathway, lead to the dysregulation of the cell cycle at the G1–S checkpoint (Curtin et al. 2006). In a recent studies, CM from intermittently sun-exposed skin showed mutations in BRAF V600E (50%), NRAS (15%–20%), and KIT (1–2%) (Davis et al. 2018). Conversely, it was reported that AM had BRAF mutations (3.6–17%), activating mutations or amplifications of wild-type KIT (0–36%), and NRAS mutations (0–15%) (Shim et al. 2017; Yun et al. 2011; Si et al. 2012; Jin et al. 2013; Kong et al. 2011; Zhou et al. 2012; Ashida et al. 2009; Ashida et al. 2012; Curtin et al. 2006; Takata et al. 2007).

BRAF Most BRAF mutations detected in AM were T1799A (V600E) in exon 15, which is a prevalent mutation in CM. There are only a few reports about the prevalence of BRAF mutation of AM in Caucasians: 17% in Sweden (Zebary et al. 2013) and 9.5% and 15% in the United Kingdom (Saldanha et al. 2006; Lang and MacKie 2005). The prevalence rate of BRAF mutation of AM among Eastern Asians was reported to be 3.6–15.5%. It is assumed that there is no racial difference in the frequency of BRAF mutation in AM.

Y. Teramoto et al.

In CM, the common characteristics of primary tumors with BRAF mutation in all previous studies are only two: younger age and non-CSD site. Other clinicopathological features, such as presence of mitosis, ulceration, and tumor thickness, vary depending on each study. Conversely, BRAF mutations in primary CM have no apparent impact on disease-free survival (DFS) and overall survival (OS) as documented in the majority of previous reports before the BRAF inhibitor era (Long et al. 2011). However, there are no reports regarding clinicopathological features of AM. One study reported the presence of BRAF mutations with a favorable prognosis in ALM (Hong et al. 2014). Of course, this study has many limitations, and further study is needed to confirm the role of BRAF mutation in AM clinically (see also chapter ▶ “Molecular Epidemiology of Melanoma”).

NRAS The frequency of NRAS mutations in AM was reported to be between 0 and 15% in previous studies (Shim et al. 2017; Zhou et al. 2012; Si et al. 2012; Ashida et al. 2012; Takata et al. 2007; Zebary et al. 2013). In AM, codon 61 mutations are heterogeneous, with C181A (Q61K) and A182G (Q61R) as the most frequent mutations, similar to other types of CM. Patients with tumors having NRAS mutation were reported to show more advanced American Joint Committee on Cancer stages, as well as lower median OS, compared with patients with wild-type tumors (Jakob et al. 2012; Lee et al. 2011; Thumar et al. 2014).

KIT KIT mutations have been identified in invasive CM, including AM and amelanotic AM (Choi et al. 2013). KIT mutations and/or amplifications are more frequently identified in AM than in CM on sun-exposed areas. Previous studies reported that KIT mutations were detected in 0–13.8% of AM patients (Lin et al. 2013; Kong et al. 2011; Zhou et al. 2012; Ashida et al. 2009; Ashida et al. 2012; Yun et al. 2011; Shim et al. 2017; Jin et al.

Acral Lentiginous Melanoma

2013). KIT mutations in AM are most frequent in exon 11 (L576P), followed by exon 13 (K642E) (Hodi et al. 2013; Moon et al. 2018). In the near future, target therapy of tyrosine kinase inhibitors is anticipated for KIT mutated AMs. Furthermore, no correlation existed between intensity of c-KIT (based on immunohistochemistry study) and KIT mutation status, regardless of the silent mutations. Kong et al. reported that there was a correlation between overexpression of c-KIT in immunohistochemistry and the increase in the copy number of KIT, rather than KIT mutations (Kong et al. 2011). This suggests that KIT aberration status is not as reliable as increasing KIT copy number for screening KIT overexpression (Jin et al. 2013).

TERT Telomerase reverse transcriptase (TERT), a protein-coding gene, is important in CM. It is reported that cell death results from the pharmacological repression of TERT expression in melanoma cells. Up to 50% of CM cases have been reported to have TERT promoter mutations (Horn et al. 2013; Huang et al. 2013), but only 0–7% have been reported in AM (Macerola et al. 2015; Heidenreich et al. 2014; Liau et al. 2014). This difference may indicate a potential discrepancy in CM subgroups. Besides, in Caucasian populations, correlations have been reported between TERT promoter mutations and BRAF and NRAS mutations (Heidenreich et al. 2014; Macerola et al. 2015). Liang et al. further showed that there are cytotoxic effects of in vitro TERT inhibition on primary AM cells (Liang et al. 2017). These findings might reveal the role of TERT in AM tumorigenesis; and TERT inhibition is also anticipated as a therapeutic strategy in AM (Bai et al. 2017).

909

highest mutation burden and the strongest ultraviolet signature. They are also detected in 14–17% of all melanomas (Cirenajwis et al. 2017; Hayward et al. 2017). According to Rawson et al., AM with high ultraviolet signatures occurs, most frequently, on subungual sites and has BRAF or NF1 mutations (Rawson et al. 2017).

Mutation Burden In recent years, whole genome studies revealed that AM has a markedly different genomic landscape from CM. Firstly, AM has a low mutation burden. Furney et al. reported that the frequency of structural variations in AM is approximately half of that in non-ALM by whole genome sequencing (Furney et al. 2014). They also reported that the SNV (single nucleotide variants) mutation rate of acral (5000 mutations/genome) is markedly lower than that in non-ALM (30,000 mutations/ genome). Furthermore, several whole genome sequencing studies revealed that a high degree of ultraviolet signature mutations (C > T changes associated with pyrimidine dimer formation following exposure to ultraviolet light) were found in non-AM, compared with AM (Krauthammer et al. 2012; Hodis et al. 2012; Hayward et al. 2017). These findings suggest that UVR is an unlikely pathogenic factor of ALM. In fact, acral lesions occur on sun-exposed sites. Interestingly, these recent genomic studies indicate that AM acts more similar to mucosal melanoma, than to other subtypes of CM in the genomic landscapes. However, further analysis of AM is required in order to determine molecular factors involved in aggressive disease.

Diagnosis NF1 NF1, a GTPase-activating protein, is known to downregulate RAS activity through its intrinsic GTPase activity. The NF1 gene loss-of-function mutation can activate the MAPK signaling pathway. NF1 mutations are associated with the

Early diagnosis of ALM is very important for better prognosis; therefore, the correct diagnosis is essential. However, its identification is often difficult for dermatologists who have less experience with ALM, since it is an extremely rare disease. This easily leads to misdiagnosis. On the other hand, ALM with typical clinical

910

features, i.e., irregular pigmentation, can be diagnosed easily by the well-known “ABCDE rule” (Asymmetrical shape, irregular and poor defined Border, varied Color, Diameter > 6 mm, and Evolution). Its characteristic features and occurrence site give prompt diagnosis, and diagnostic accuracy can be enhanced with the use of dermoscopy. Typical dermoscopic findings can exclude ALM from other pigmented lesions (e.g., hematoma, hemorrhage, and melanocytic nevus) (Fig. 3, and 5). However, ALM sometimes shows amelanotic or hypopigmented clinical features, and this may cause confusion for many dermatologists. While amelanotic melanoma is said to represent 2–8% of all melanomas, amelanotic subungual melanomas are observed more frequently, representing up to 25% of all subungual melanomas (Pizzichetta et al. 2004). The lack of macroscopic pigmentation causes difficulty in accurate diagnosis of ALM. Metzger et al. reported a prevalence of misdiagnosis in 52% of subungual melanoma and 20% of palmoplantar melanoma, which was notably associated with greater tumor thickness (Metzger et al. 1998). Advanced ulcerated ALM tumor tends to show amelanotic or hypopigmented clinical features. Furthermore, Phan et al. reported that amelanotic lesions led to incorrect medical treatment in 20% of cases and inappropriate surgical procedure in 2% (Phan et al. 2006). The amelanotic lesions mimic various conditions clinically, such as benign tumors (e.g., poroma, plantar wart, and pyogenic granuloma), skin cancers (e.g., squamous cell carcinoma and porocarcinoma), inflammatory lesions (e.g., diabetic foot ulcer and chronic paronychia), and fungal infections (e.g., onychomycosis and tinea).

Prognosis Compared with Other Subtypes of CM ALM presents a more severe prognosis than other CM subtypes, as reported in previous studies. In our report, which analyzed 58,949 CM patients, the 5-year survival rates of each subtype of CM, excluding in situ cases, were 94.7% (SSM),

Y. Teramoto et al.

80.3% (NM), 95.45 (LMM), and 79.8% (ALM) (Teramoto et al. 2018). Bradford et al. also reported that the 10-year survival rates of ALM patients at stages II and III were 10–15% lower than that of all CM patients, with adjustment for stage (Bradford et al. 2009). Conversely, there were few reports based on small sample size analyses, and using the Kaplan-Meier survival curves, that indicated no significant difference in melanoma-specific survival and/or DFS between patients with ALM and non-ALM (Paolino et al. 2016, Wada et al. 2017). To date, previous studies have reported many clinicopathological parameters as significant prognostic factors in ALM, such as sex (Phan et al. 2006), tumor thickness (Bradford et al. 2009; Lv et al. 2016; Bello et al. 2013; Phan et al. 2006; Paolino et al. 2016; Teramoto et al. 2018), ulceration (Paolino et al. 2016; Bello et al. 2013; Teramoto et al. 2018), mitotic rate (Phan et al. 2006, Lv et al. 2016), invasion level (Clark level) (Phan et al. 2006; Kuchelmeister et al. 2000), microsatellites (Phan et al. 2006), vascular invasion (Lv et al. 2016), pathological stage (Bradford et al. 2009; Bello et al. 2013; Lv et al. 2016), sentinel lymph node positivity (Bello et al. 2013), regional lymph node metastasis (Lv et al. 2016), and pigmentation of the primary lesion (Phan et al. 2006). Moreover, there are efforts being made to investigate other factors (including molecular characteristics, immune status, DNA methylation, and gene expression signatures) associated with more biologically aggressive primary melanoma tumors. However, there are no specific parameters of poor prognosis for ALM only compared with other subtypes of CM. Most studies concluded that poor ALM prognosis was due to the delay in diagnosis. ALM tends to show a thicker primary tumor, a higher rate of ulceration of primary tumor, and a higher rate of metastasized cases at diagnosis than other subtypes of CM, especially SSM and LMM (Teramoto et al. 2018).

Comparison in the Localization of ALM The difference in localizations (i.e., sole, hand, or nail) of the primary tumor showed no prognostic

Acral Lentiginous Melanoma

significance. Although several studies analyzed characteristics and prognostic differences between ALM of the hands and ALM of the feet (Paolino et al. 2016; Boriani et al. 2014) or between subungual and nonsubungual ALM (Phan et al. 2006; Jung et al. 2013; Kato et al. 1996), no significant differences between the two groups have been shown in all these studies.

Compared in Racial Groups Bradford et al. reported that among the different racial groups (non-Hispanic Whites, Hispanic Whites, Blacks, and Asian Pacific Islanders), there were no statistical significant differences between the 5- and 10-year survival rates (Bradford et al. 2009).

Treatment Surgery Excision margins for ALM in volar skin: To date, several large randomized trials have been conducted to determine optimal surgical side margins for cutaneous melanoma (Table 1). The World Health Organization Melanoma Program compared side 1 cm and 3 cm margins in 612 melanoma patients, with thicknesses ranging from 0.8 to 2 mm (Veronesi et al. 1988). No significant differences in terms of local recurrence, DFS, and OS rates were observed between the two groups. The following two trials in Europe compared side margins of 2 or 5 cm. The Scandinavian Melanoma Group Study and the French Cooperative Group Trial were conducted in melanoma patients with 0.8–2.0 mm and 2 mm,

911

respectively. They also showed that there was no significant difference in local recurrence and survival rates between the two groups in both trials (Balch et al. 2001; Gillgren et al. 2011). The UK Melanoma Study Group performed a randomized trial in 900 melanoma patients with tumors on the trunk or limbs that were > 2 mm thick to compare side margins of 1 cm with 3 cm (Thomas et al. 2004; Hayes et al. 2016). Although no significant changes in local recurrence were recognized in both groups, locoregional recurrences were greater in the group with 1 cm side margin than in the group with 3 cm margins (hazard ratio, 1.26; P = 0.05). At a median follow-up of 8.8 years, the 3 cm side margin group had improved melanoma-specific survival, although this wider margin did not significantly improve OS. They concluded that a 1 cm side margin is insufficient for >2-mm-thick melanoma on the trunk and limbs. As for melanoma in situ, a side margin of 5–10 mm is widely accepted in some national guidelines. A recent prospective study also showed that 86% of 1120 in situ lesions were successfully excised with side margins of 6 mm and 98.9% were successful with side margins of 9 mm. The conclusion was that the 9 mm side margin was adequate for in situ lesions (Kunishige et al. 2012). The adequate depth of excision remains controversial. The excision depth is recommended to be at least to the muscle fascia level; however, deeper excisions have not improved survival outcomes (Kenady et al. 1982; Holmstrom 1992). A recent study showed no advantage of the excision of the deep muscle fascia but an increased risk of intralymphatic recurrences (Grotz et al. 2013). Based on these evidence, the current National Comprehensive Cancer Network (NCCN) Guidelines recommended adequate side margins, depending upon the tumor thickness of the primary lesion (NCCN Clinical Practice Guidelines in Oncology (NCCN Guidelines) Melanoma Version2.2018-January 19). However, the most available evidence was derived from clinical trials that mainly included SSM or NM patients. ALM patients were excluded from most clinical trials or were included in very few numbers because of the rarity of ALM in western countries. Therefore, it is still unknown whether the current recommended

912

Y. Teramoto et al.

Table 1 Randomized control trial for side margins of primary melanoma Published year 1998

Patient no. 612

Tumor thickness 0.8–2.0 mm

989

2 mm

900

>2 mm

2001

Authors Veronesi et al. CohnCedermark et al. Balch et al.

2003 2011

2000

2004 2016

Side margin 1 cm vs. 3 cm 2 cm vs. 5 cm 2 cm vs. 4 cm 2 cm vs. 5 cm 2 cm vs. 4 cm 1 cm vs. 3 cm

Local recurrence N.S.

RFS OS N.S. N.S.

Accrual of ALM patients fingers and toes excluded hands and feet excluded

N.S.

N.S. N.S.

N.S.

N.S. N.S.

N.S.

N.S. N.S.

None (trunk or proximal extremity) None

N.S.

N.S. N.S.

2 patients (0.2%)

N.S.a

N.S. N.S.b None

RPS relapse-free survival, OS overall survival, N.S. not significant, N.D. not described Locoregional recurrence was greater in 1-cm side margin group b Melanoma-specific survival was shorter in 1-cm side margin group a

excision margins should be applied to the treatment of ALM (that is clinically and biologically different from other subtypes). Moreover, if the current recommended side margins in the NCCN Guidelines (NCCN Clinical Practice Guidelines in Oncology (NCCN Guidelines) Melanoma Version2.2018-January 19) are applied to ALM lesions, the surgeons would confront another issue. The peripheral border of the in situ component of ALM is often discontinuous and poorly defined, and the excision margins can be difficult to ascertain. Furthermore, surgeons are also confused about the determination of side margins in other ways. ALMs often consist of in situ and invasive components, in the solitary tumor. It is still unclear whether side margins of 0.5–1 cm, from the peripheral border of the lesion, should be selected, as the peripheral lesion is considered an in situ lesion, or if side margins of 2 cm from the tumor periphery should be appropriate, since the lesion is considered a thick, invasive lesion. Thus, the adequacy of side margins for ALM should be addressed in future randomized trials (Fig. 9). Excision margin for subungual (nail apparatus) melanoma: The nail apparatus anatomy is characteristic and complex, comprising the nail plate, nail matrix, nail bed, proximal nail fold, and hyponychium. The proximal nail fold is continuous with the dorsal nail matrix, distally. The dorsal nail

matrix is continuous with the nail bed, beneath the nail plate, and the extensor digital skin, proximally. The nail plate is the translucent keratin protein produced from the nail matrix. The lunula, forming whitish crescent-shaped base of the visible nail, corresponds to the visible part of the nail matrix. The hyponychium is the distal part of the nail bed between the nail bed and tip of the digital skin. One characteristic feature of this anatomy is that the distance from the nail matrix or nail bed to the surface of the underlying bone (distal phalanx) is extremely narrow. Based on the characteristic anatomy of the nail apparatus, the surgeons have to consider optimal surgical treatment. With regard to in situ or  0.5mm-thick lesions, several authors have proposed digit-sparing, wide local excision with the periosteum of the distal phalanx and the peritendon (Fig. 10) (Clarkson et al. 2002; Lazar et al. 2005; High et al. 2004; Cohen et al. 2008; Imakado et al. 2008; Sureda et al. 2011). There have also been several reports proposing a two-stage surgery using artificial dermis tissue for the tentative coverage of the tissue defect produced by a wide local excision, followed by skin grafting after the histological evaluation for negative surgical margins (Hayashi et al. 2012; Smock et al. 2010). These are becoming common strategies for these stages of subungual melanoma (SUM).

Acral Lentiginous Melanoma

Fig. 9 Confusion of side margins with acral lentiginous melanoma that has in situ and invasive lesions in the same tumor. (1) The dotted yellow line indicates the invasive lesion. (2) The surgeon’s decision on the excisional line of

913

the tumor (black line, 2 cm side margin). (3) Red and blue dotted lines indicate side margins of 0.5 cm or 2 cm from the border of the in situ or invasive lesion, respectively

Fig. 10 Digit-sparing wide local excision for subungual melanoma. (1) Incision line. (2) Tumor resection includes the periosteum of the distal phalanx and the peritendon

In contrast, historically, amputation has commonly been performed in invasive SUM patients because of the proximity between the nail matrix and the distal phalanx (Daly et al. 1987). Haneke measured the distance from the deepest part of the nail matrix to the surface of insertion of extensor tendon of the middle finger in a young male, and

the measured distance was 0.8 mm (Haneke 2006). The shortest distances between the deepest part of the nail matrix and the distal phalanx surfaces were also investigated in cadavers (Kim et al. 2011). The mean measured distance of all digits was 0.90 mm. As for amputation level, several studies comparing various amputation levels showed that more

914

aggressive, proximal amputation had no clinical benefit (Haneke 2012). The recent trend has been to select more distal amputations without compromising recurrence or survival (Finley et al. 1994; Banfield et al. 1998). Despite the anatomic feature of the narrow distance between the nail matrix and the underlying bone surface, more conservative surgery for invasive SUM is being pursued. This is due to the fact that patients with invasive SUM, with no invasion or attachment to the distal phalanx, are often encountered in clinical practice. Nakamura et al. measured the distances between the deepest base of the tumor and the surface of the distal phalanx using 30 amputated specimens of invasive SUM (Nakamura et al. 2014). In all specimens with 0.5-mm-thick SUM, who underwent digit-sparing surgeries with side margins of 5–10 mm and had favorable prognosis (Cohen et al. 2008; Smock et al. 2010; Rayatt et al. 2007; Nakamura et al. 2015). Based on these encouraging reports, prospective confirmatory trial for evaluating the safety and efficacy of digit-sparing wide local excision is now ongoing (J-NAIL study, UMIN000029997) (Nakamura et al. 2018). Regarding other conservative excision procedures for invasive SUM, Moehrle et al. described the excision of SUM with the processus uncinatus, which is the distal part of the distal phalanx, as a “functional” surgery (Moehrle et al. 2003). Maekawa et al. also reported on digit-sparing surgeries with deeper excision procedures involving the nail apparatus, together with a layer of the cortical bone, parallel to the nail bed, preserving the length of the distal phalanx (Maekawa et al. 2014). Reconstruction: ALM is usually located on the palm and sole, which are cosmetically and functionally sensitive anatomical sites. Furthermore, it is often difficult to close the surgical defect, primarily because of the lack of skin

Y. Teramoto et al.

mobility on the palmoplantar lesion and because of substantial defect after wide local excision. Therefore, various reconstructive procedures including skin grafting, flaps, and secondary intention healing need to be used depending upon the defect size and site, the medical condition, and lifestyle of the patients. In most patients with ALMs on the palm, diagnosis is made at a small size since the patient easily notices the palmar lesion at an earlier stage. Despite this, the defect after a wide local excision is relatively large, and the full or splitthickness skin grafting is usually required. As for the donor site, the skin taken from the plantar arch areas will produce similar skin quality on the palm. However, the difference in skin quality between the palmar skin and the skin from other areas may be inconspicuous, particularly in elderly patients. The reconstruction of the sole is more complex than the palm because the sole largely involves weight-bearing areas. Additionally, ALM on this site is often diagnosed at a larger size because of its “out-of-site” location. A flap reconstruction with well-padded tissue is generally recommended for coverage of weight-bearing areas, if the excised margin of the tumor is deeper, and the most subcutaneous layer is excised close to the calcaneal bone. The medial plantar flap is the most appropriate reconstruction, which can provide the same skin quality, as well as some sensation, for weight-bearing areas. Although, skin grafting is needed to cover the planter arch, the donor site of this flap (Liu et al. 2014). An optional procedure for reconstruction of the heel includes the distally based sural artery neurocutaneous flap (Liu et al. 2014). However, the donor site is the posterior lower leg, which leads to a different quality of skin is most frequently the upper leg. Furthermore, a comparative study between the medial plantar flap and distally based sural artery flap showed that the incidence rate of postoperative complications was higher in the sural artery neurocutaneous flap group (Rashid et al. 2003). If the tumor thickness of the primary lesion is thinner, and the removal of the thick subcutaneous fat pad with the primary tumor is unnecessary, both split or full-thickness

Acral Lentiginous Melanoma

skin grafts may be acceptable, even for weightbearing reconstruction. Secondary intention healing is another option of reconstruction, but it takes a long time to epithelialize completely, and the patients need prolonged care, such as regular dressing changes and frequent observation. Actually, several investigators demonstrated that it took approximately 12–18 weeks to close the defects (mean defect size 32.6–36.5 cm2) (Jung et al. 2011, Oh et al. 2013). Despite these disadvantages, it still has the following advantages: (1) avoiding a secondary wound produced by tissue harvesting that needs skin grafting or skin flap; (2) a smaller scar, because of the natural wound contraction; and (3) granulation tissue formation that will be a cushion to absorb impact while walking (Jung et al. 2011). Recently, the effect of secondary intention healing, in combination with negative pressure wound therapy, has been studied (Oh et al. 2013). Although there was no difference in the time needed to close the defect between the two groups, the vascularity score and height of the scars were significantly better, and there were no cases of wound infections during treatment in the combination group. Reconstruction after wide local excision of nail apparatus has unique anatomical characteristics, because the surgical field, after excision, involves the exposed bone surface of the distal phalanx. However, general reconstructive procedure for this area’s coverage is skin grafting (Sureda et al. 2011) or covering with artificial dermis followed by skin grafting (Hayashi et al. 2012). Sentinel lymph node biopsy (SLNB): Lymphatic mapping and SLNB were first introduced in the early 1990s as a less invasive procedure to identify microscopic regional lymph node metastases (Morton et al. 1992). SLNB is derived from the concept that the lymphatic drainage from the primary lesion to the regional lymph node basins flows in an orderly, stepwise fashion, and these lymphatic drainage patterns would be considered in the same way as the spread of melanoma cells through the lymphatics. Since the detailed technique of the SLNB using blue dye injection was first reported and the reported SLN identification rate is 82% (Morton et al. 1992), various additional techniques for SLNB raised SLN detection

915

rate up to approximately 100%. Such techniques include lymphoscintigraphy using radioisotope and intraoperative use of handheld gamma probe (Kapteijn et al. 1997, Gershenwald et al. 1998), indocyanine green (ICG) injection and intraoperative use of a near-infrared camera (Namikawa and Yamazaki 2011; Stoffels et al. 2012; Polom et al. 2012), and SPECT/CT (Vermeeren et al. 2010). The main areas of lymphatic basin are limited in the extremities in ALM patients, which indicate that the targeted regional lymph nodes are in the inguinal and axillary areas in most cases. Rarely, there are possibilities that SLNs are identified in the popliteal, pelvic, epitrochlear, or other interval nodal areas (Duprat et al. 2005; Teramoto et al. 2016). Although SLNB is also widely accepted and applied to ALM patients, ALM has extensive differences in genetic, biological, and clinicopathological aspects, from other clinical subtypes. The actual usefulness of SLNB in this cohort is still unclear, because the number of ALM patients included in the several large prospective trials of SLNB, which were conducted in western countries, is unclear and seems small (Morton et al. 2014; Faries et al. 2017). Jeon et al. reported the prognosis and survival of 34 Korean ALM patients based on SLN status (Jeon et al. 2014). The rate of positive SLNs was 41.2%. Patients with negative SLNs had better prolonged DFS and OS (p < 0.05) and increased tumor thickness correlating with shorter DFS and OS (p < 0.05). This led to their conclusion that SLN status is a crucial prognostic factor for predicting OS and DFS. Ito et al. also investigated retrospectively 116 Japanese ALM patients who underwent SLNB (Ito et al. 2015). Patients with positive SLN showed significant shorter melanoma-specific survival and DFS (5-year survival rate, 37.5% vs 84.3% and 37.5% vs 77.9%; p < 0.0001 and = 0.0024, respectively). The impact of positive SLNs on melanomaspecific survival was increased in ALM patients with >1 mm thickness (5-year survival, 22.7% vs 80.8%; P = 0.0005). Although the sample size in these studies was small, the results of positive SLN status in association with recurrence and worsened survival in ALM patients show similar trend with the results in larger prospective randomized trial

916

(Morton et al. 2014) (see also chapter ▶ “Biopsy of the Sentinel Lymph Node”). Regional lymph node dissection (LND): The main target areas for LND in ALM patients are axillary and inguinal area. Generally, LND is performed for local control and staging (Balch 1990); however, the therapeutic value is still unclear. Although, until recently, immediate LND has been recommended usually, for patients with positive SLN metastases, clear evidence of the efficacy of LND in such patients was lacking. In 2018, the MSLT-II trial compared patients with SLN micrometastases who had immediate LND (dissection group) with those having nodal observation using ultrasonography (observation group). The results indicated that immediate LND was not associated with prolonged melanoma-specific survival and lymphedema was more frequent in the dissection group than the observation group (24.1% vs 6.3%) (Faries et al. 2017). However, as described above, the number of ALM patients included in this trial is unclear and seems small. Therefore, the true effect of immediate LND in ALM patients is still unclear. Postlewait et al. reported that lymphedema occurred in only 13 of 254 (5%) axillary LNDs, according to their retrospective study. Although lymphedema due to LND is often cited as a reason to change or forgo surgical intervention, this may be less of a concern for axillary LND. In contrast, LND is applied to ALM patients who have clinical nodal disease, as well as patients with other clinical subtypes, although there are no definitive studies investigating therapeutic value of LND for nodal disease in ALM patients. It is generally recommended that the extent of axillary LND is from level I to level III lymph nodes because of the variety of drainage patterns in the second-tier nodes as well as the possibility of nodal recurrence in case of narrow dissection areas (Garbe et al. 2008). In contrast, several investigators have questioned the necessity of a level III dissection in all cases including a positive SLN. They proposed that level III dissection may be added only when suspicious nodes are involved in this level (Karakousis 1998; Meyer et al. 2002; Serpell et al. 2003). LND in the

Y. Teramoto et al.

inguinal areas is still a controversial issue; whether surgeons should perform inguinal LND alone or pelvic LND in addition to inguinal LND is unclear. As predictive factors for pelvic nodal disease, the current NCCN Guidelines for melanoma (Version 2.2018) proposed the following four items, (1) a metastatic Cloquet’s node (CN), (2) four or more metastatic inguinal nodes, (3) metastatic inguinal nodes with extracapsular extension, and (4) palpable inguinal nodes, and that additional pelvic LND in patients with these factors should be considered (NCCN Clinical Practice Guidelines in Oncology (NCCN Guidelines) Melanoma Version2.2018-January 19). Conversely, a study investigating lymphatic drainage patterns using lymphoscintigraphy and/or SPECT/CT revealed that over 50% of patients showed second-tier nodal drainage from the inguinal SLNs to the pelvic nodes (van der Ploeg et al. 2009). The other study investigating drainage patterns from lower limb melanoma also showed that the lymphatic drainage directly from the inguinal to the pelvic nodes bypassing CN was observed in 37.5% of patients, indicating that negative CN status alone is of limited value as a predictive factor for withholding pelvic LND (Teramoto et al. 2016) (see also chapter ▶ “Inguinofemoral, Iliac/Obturator, and Popliteal Lymphadenectomy for Melanoma”).

Immunotherapy In the last decade, the treatment of advanced melanoma changed dramatically. Three immunecheckpoint inhibitors were approved by the US Food and Drug Administration (FDA) for metastatic or unresectable melanoma patients: antiPD-1 agents (nivolumab and pembrolizumab) and a monoclonal antibody against CTLA-4 (ipilimumab).These drugs are used as monotherapy for anti-PD-1 agents or combination therapy as nivolumab plus ipilimumab. The immunotherapy has improved the response rate and the OS rate compared to traditional cytotoxic chemotherapies (see also chapter ▶ “Checkpoint Inhibitors in the Treatment of Metastatic Melanoma”).

Acral Lentiginous Melanoma

However, the evidence is limited to support the efficacy of immunotherapy in ALM. This is because major clinical trials investigating the efficacy of these new drugs include very small number of ALM patients, because of its rarity. D’Angelo et al. retrospectively analyzed melanoma patients who received nivolumab therapy in six clinical trials (CheckMate-037, CheckMate066, CheckMate-067, CheckMate-069, CA209–038, and CA209–003). In these pooled data, only 11–15% of all melanoma patients were diagnosed with acral melanoma, uveal melanoma, or unknown primaries (D’Angelo et al. 2017). The rate of ALM in these trials may be less than 10%. Two retrospective studies reported objective response rates to anti-PD-1 antibody in ALM patients (Choi et al. 2013; Shoushtari et al. 2016). The treatment outcome to anti-PD-1 antibody was not different in ALM compared with CM. However, these studies have several limitations, for example, being retrospective studies and having small number of cases. Therefore, it is possible that ALM shows different efficacy to immunotherapy by common subtypes of melanoma. Meanwhile, based on some hypotheses tested, immunotherapy shows low efficacy for ALM. Firstly, Farney et al. reported the lower somatic mutation rates of ALM versus CM (Furney et al. 2014). Some studies have indicated that mutation burden has been clearly linked with some tumor types with response to immunotherapy (Snyder et al. 2014; Rizvi et al. 2015). The reason is that putative “passenger” mutations actually function to generate so-called neoantigens, and peptides may be recognized as foreign to the adaptive immune system (Meng et al. 2015). Secondly, several studies reported that PD-1 ligand (PD-L1) tumor staining was associated with responses to nivolumab in melanoma as with non-small cell lung cancer (Robert et al. 2015; Taube et al. 2014; Weber et al. 2013). Nevertheless, standardization to develop appropriate scoring methods for immunohistochemistry based PD-L1 expression is still problematic. Kaunitz et al. reported the low PD-L1 tumor expression in ALM, compared to CDS melanoma (Kaunitz et al. 2017). This might suggest that

917

immunotherapy by anti-PD-1 therapy is not more effective in ALM than other types of CM. Finally, we need further investigations to evaluate the true efficacy of immunotherapy in ALM and the specific immune biology of ALM.

Targeted Therapy There are a limited number of ALM patients who obtain the benefits from targeted therapy. This is due to the much lower frequency of BARF mutations in ALM, than in CM. However, several new molecular targets are being researched, which are promising therapies for metastatic or unresectable ALM (see also chapter ▶ “Molecularly Targeted Therapy for Patients with BRAF Wild-Type Melanoma”). Tyrosine kinase inhibitors: KIT mutations are detected in a certain number of ALM patients (0–13.8%) (Lin et al. 2013; Kong et al. 2011; Zhou et al. 2012; Yun et al. 2011; Shim et al. 2017; Jin et al. 2013; Ashida et al. 2009; Ashida et al. 2012). There are several phase 2 trials of KIT (e.g., imatinib, dasatinib, nilotinib, or masatinib) for melanoma with KIT amplifications and/or mutations (Hodi et al. 2013; Guo et al. 2011). Enrolled patients in these trials include ALM patients. In phase 2 trial of imatinib for mucosal, acral, and CSD melanoma with KIT mutation or amplification, responses to treatment occurred in 53.8% of patients with KIT mutations, although 92.3% of patients acquired resistance within 1 year. Conversely, KIT amplification does not appear to regulate imatinib sensitivity (Hodi et al. 2013). These trials suggest that KIT inhibitors act as a favorable therapeutic strategy in advanced melanoma patients possessing KIT aberrations, especially KIT mutations. However, more clinical trials are needed to obtain a definite answer for efficacy and safety. Sunitinib is a multi-targeted orally administered inhibitor of tyrosine kinases, including VEGF receptor, KIT, platelet-derived growth factor receptor, RET, colony-stimulating factor 1 receptor, and fms-related tyrosine kinase 3 (Imbulgoda et al. 2014). A multicenter phase 2 trial of sunitinib was performed in patients with

918

unresectable stage III or IV mucosal melanoma or ALM. Sunitinib showed response for mucosal melanoma and ALM regardless of the presence of KIT mutation. However, the medication was poorly tolerated, and there were no prolonged responses (Buchbinder et al. 2015). MEK inhibitors: There are no therapies targeted at NRAS directly at present. However, MEK inhibitors are being researched as new NRAS targeted therapeutic strategies. Binimetinib is a MEK1/2 inhibitor. A phase 3 study showed that binimetinib improved the progression-free survival compared with dacarbazine in NRASmutant melanoma and binimetinib was tolerable. Binimetinib might become a new treatment option in patients with NRAS-mutant ALM after failure of immunotherapy (Dummer et al. 2017). The combination of MEK inhibitors with RAF, EGFR-PI3K-AKT, and CDK4/6 inhibitors is currently being evaluated in several clinical trials. These combination therapies are involved in two particular pathways. Therefore, it holds out a promise that these combination therapies could achieve the desired effects for NRAS-mutant ALM in the near future.

Conclusion It has been more than 30 years since Reed first described ALM. It has been clarified that ALM has many unique characteristics compared with other subtypes of CM. ALM has more advanced status at diagnosis and poorer prognosis than other types of CM. However, the reason for its aggressive features still remains unknown. Although ALM is the rarest subtype of CM worldwide, there is need to understand the nature of ALM and to establish its original diagnosis and identify treatment modalities for ALM in the future to effectively address its poor prognosis.

References Arrington JH 3rd, Reed RJ, Ichinose H, Krementz ET (1977) Plantar lentiginous melanoma: a distinctive variant of human cutaneous malignant melanoma. Am J Surg Pathol 1:131–143

Y. Teramoto et al. Ashida A, Takata M, Murata H, Kido K, Saida T (2009) Pathological activation of KIT in metastatic tumors of acral and mucosal melanomas. Int J Cancer 124:862–868 Ashida A, Uhara H, Kiniwa Y, Oguchi M, Murata H, Goto Y, Uchiyama A, Ogawa E, Hayashi K, Koga H, Okuyama R (2012) Assessment of BRAF and KIT mutations in Japanese melanoma patients. J Dermatol Sci 66:240–242 Bai X, Kong Y, Chi Z, Sheng X, Cui C, Wang X, Mao L, Tang B, Li S, Lian B, Yan X, Zhou L, Dai J, Guo J, Si L (2017) MAPK pathway and TERT promoter gene mutation pattern and its prognostic value in melanoma patients: a retrospective study of 2,793 cases. Clin Cancer Res 23:6120–6127 Balch CM (1990) Axillary lymph node dissection: differences in goals and techniques when treating melanoma and breast cancer. Surgery 108:118–119 Balch CM, Soong SJ, Smith T, Ross MI, Urist MM, Karakousis CP, Temple WJ, Mihm MC, Barnhill RL, Jewell WR, Wanebo HJ, Desmond R (2001) Long-term results of a prospective surgical trial comparing 2 cm vs. 4 cm excision margins for 740 patients with 1–4 mm melanomas. Ann Surg Oncol 8:101–108 Banfield CC, Redburn JC, Dawber RP (1998) The incidence and prognosis of nail apparatus melanoma. A retrospective study of 105 patients in four English regions. Br J Dermatol 139:276–279 Bello DM, Chou JF, Panageas KS, Brady MS, Coit DG, Carvajal RD, Ariyan CE (2013) Prognosis of acral melanoma: a series of 281 patients. Ann Surg Oncol 20:3618–3625 Blessing K, Kernohan NM, Park KG (1991) Subungual malignant melanoma: clinicopathological features of 100 cases. Histopathology 19:425–429 Boriani F, O’leary F, Tohill M, Orlando A (2014) Acral Lentiginous Melanoma – misdiagnosis, referral delay and 5 years specific survival according to site. Eur Rev Med Pharmacol Sci 18:1990–1996 Bradford PT, Goldstein AM, Mcmaster ML, Tucker MA (2009) Acral lentiginous melanoma: incidence and survival patterns in the United States, 1986–2005. Arch Dermatol 145:427–434 Buchbinder EI, Sosman JA, Lawrence DP, Mcdermott DF, Ramaiya NH, Van Den Abbeele AD, Linette GP, Giobbie-Hurder A, Hodi FS (2015) Phase 2 study of sunitinib in patients with metastatic mucosal or acral melanoma. Cancer 121:4007–4015 Busam KJ, Iversen K, Coplan KC, Jungbluth AA (2001) Analysis of microphthalmia transcription factor expression in normal tissues and tumors, and comparison of its expression with S-100 protein, gp100, and tyrosinase in desmoplastic malignant melanoma. Am J Surg Pathol 25:197–204 Chang JW, Yeh KY, Wang CH, Yang TS, Chiang HF, Wei FC, Kuo TT, Yang CH (2004) Malignant melanoma in Taiwan: a prognostic study of 181 cases. Melanoma Res 14:537–541 Choi YD, Chun SM, Jin SA, Lee JB, Yun SJ (2013) Amelanotic acral melanomas: clinicopathological,

Acral Lentiginous Melanoma BRAF mutation, and KIT aberration analyses. J Am Acad Dermatol 69:700–707 Cirenajwis H, Lauss M, Ekedahl H, Torngren T, Kvist A, Saal LH, Olsson H, Staaf J, Carneiro A, Ingvar C, Harbst K, Hayward NK, Jonsson G (2017) NF1-mutated melanoma tumors harbor distinct clinical and biological characteristics. Mol Oncol 11:438–451 Clark WH Jr, Elder DE, Van Horn M (1986) The biologic forms of malignant melanoma. Hum Pathol 17:443–450 Clarkson JH, Mcallister RM, Cliff SH, Powell B (2002) Subungual melanoma in situ: two independent streaks in one nail bed. Br J Plast Surg 55:165–167 Cohen T, Busam KJ, Patel A, Brady MS (2008) Subungual melanoma: management considerations. Am J Surg 195:244–248 Cohn-Cedermark G, Rutqvist LE, Andersson R, Breivald M, Ingvar C, Johansson H, Jonsson PE, Krysander L, Lindholm C, Ringborg U (2000) Long term results of a randomized study by the Swedish Melanoma Study Group on 2-cm versus 5-cm resection margins for patients with cutaneous melanoma with a tumor thickness of 0.8–2.0 mm. Cancer 89:1495–1501 Cooper C, Arva NC, Lee C, Yelamos O, Obregon R, Sholl LM, Wagner A, Shen L, Guitart J, Gerami P (2015) A clinical, histopathologic, and outcome study of melanonychia striata in childhood. J Am Acad Dermatol 72:773–779 Cress RD, Holly EA (1997) Incidence of cutaneous melanoma among non-Hispanic whites, Hispanics, Asians, and blacks: an analysis of California Cancer Registry data, 1988-93. Cancer Causes Control 8:246–252 Cronin KA, Lake AJ, Scott S (2018) Annual Report to the Nation on the Status of Cancer, part I: National cancer statistics. Cancer 124:2785–2800 Curtin JA, Busam K, Pinkel D, Bastian BC (2006) Somatic activation of KIT in distinct subtypes of melanoma. J Clin Oncol 24:4340–4346 D’angelo SP, Larkin J, Sosman JA, Lebbe C, Brady B, Neyns B, Schmidt H, Hassel JC, Hodi FS, Lorigan P, Savage KJ, Miller WH Jr, Mohr P, Marquez-Rodas I, Charles J, Kaatz M, Sznol M, Weber JS, Shoushtari AN, Ruisi M, Jiang J, Wolchok JD (2017) Efficacy and safety of nivolumab alone or in combination with ipilimumab in patients with mucosal melanoma: a pooled analysis. J Clin Oncol 35:226–235 Daly JM, Berlin R, Urmacher C (1987) Subungual melanoma: a 25-year review of cases. J Surg Oncol 35:107–112 Davis EJ, Johnson DB, Sosman JA, Chandra S (2018) Melanoma: what do all the mutations mean? Cancer 124(17):3490–3499 Desai A, Ugorji R, Khachemoune A (2018) Acral melanoma foot lesions. Part 2: clinical presentation, diagnosis, and management. Clin Exp Dermatol 43:117–123 Dummer R, Schadendorf D, Ascierto PA, Arance A, Dutriaux C, Di Giacomo AM, Rutkowski P, Del Vecchio M, Gutzmer R, Mandala M, Thomas L, Demidov L, Garbe C, Hogg D, Liszkay G, Queirolo P, Wasserman E, Ford J, Weill M, Sirulnik LA, Jehl V, Bozon V, Long GV, Flaherty K (2017) Binimetinib versus dacarbazine in patients with

919 advanced NRAS-mutant melanoma (NEMO): a multicentre, open-label, randomised, phase 3 trial. Lancet Oncol 18:435–445 Duprat JP, Silva DC, Coimbra FJ, Lima IA, Lima EN, Almeida OM, Brechtbuhl ER, Landman G, Scramim AP, Neves RI (2005) Sentinel lymph node biopsy in cutaneous melanoma: analysis of 240 consecutive cases. Plast Reconstr Surg 115:1944–1951 discussion 1952–3 Durbec F, Martin L, Derancourt C, Grange F (2012) Melanoma of the hand and foot: epidemiological, prognostic and genetic features. A systematic review. Br J Dermatol 166:727–739 Faries MB, Thompson JF, Cochran AJ, Andtbacka RH, Mozzillo N, Zager JS, Jahkola T, Bowles TL, Testori A, Beitsch PD, Hoekstra HJ, Moncrieff M, Ingvar C, Wouters M, Sabel MS, Levine EA, Agnese D, Henderson M, Dummer R, Rossi CR, Neves RI, Trocha SD, Wright F, Byrd DR, Matter M, Hsueh E, Mackenzie-Ross A, Johnson DB, Terheyden P, Berger AC, Huston TL, Wayne JD, Smithers BM, Neuman HB, Schneebaum S, Gershenwald JE, Ariyan CE, Desai DC, Jacobs L, Mcmasters KM, Gesierich A, Hersey P, Bines SD, Kane JM, Barth RJ, Mckinnon G, Farma JM, Schultz E, Vidal-Sicart S, Hoefer RA, Lewis JM, Scheri R, Kelley MC, Nieweg OE, Noyes RD, Hoon DSB, Wang HJ, Elashoff DA, Elashoff RM (2017) Completion dissection or observation for sentinel-node metastasis in melanoma. N Engl J Med 376:2211–2222 Fernandes JD, Hsieh R, De Freitas LA, Brandao MA, Lourenco SV, Sangueza M, Nico MM (2015) MAP kinase pathways: molecular roads to primary acral lentiginous melanoma. Am J Dermatopathol 37:892–897 Finley RK 3rd, Driscoll DL, Blumenson LE, Karakousis CP (1994) Subungual melanoma: an eighteen-year review. Surgery 116:96–100 Furney SJ, Turajlic S, Stamp G, Thomas JM, Hayes A, Strauss D, Gavrielides M, Xing W, Gore M, Larkin J, Marais R (2014) The mutational burden of acral melanoma revealed by whole-genome sequencing and comparative analysis. Pigment Cell Melanoma Res 27:835–838 Garbe C, Hauschild A, Volkenandt M, Schadendorf D, Stolz W, Reinhold U, Kortmann RD, Kettelhack C, Frerich B, Keilholz U, Dummer R, Sebastian G, Tilgen W, Schuler G, Mackensen A, Kaufmann R (2008) Evidence-based and interdisciplinary consensus-based German guidelines: systemic medical treatment of melanoma in the adjuvant and palliative setting. Melanoma Res 18:152–160 Gershenwald JE, Tseng CH, Thompson W, Mansfield PF, Lee JE, Bouvet M, Lee JJ, Ross MI (1998) Improved sentinel lymph node localization in patients with primary melanoma with the use of radiolabeled colloid. Surgery 124:203–210 Gillgren P, Drzewiecki KT, Niin M, Gullestad HP, Hellborg H, Mansson-Brahme E, Ingvar C, Ringborg U (2011) 2-cm versus 4-cm surgical excision margins

920 for primary cutaneous melanoma thicker than 2 mm: a randomised, multicentre trial. Lancet 378:1635–1642 Green A, Mccredie M, Mackie R, Giles G, Young P, Morton C, Jackman L, Thursfield V (1999) A case-control study of melanomas of the soles and palms (Australia and Scotland). Cancer Causes Control 10:21–25 Grotz TE, Glorioso JM, Pockaj BA, Harmsen WS, Jakub JW (2013) Preservation of the deep muscular fascia and locoregional control in melanoma. Surgery 153:535–541 Gumaste PV, Fleming NH, Silva I, Shapiro RL, Berman RS, Zhong J, Osman I, Stein JA (2014) Analysis of recurrence patterns in acral versus nonacral melanoma: should histologic subtype influence treatment guidelines? J Natl Compr Cancer Netw 12:1706–1712 Guo J, Si L, Kong Y, Flaherty KT, Xu X, Zhu Y, Corless CL, Li L, Li H, Sheng X, Cui C, Chi Z, Li S, Han M, Mao L, Lin X, Du N, Zhang X, Li J, Wang B, Qin S (2011) Phase II, open-label, single-arm trial of imatinib mesylate in patients with metastatic melanoma harboring c-Kit mutation or amplification. J Clin Oncol 29:2904–2909 Guy GP Jr, Thomas CC, Thompson T, Watson M, Massetti GM, Richardson LC (2015) Vital signs: melanoma incidence and mortality trends and projections – United States, 1982–2030. MMWR Morb Mortal Wkly Rep 64:591–596 Haneke E (2006) Surgical anatomy of the nail apparatus. Dermatol Clin 24:291–296 Haneke E (2012) Ungual melanoma – controversies in diagnosis and treatment. Dermatol Ther 25:510–524 Hayashi K, Uhara H, Koga H, Okuyama R, Saida T (2012) Surgical treatment of nail apparatus melanoma in situ: the use of artificial dermis in reconstruction. Dermatol Surg 38:692–694 Hayes AJ, Maynard L, Coombes G, Newton-Bishop J, Timmons M, Cook M, Theaker J, Bliss JM, Thomas JM, Group, U. K. M. S., British Association of, P., Reconstructive, Aesthetic, S. & The Scottish Cancer Therapy, N (2016) Wide versus narrow excision margins for high-risk, primary cutaneous melanomas: longterm follow-up of survival in a randomised trial. Lancet Oncol 17:184–192 Hayward NK, Wilmott JS, Waddell N, Johansson PA, Field MA, Nones K, Patch AM, Kakavand H, Alexandrov LB, Burke H, Jakrot V, Kazakoff S, Holmes O, Leonard C, Sabarinathan R, Mularoni L, Wood S, Xu Q, Waddell N, Tembe V, Pupo GM, De PaoliIseppi R, Vilain RE, Shang P, Lau LMS, Dagg RA, Schramm SJ, Pritchard A, Dutton-Regester K, Newell F, Fitzgerald A, Shang CA, Grimmond SM, Pickett HA, Yang JY, Stretch JR, Behren A, Kefford RF, Hersey P, Long GV, Cebon J, Shackleton M, Spillane AJ, Saw RPM, Lopez-Bigas N, Pearson JV, Thompson JF, Scolyer RA, Mann GJ (2017) Wholegenome landscapes of major melanoma subtypes. Nature 545:175–180 Heidenreich B, Nagore E, Rachakonda PS, Garcia-CasadoZ, Requena C, Traves V, Becker J, Soufir N, Hemminki K, Kumar R (2014) Telomerase reverse

Y. Teramoto et al. transcriptase promoter mutations in primary cutaneous melanoma. Nat Commun 5:3401 High WA, Quirey RA, Guillen DR, Munoz G, Taylor RS (2004) Presentation, histopathologic findings, and clinical outcomes in 7 cases of melanoma in situ of the nail unit. Arch Dermatol 140:1102–1106 Hodi FS, Corless CL, Giobbie-Hurder A, Fletcher JA, Zhu M, Marino-Enriquez A, Friedlander P, Gonzalez R, Weber JS, Gajewski TF, O’day SJ, Kim KB, Lawrence D, Flaherty KT, Luke JJ, Collichio FA, Ernstoff MS, Heinrich MC, Beadling C, Zukotynski KA, Yap JT, Van Den Abbeele AD, Demetri GD, Fisher DE (2013) Imatinib for melanomas harboring mutationally activated or amplified KIT arising on mucosal, acral, and chronically sun-damaged skin. J Clin Oncol 31:3182–3190 Hodis E, Watson IR, Kryukov GV, Arold ST, Imielinski M, Theurillat JP, Nickerson E, Auclair D, Li L, Place C, Dicara D, Ramos AH, Lawrence MS, Cibulskis K, Sivachenko A, Voet D, Saksena G, Stransky N, Onofrio RC, Winckler W, Ardlie K, Wagle N, Wargo J, Chong K, Morton DL, Stemke-Hale K, Chen G, Noble M, Meyerson M, Ladbury JE, Davies MA, Gershenwald JE, Wagner SN, Hoon DS, Schadendorf D, Lander ES, Gabriel SB, Getz G, Garraway LA, Chin L (2012) A landscape of driver mutations in melanoma. Cell 150:251–263 Holmstrom H (1992) Surgical management of primary melanoma. Semin Surg Oncol 8:366–369 Hong JW, Lee S, Kim DC, Kim KH, Song KH (2014) Prognostic and clinicopathologic associations of BRAF mutation in primary acral lentiginous melanoma in Korean patients: a preliminary study. Ann Dermatol 26:195–202 Horn S, Figl A, Rachakonda PS, Fischer C, Sucker A, Gast A, Kadel S, Moll I, Nagore E, Hemminki K, Schadendorf D, Kumar R (2013) TERT promoter mutations in familial and sporadic melanoma. Science 339:959–961 Huang FW, Hodis E, Xu MJ, Kryukov GV, Chin L, Garraway LA (2013) Highly recurrent TERT promoter mutations in human melanoma. Science 339:957–959 Hutchinson J (1886) Melanosis often not black: melanotic whitlow. Br Med J 1:491–494 Imakado S, Sato H, Hamada K (2008) Two cases of subungual melanoma in situ. J Dermatol 35:754–758 Imbulgoda A, Heng DY, Kollmannsberger C (2014) Sunitinib in the treatment of advanced solid tumors. Recent Results Cancer Res 201:165–184 Ishihara K, Saida T, Otsuka F, Yamazaki N (2008) Statistical profiles of malignant melanoma and other skin cancers in Japan: 2007 update. Int J Clin Oncol 13:33–41 Ito T, Wada M, Nagae K, Nakano-Nakamura M, Nakahara T, Hagihara A, Furue M, Uchi H (2015) Acral lentiginous melanoma: who benefits from sentinel lymph node biopsy? J Am Acad Dermatol 72:71–77 Jakob JA, Bassett RL Jr, Ng CS, Curry JL, Joseph RW, Alvarado GC, Rohlfs ML, Richard J, Gershenwald JE, Kim KB, Lazar AJ, Hwu P, Davies MA (2012) NRAS

Acral Lentiginous Melanoma mutation status is an independent prognostic factor in metastatic melanoma. Cancer 118:4014–4023 Jeon SY, Hong JW, Lee S, Oh SY, Hong YS, Kim KH, Song KH (2014) Long-term survival analysis and clinical follow-up in acral lentiginous malignant melanoma undergoing sentinel lymph node biopsy in korean patients. Ann Dermatol 26:177–183 Jin SA, Chun SM, Choi YD, Kweon SS, Jung ST, Shim HJ, Yun SJ (2013) BRAF mutations and KIT aberrations and their clinicopathological correlation in 202 Korean melanomas. J Invest Dermatol 133:579–582 Jung HJ, Kweon SS, Lee JB, Lee SC, Yun SJ (2013) A clinicopathologic analysis of 177 acral melanomas in Koreans: relevance of spreading pattern and physical stress. JAMA Dermatol 149:1281–1288 Jung JY, Roh HJ, Lee SH, Nam K, Chung KY (2011) Comparison of secondary intention healing and fullthickness skin graft after excision of acral lentiginous melanoma on foot. Dermatol Surg 37:1245–1251 Kapteijn BA, Nieweg OE, Liem I, Mooi WJ, Balm AJ, Muller SH, Peterse JL, Valdes Olmos RA, Hoefnagel CA, Kroon BB (1997) Localizing the sentinel node in cutaneous melanoma: gamma probe detection versus blue dye. Ann Surg Oncol 4:156–160 Karakousis CP (1998) Therapeutic node dissections in malignant melanoma. Ann Surg Oncol 5:473–482 Kato T, Suetake T, Sugiyama Y, Tabata N, Tagami H (1996) Epidemiology and prognosis of subungual melanoma in 34 Japanese patients. Br J Dermatol 134:383–387 Kaunitz GJ, Cottrell TR, Lilo M, Muthappan V, Esandrio J, Berry S, Xu H, Ogurtsova A, Anders RA, Fischer AH, Kraft S, Gerstenblith MR, Thompson CL, Honda K, Cuda JD, Eberhart CG, Handa JT, Lipson EJ, Taube JM (2017) Melanoma subtypes demonstrate distinct PD-L1 expression profiles. Lab Investig 97:1063–1071 Kenady DE, Brown BW, Mcbride CM (1982) Excision of underlying fascia with a primary malignant melanoma: effect on recurrence and survival rates. Surgery 92:615–618 Khayat D, Rixe O, Martin G, Soubrane C, Banzet M, Bazex JA, Lauret P, Verola O, Auclerc G, Harper P, Banzet P (2003) Surgical margins in cutaneous melanoma (2 cm versus 5 cm for lesions measuring less than 2.1-mm thick). Cancer 97:1941–1946 Kim JY, Jung HJ, Lee WJ, Kim Do W, Yoon GS, Kim DS, Park MJ, Lee SJ (2011) Is the distance enough to eradicate in situ or early invasive subungual melanoma by wide local excision? From the point of view of matrix-to-bone distance for safe inferior surgical margin in Koreans. Dermatology 223:122–123 Kim YC, Lee MG, Choe SW, Lee MC, Chung HG, Cho SH (2003) Acral lentiginous melanoma: an immunohistochemical study of 20 cases. Int J Dermatol 42:123–129 Koga H, Saida T (2011) Revised 3-step dermoscopic algorithm for the management of acral melanocytic lesions. Arch Dermatol 147:741–743 Kong Y, Si L, Zhu Y, Xu X, Corless CL, Flaherty KT, Li L, Li H, Sheng X, Cui C, Chi Z, Li S, Han M, Mao L, Lu A, Guo J (2011) Large-scale analysis of KIT

921 aberrations in Chinese patients with melanoma. Clin Cancer Res 17:1684–1691 Krauthammer M, Kong Y, Ha BH, Evans P, Bacchiocchi A, Mccusker JP, Cheng E, Davis MJ, Goh G, Choi M, Ariyan S, Narayan D, Dutton-Regester K, Capatana A, Holman EC, Bosenberg M, Sznol M, Kluger HM, Brash DE, Stern DF, Materin MA, Lo RS, Mane S, Ma S, Kidd KK, Hayward NK, Lifton RP, Schlessinger J, Boggon TJ, Halaban R (2012) Exome sequencing identifies recurrent somatic RAC1 mutations in melanoma. Nat Genet 44:1006–1014 Kuchelmeister C, Schaumburg-Lever G, Garbe C (2000) Acral cutaneous melanoma in caucasians: clinical features, histopathology and prognosis in 112 patients. Br J Dermatol 143:275–280 Kunishige JH, Brodland DG, Zitelli JA (2012) Surgical margins for melanoma in situ. J Am Acad Dermatol 66:438–444 Lang J, Mackie RM (2005) Prevalence of exon 15 BRAF mutations in primary melanoma of the superficial spreading, nodular, acral, and lentigo maligna subtypes. J Invest Dermatol 125:575–579 Lazar A, Abimelec P, Dumontier C (2005) Full thickness skin graft for nail unit reconstruction. J Hand Surg Br 30:194–198 Lee JH, Choi JW, Kim YS (2011) Frequencies of BRAF and NRAS mutations are different in histological types and sites of origin of cutaneous melanoma: a metaanalysis. Br J Dermatol 164:776–784 Liang WS, Hendricks W, Kiefer J, Schmidt J, Sekar S, Carpten J, Craig DW, Adkins J, Cuyugan L, Manojlovic Z, Halperin RF, Helland A, Nasser S, Legendre C, Hurley LH, Sivaprakasam K, Johnson DB, Crandall H, Busam KJ, Zismann V, Deluca V, Lee J, Sekulic A, Ariyan CE, Sosman J, Trent J (2017) Integrated genomic analyses reveal frequent TERT aberrations in acral melanoma. Genome Res 27:524–532 Liau JY, Tsai JH, Jeng YM, Chu CY, Kuo KT, Liang CW (2014) TERT promoter mutation is uncommon in acral lentiginous melanoma. J Cutan Pathol 41:504–508 Lin YC, Chang YM, Ho JY, Lin HC, Tsai YM, Chiang CP, Wang WM, Gao HW (2013) C-kit expression of melanocytic neoplasm and association with clinicopathological parameters and anatomic locations in Chinese people. Am J Dermatopathol 35:569–575 Lino-Silva LS, Dominguez-Rodriguez JA, AguilarRomero JM, Martinez-Said H, Salcedo-Hernandez RA, Garcia-Perez L, Herrera-Gomez A, CuellarHubbe M (2016) Melanoma in Mexico: clinicopathologic features in a population with predominance of acral lentiginous subtype. Ann Surg Oncol 23:4189–4194 Liu JF, Zhao LR, Lu LJ, Chen L, Liu ZG, Gong X, Liu B (2014) Limb salvage surgery following resection of a melanoma: foot and ankle reconstruction using cutaneous flaps. Oncol Lett 8:1966–1972 Long GV, Menzies AM, Nagrial AM, Haydu LE, Hamilton AL, Mann GJ, Hughes TM, Thompson JF, Scolyer RA, Kefford RF (2011) Prognostic and clinicopathologic

922 associations of oncogenic BRAF in metastatic melanoma. J Clin Oncol 29:1239–1246 Luk NM, Ho LC, Choi CL, Wong KH, Yu KH, Yeung WK (2004) Clinicopathological features and prognostic factors of cutaneous melanoma among Hong Kong Chinese. Clin Exp Dermatol 29:600–604 Lv J, Dai B, Kong Y, Shen X, Kong J (2016) Acral melanoma in Chinese: a clinicopathological and prognostic study of 142 cases. Sci Rep 6:31432 Macerola E, Loggini B, Giannini R, Garavello G, Giordano M, Proietti A, Niccoli C, Basolo F, Fontanini G (2015) Coexistence of TERT promoter and BRAF mutations in cutaneous melanoma is associated with more clinicopathological features of aggressiveness. Virchows Arch 467:177–184 Maekawa T, Komine M, Murata S, Ohtsuki M (2014) Surgical treatment of subungual melanoma: a case treated with bone splitting of the distal phalanx. J Dermatol 41:944–946 Meng X, Huang Z, Teng F, Xing L, Yu J (2015) Predictive biomarkers in PD-1/PD-L1 checkpoint blockade immunotherapy. Cancer Treat Rev 41:868–876 Metzger S, Ellwanger U, Stroebel W, Schiebel U, Rassner G, Fierlbeck G (1998) Extent and consequences of physician delay in the diagnosis of acral melanoma. Melanoma Res 8:181–186 Meyer T, Merkel S, Gohl J, Hohenberger W (2002) Lymph node dissection for clinically evident lymph node metastases of malignant melanoma. Eur J Surg Oncol 28:424–430 Minagawa A, Omodaka T, Okuyama R (2016) Melanomas and mechanical stress points on the plantar surface of the foot. N Engl J Med 374:2404–2406 Moehrle M, Metzger S, Schippert W, Garbe C, Rassner G, Breuninger H (2003) “Functional” surgery in subungual melanoma. Dermatol Surg 29:366–374 Moon KR, Choi YD, Kim JM, Jin S, Shin MH, Shim HJ, Lee JB, Yun SJ (2018) Genetic alterations in primary acral melanoma and acral melanocytic nevus in Korea: common mutated genes show distinct cytomorphological features. J Invest Dermatol 138:933–945 Morton DL, Thompson JF, Cochran AJ, Mozzillo N, Nieweg OE, Roses DF, Hoekstra HJ, Karakousis CP, Puleo CA, Coventry BJ, Kashani-Sabet M, Smithers BM, Paul E, Kraybill WG, Mckinnon JG, Wang HJ, Elashoff R, Faries MB, Group, M (2014) Final trial report of sentinel-node biopsy versus nodal observation in melanoma. N Engl J Med 370:599–609 Morton DL, Wen DR, Wong JH, Economou JS, Cagle LA, Storm FK, Foshag LJ, Cochran AJ (1992) Technical details of intraoperative lymphatic mapping for early stage melanoma. Arch Surg 127:392–399 Nakamura Y, Fujisawa Y, Teramoto Y, Sato S, Yamada K, Sekine K, Fujimoto M, Otsuka F, Yamamoto A (2014) Tumor-to-bone distance of invasive subungual melanoma: an analysis of 30 cases. J Dermatol 41:872–877 Nakamura Y, Ohara K, Kishi A, Teramoto Y, Sato S, Fujisawa Y, Fujimoto M, Otsuka F, Hayashi N, Yamazaki N, Yamamoto A (2015) Effects of

Y. Teramoto et al. non-amputative wide local excision on the local control and prognosis of in situ and invasive subungual melanoma. J Dermatol 42:861–866 Nakamura Y, Tanaka K, Shibata T, Mizusawa J, Mizutani T, Fukuda H, Tsutsumida A, Namikawa K, Takahashi A, Fujisawa T, Yoshikawa S, Matsushita S, Yamazaki N (2018) Confirmatory trial of non-amputative digit preservation surgery in subungual melanoma: JCOG1602 (J-NAIL study). J Clin Oncol 36:515S, abstr TPS9607 Namikawa K, Yamazaki N (2011) Sentinel lymph node biopsy guided by indocyanine green fluorescence for cutaneous melanoma. Eur J Dermatol 21:184–190 NCCN Clinical Practice Guidelines in Oncology (NCCN Guidelines) Melanoma Version2.2018-January 19. https://www.nccn.org/professionals/physician_gls/pdf/ melanoma.pdf Oh BH, Lee SH, Nam KA, Lee HB, Chung KY (2013) Comparison of negative pressure wound therapy and secondary intention healing after excision of acral lentiginous melanoma on the foot. Br J Dermatol 168:333–338 Ohn J, Choe YS, Mun JH (2016) Dermoscopic features of nail matrix nevus (NMN) in adults and children: a comparative analysis. J Am Acad Dermatol 75:535–540 Ohsie SJ, Sarantopoulos GP, Cochran AJ, Binder SW (2008) Immunohistochemical characteristics of melanoma. J Cutan Pathol 35:433–444 Paolino G, Bekkenk MW, Didona D, Eibenschutz L, Richetta AG, Cantisani C, Viti G, Carbone A, Buccini P, De Simone P, Ferrari A, Scali E, Calvieri S, Silipo V, Cigna E, Viti GP, Bottoni U (2016) Is the prognosis and course of acral melanoma related to site-specific clinicopathological features? Eur Rev Med Pharmacol Sci 20:842–848 Phan A, Touzet S, Dalle S, Ronger-Savle S, Balme B, Thomas L (2006) Acral lentiginous melanoma: a clinicoprognostic study of 126 cases. Br J Dermatol 155:561–569 Piliang MP (2011) Acral lentiginous melanoma. Clin Lab Med 31:281–288 Pizzichetta MA, Talamini R, Stanganelli I, Puddu P, Bono R, Argenziano G, Veronesi A, Trevisan G, Rabinovitz H, Soyer HP (2004) Amelanotic/hypomelanotic melanoma: clinical and dermoscopic features. Br J Dermatol 150:1117–1124 Polom K, Murawa D, Rho YS, Spychala A, Murawa P (2012) Skin melanoma sentinel lymph node biopsy using real-time fluorescence navigation with indocyanine green and indocyanine green with human serum albumin. Br J Dermatol 166:682–683 Rashid M, Hussain SS, Aslam R, Illahi I (2003) A comparison of two fasciocutaneous flaps in the reconstruction of defects of the weight-bearing heel. J Coll Physicians Surg Pak 13:216–218 Rawson RV, Johansson PA, Hayward NK, Waddell N, Patch AM, Lo S, Pearson JV, Thompson JF, Mann GJ, Scolyer RA, Wilmott JS (2017) Unexpected UVR

Acral Lentiginous Melanoma and non-UVR mutation burden in some acral and cutaneous melanomas. Lab Investig 97:130–145 Rayatt SS, Dancey AL, Davison PM (2007) Thumb subungual melanoma: is amputation necessary? J Plast Reconstr Aesthet Surg 60:635–638 Reed RJ (1976) Acral lentiginous melanoma. Wiley, New York Rizvi NA, Hellmann MD, Snyder A, Kvistborg P, Makarov V, Havel JJ, Lee W, Yuan J, Wong P, Ho TS, Miller ML, Rekhtman N, Moreira AL, Ibrahim F, Bruggeman C, Gasmi B, Zappasodi R, Maeda Y, Sander C, Garon EB, Merghoub T, Wolchok JD, Schumacher TN, Chan TA (2015) Cancer immunology. Mutational landscape determines sensitivity to PD-1 blockade in non-small cell lung cancer. Science 348:124–128 Robert C, Long GV, Brady B, Dutriaux C, Maio M, Mortier L, Hassel JC, Rutkowski P, Mcneil C, KalinkaWarzocha E, Savage KJ, Hernberg MM, Lebbe C, Charles J, Mihalcioiu C, Chiarion-Sileni V, Mauch C, Cognetti F, Arance A, Schmidt H, Schadendorf D, Gogas H, Lundgren-Eriksson L, Horak C, Sharkey B, Waxman IM, Atkinson V, Ascierto PA (2015) Nivolumab in previously untreated melanoma without BRAF mutation. N Engl J Med 372:320–330 Saida T, Koga H, Uhara H (2011) Key points in dermoscopic differentiation between early acral melanoma and acral nevus. J Dermatol 38:25–34 Saldanha G, Potter L, Daforno P, Pringle JH (2006) Cutaneous melanoma subtypes show different BRAF and NRAS mutation frequencies. Clin Cancer Res 12:4499–4505 Serpell JW, Carne PW, Bailey M (2003) Radical lymph node dissection for melanoma. ANZ J Surg 73:294–299 Shaw JH, Koea JB (1988) Acral (volar-subungual) melanoma in Auckland, New Zealand. Br J Surg 75:69–72 Shim JH, Shin HT, Park J, Park JH, Lee JH, Yang JM, Kim DH, Jang KT, Lee DY (2017) Mutational profiling of acral melanomas in Korean populations. Exp Dermatol 26:883–888 Shoushtari AN, Munhoz RR, Kuk D, Ott PA, Johnson DB, Tsai KK, Rapisuwon S, Eroglu Z, Sullivan RJ, Luke JJ, Gangadhar TC, Salama AK, Clark V, Burias C, Puzanov I, Atkins MB, Algazi AP, Ribas A, Wolchok JD, Postow MA (2016) The efficacy of anti-PD-1 agents in acral and mucosal melanoma. Cancer 122:3354–3362 Si L, Kong Y, Xu X, Flaherty KT, Sheng X, Cui C, Chi Z, Li S, Mao L, Guo J (2012) Prevalence of BRAF V600E mutation in Chinese melanoma patients: large scale analysis of BRAF and NRAS mutations in a 432-case cohort. Eur J Cancer 48:94–100 Smock ED, Barabas AG, Geh JL (2010) Reconstruction of a thumb defect with Integra following wide local excision of a subungual melanoma. J Plast Reconstr Aesthet Surg 63:e36–e37 Snyder A, Makarov V, Merghoub T, Yuan J, Zaretsky JM, Desrichard A, Walsh LA, Postow MA, Wong P, Ho TS, Hollmann TJ, Bruggeman C, Kannan K, Li Y,

923 Elipenahli C, Liu C, Harbison CT, Wang L, Ribas A, Wolchok JD, Chan TA (2014) Genetic basis for clinical response to CTLA-4 blockade in melanoma. N Engl J Med 371:2189–2199 Stoffels I, Von Der Stuck H, Boy C, Poppel T, Korber N, Weindorf M, Dissemond J, Schadendorf D, Klode J (2012) Indocyanine green fluorescence-guided sentinel lymph node biopsy in dermato-oncology. J Dtsch Dermatol Ges 10:51–57 Sureda N, Phan A, Poulalhon N, Balme B, Dalle S, Thomas L (2011) Conservative surgical management of subungual (matrix derived) melanoma: report of seven cases and literature review. Br J Dermatol 165:852–858 Svedman FC, Pillas D, Taylor A, Kaur M, Linder R, Hansson J (2016) Stage-specific survival and recurrence in patients with cutaneous malignant melanoma in Europe – a systematic review of the literature. Clin Epidemiol 8:109–122 Takata M, Lin J, Takayanagi S, Suzuki T, Ansai S, Kimura T, Cerroni L, Saida T (2007) Genetic and epigenetic alterations in the differential diagnosis of malignant melanoma and spitzoid lesion. Br J Dermatol 156:1287–1294 Takematsu H, Obata M, Tomita Y, Kato T, Takahashi M, Abe R (1985) Subungual melanoma. A clinicopathologic study of 16 Japanese cases. Cancer 55:2725–2731 Taube JM, Klein A, Brahmer JR, Xu H, Pan X, Kim JH, Chen L, Pardoll DM, Topalian SL, Anders RA (2014) Association of PD-1, PD-1 ligands, and other features of the tumor immune microenvironment with response to anti-PD-1 therapy. Clin Cancer Res 20:5064–5074 Teramoto Y, Keim U, Gesierich A, Schuler G, Fiedler E, Tuting T, Ulrich C, Wollina U, Hassel JC, Gutzmer R, Goerdt S, Zouboulis C, Leiter U, Eigentler TK, Garbe C (2018) Acral lentiginous melanoma: a skin cancer with unfavourable prognostic features. A study of the German central malignant melanoma registry (CMMR) in 2050 patients. Br J Dermatol 178:443–451 Teramoto Y, Nakamura Y, Sato S, Yamazaki N, Yamamoto A (2016) Low Probability of Lymphatic Drainage to Cloquet's Node Is of Limited Value as Indicator for Pelvic Lymph Node Dissection in Patients with Lower Limb Melanoma. Lymphat Res Biol 14:109–114 Thomas JM, Newton-Bishop J, A’hern R, Coombes G, Timmons M, Evans J, Cook M, Theaker J, Fallowfield M, O’neill T, Ruka W, Bliss JM, United Kingdom Melanoma Study, G., British Association of Plastic, S. & Scottish Cancer Therapy, N (2004) Excision margins in high-risk malignant melanoma. N Engl J Med 350:757–766 Thumar J, Shahbazian D, Aziz SA, Jilaveanu LB, Kluger HM (2014) MEK targeting in N-RAS mutated metastatic melanoma. Mol Cancer 13:45 Van Der Ploeg IM, Kroon BB, Valdes Olmos RA, Nieweg OE (2009) Evaluation of lymphatic drainage patterns to the groin and implications for the extent of groin dissection in melanoma patients. Ann Surg Oncol 16:2994–2999

924 Vermeeren L, Van Der Ploeg IM, Olmos RA, Meinhardt W, Klop WM, Kroon BB, Nieweg OE (2010) SPECT/CT for preoperative sentinel node localization. J Surg Oncol 101:184–190 Veronesi U, Cascinelli N, Adamus J, Balch C, Bandiera D, Barchuk A, Bufalino R, Craig P, De Marsillac J, Durand JC, Et AL (1988) Thin stage I primary cutaneous malignant melanoma. Comparison of excision with margins of 1 or 3 cm. N Engl J Med 318:1159–1162 Wada M, Ito T, Tsuji G, Nakahara T, Hagihara A, Furue M, Uchi H (2017) Acral lentiginous melanoma versus other melanoma: A single-center analysis in Japan. J Dermatol 44:932–938 Wang Y, Zhao Y, Ma S (2016) Racial differences in six major subtypes of melanoma: descriptive epidemiology. BMC Cancer 16:691 Weber JS, Kudchadkar RR, Yu B, Gallenstein D, Horak CE, Inzunza HD, Zhao X, Martinez AJ, Wang W, Gibney G, Kroeger J, Eysmans C, Sarnaik AA, Chen YA (2013) Safety, efficacy, and biomarkers of nivolumab with vaccine in ipilimumab-refractory or – naive melanoma. J Clin Oncol 31:4311–4318

Y. Teramoto et al. Yamazaki N, Tanaka R, Tsutsumida A, Namikawa K, Eguchi H, Omata W, Oashi K, Ogawa T, Hayashi A, Nakamura N, Tsuta K (2015) BRAF V600 mutations and pathological features in Japanese melanoma patients. Melanoma Res 25:9–14 Yun J, Lee J, Jang J, Lee EJ, Jang KT, Kim JH, Kim KM (2011) KIT amplification and gene mutations in acral/ mucosal melanoma in Korea. APMIS 119:330–335 Zebary A, Omholt K, Vassilaki I, Hoiom V, Linden D, Viberg L, Kanter-Lewensohn L, Johansson CH, Hansson J (2013) KIT, NRAS, BRAF and PTEN mutations in a sample of Swedish patients with acral lentiginous melanoma. J Dermatol Sci 72:284–289 Zhang N, Wang L, Zhu GN, Sun DJ, He H, Luan Q, Liu L, Hao F, Li CY, Gao TW (2014) The association between trauma and melanoma in the Chinese population: a retrospective study. J Eur Acad Dermatol Venereol 28:597–603 Zhou QM, Li W, Zhang X, Chen YB, Chen XC, Guan YX, Ding Y, Wen XZ, Xia Q, Zhou Q, Peng RQ, Hou JH, Zhu XF, Zeng YX, Zhang XS (2012) The mutation profiles of common oncogenes involved in melanoma in southern China. J Invest Dermatol 132:1935–1937

Lentigo Maligna Melanoma Cristian Navarrete-Dechent, Kelly C. Nelson, Anthony M. Rossi, Erica H. Lee, Christopher A. Barker, Kishwer S. Nehal, and Susan M. Swetter

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 927 Epidemiology and Risk Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 927 Natural Course . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 927

C. Navarrete-Dechent Melanoma and Skin Cancer Unit, Department of Dermatology, Facultad de Medicina, Pontificia Universidad Catolica de Chile, Santiago, Chile Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA e-mail: [email protected] K. C. Nelson Department of Dermatology, MD Anderson Cancer Center, The University of Texas, Houston, TX, USA e-mail: [email protected] A. M. Rossi · E. H. Lee · K. S. Nehal Dermatology Service, Memorial Sloan Kettering Cancer Center, New York, NY, USA e-mail: [email protected]; [email protected]; [email protected] C. A. Barker Department of Radiation Oncology, Memorial Sloan Kettering Cancer Center, New York, NY, USA e-mail: [email protected] S. M. Swetter (*) Department of Dermatology, Pigmented Lesion and Melanoma Program, Stanford University Medical Center and Cancer Institute, Stanford, CA, USA Dermatology Service, Veterans Affairs Palo Alto Health Care System, Palo Alto, CA, USA e-mail: [email protected] © This is a U.S. Government work and not under copyright protection in the U.S.; foreign copyright protection may apply 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_68

925

926

C. Navarrete-Dechent et al. Clinical Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928 Dermoscopic Features of Lentigo Maligna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 929 Reflectance Confocal Microscopy Diagnostic Features of Lentigo Maligna . . . . . . . . . . 931 Diagnostic Biopsy Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 932 Histopathologic Diagnostic Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 934 Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935 Surgical Modalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936 Standard Wide Excision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936 Surgical Techniques with Complete Peripheral Margin Assessment . . . . . . . . . . . . . . 937 Staged Excision Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 937 Mohs Surgery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 939 Histopathological Challenges Associated with Lentigo Maligna Surgical Margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940 Nonsurgical Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940 Topical Imiquimod . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Neoadjuvant Use of Imiquimod Prior to Surgery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Use of Imiquimod as Primary Therapy for LM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Use of Imiquimod as Adjuvant Therapy for LM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Monitoring for Response and Recurrence During Imiquimod Treatment . . . . . . . . . . . . . Radiation Therapy (Radiotherapy) for LM/LMM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

940 941 941 942 942 942

Long-Term Follow-Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 943 Potential Role of Reflectance Confocal Microscopy for LM Management . . . . . . . . 943 Quality of Life Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 944 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 947

Abstract

Lentigo maligna melanoma is an increasingly common melanoma subtype worldwide, occurring mainly on the head and neck of older, lightskinned individuals. This subtype poses specific challenges in diagnosis and management, including histologic differentiation from surrounding sun-damaged skin with atypical melanocytes, optimizing treatment in anatomically constrained sites where this melanoma subtype tends to occur, and reducing local recurrence rates related to subclinical extensions, all while taking into account quality of life issues and suitability of surgical, topical, and/or radiation therapy in an elderly population. Noninvasive tools such as dermoscopy and reflectance confocal microscopy can facilitate early diagnosis. Treatment options include surgical and nonsurgical modalities which are complicated by frequent subclinical extension of the tumor and

more complex anatomic location. Surgical options include conventional wide excision and techniques that provide exhaustive peripheral margin control, including staged excision with permanent sections and Mohs micrographic surgery, both of which may reduce rates of local recurrence. Nonsurgical therapies include off-label topical imiquimod and radiation therapy, which are generally second-line but may be preferable in certain instances based on patient comorbidities and/or limited life expectancy, quality of life issues, and/or preference for nonsurgical treatment. Recurrence may occur years after initial treatment, necessitating the need for long-term follow-up.

Keywords

Melanoma · Lentigo maligna · Lentigo maligna melanoma · Treatment ·

Lentigo Maligna Melanoma

Dermoscopy · Reflectance confocal microscopy · Diagnosis · Treatment · Staged excision · Mohs micrographic surgery · Imiquimod · Radiation therapy · Quality of life · Follow-up · Recurrence

Introduction Lentigo maligna melanoma (LMM) is a subtype of melanoma arising on chronically sun-damaged skin. It compromises 4–15% of all melanomas but is the most common melanoma affecting the head and neck (Cohen 1995; Cox et al. 1996; Swetter et al. 2005). The lentigo maligna subtype was first described in the 1890s by Jonathan Hutchinson as infective senile freckle (leading to the term “Hutchinson melanotic freckle”) and later as a “precancer” by William Dubreuilh, i.e., Lentigo malin des viellards and Melanose circonscrite precancereuse (Hutchinson 1894; Dubreuilh 1894). This terminology has resulted in confusion regarding classification of LMM as malignant versus benign, which has persisted to this day, with clinicians erroneously classifying melanoma in situ, lentigo maligna type as a “premalignant” melanocytic neoplasm (Kallini et al. 2013). In the English literature, LMM has varied nomenclature depending on the presence or absence of invasion. When there is exclusively in situ disease, it is termed lentigo maligna (LM). If invasion is present, the term lentigo maligna melanoma is used (Pralong et al. 2012; Cohen 1995). This accepted terminology will be employed throughout the chapter. There are many challenges in the management of LM and LMM, including (1) improving early diagnosis and histologic differentiation from surrounding sun-damaged skin with atypical melanocytes (i.e., actinic melanocytic hyperplasia), (2) optimizing treatment in cosmetically and functionally constrained anatomic sites, (3) reducing local recurrence rates in the setting of clinically inapparent (i.e., subclinical) extension, and (4) balancing quality of life issues related to disease and treatment extent and a largely elderly population, including the appropriateness of surgical, topical, and/or radiation therapy (Wildemore et al. 2001; Shin et al. 2017; McLeod et al. 2011).

927

Epidemiology and Risk Factors LM and LMM arise from cumulative ultraviolet radiation (UVR)-induced oncogenic changes and so typically appear on chronically sun-damaged skin of the head and neck (Cox et al. 1996; Swetter et al. 2005). There is an increased incidence in geographic regions with higher UVR exposure (Holman et al. 1980; Newell et al. 1988), and LM/ LMM incidence appears to be rising (Greveling et al. 2016) (see also chapter ▶ “Clinical Epidemiology of Melanoma”). A Netherlands study showed an increase in the incidence from 0.72 to 3.82 per 100,000 person-year for LM and from 0.24 to 1.19 per 100,000 person-year for LMM between the years 1989 and 2013 (Greveling et al. 2016). This subtype of melanoma tends to occur in the older individuals, with a mean age of 65 years and highest rates observed in the seventh and eighth decades of life (Swetter et al. 2005). LM typically presents with a slow, indolent radial expansion phase over a period of years (estimated 5–20 years). Individual risk factors for LM include lightly pigmented skin (Tiodorovic-Zivkovic et al. 2015) and increasing age (Cox et al. 1996; Swetter et al. 2005), with a slight female to male predominance of 1.7:1 in some studies (Chang et al. 1998). Additional risk factors for LM include number of lentigines (OR 15.9), skin cancer history (OR 2.84), and the presence of actinic keratosis, when compared to the superficial spreading melanoma subtype (Kvaskoff et al. 2012). Certain genetic syndromes may increase the risk for developing LM through reduced skin pigmentation (e.g., oculocutaneous albinism) (Stoll et al. 1981) or impaired UVR damage repair (e.g., xeroderma pigmentosum) (Spatz et al. 2001).

Natural Course Although LMM often progresses slowly, once invasive, it is believed to have same prognosis as other melanoma subtypes when adjusted for Breslow thickness, ulceration, and other histologic factors (Florell et al. 2003). The risk of progression from LM to LMM has been debated for decades and remains unknown, as this would

928

C. Navarrete-Dechent et al.

Fig. 1 Varied clinical appearance of lentigo maligna melanoma in the head and neck region (a–d)

necessitate prospective long-term follow-up of a cohort in which LM lesions would be subjected to intensive observation. In the 1980s, Weinstock et al. estimated the progression rate of LM based on age. Risk of progression varied within population data, ranging from 0.03% per year (from 1 to 44 years) to 0.13% (from 45 to 64 years), to 0.14% (from 65 to 74 years). The lifetime risk of developing invasive melanoma if LM was diagnosed at 45 or 65 years old was estimated at 4.7% and 2.2%, respectively (Weinstock and Sober 1987). A more recent analysis from the Netherlands, using cancer registry data from 1989 to 2013, estimated the risk of a subsequent LMM at any anatomical location after a histologically confirmed LM (Greveling et al. 2016). The cumulative risk of developing an LMM after a primary LM was estimated to be between 2% and 3% after 25 years, with an absolute risk of 2.0–2.6%. However, these data only represent an estimation of the

risk of progression to invasive disease, and larger, prospective analyses are needed.

Clinical Features LM usually presents as an ill-defined, irregularly pigmented macule or patch on sun-exposed areas. Initially similar to surrounding sun-damaged skin, LM slowly “stands out” as a solitary or outlier lesion compared to the background lentigines or seborrheic keratosis (Fig. 1) (Cohen 1995). In contrast to melanomas elsewhere on the body, the “ABCDE rule” may be less applicable for facial locations, particularly when LM mimics a slowly growing solar lentigo (Thomas et al. 1998). LM/LMM most commonly appears on the cheeks and central face in women; on men, it involves the scalp, cartilaginous portions of the ears, and neck (TiodorovicZivkovic et al. 2015; Lesage et al. 2013).

Lentigo Maligna Melanoma

In early stages, the clinical differential diagnosis of LM includes lesions that usually present on the head and neck and/or chronically sun-exposed areas, e.g., solar lentigo, pigmented actinic keratosis, lichen planus-like keratosis, flat seborrheic keratosis, pigmented basal cell carcinoma, melanocytic nevus, and lentigo simplex (see also chapter ▶ “Clinical Presentations of Melanoma”) (Cohen 1995; Lallas et al. 2014; Stolz et al. 2002). One important feature to distinguish LM from pigmented actinic keratosis is the presence of a smooth surface on palpation in the former and a rough and/or scaly surface in the latter (Lallas et al. 2014). Clinically, it may not be possible to make the distinction between LM and its mimickers, necessitating a biopsy to confirm or exclude the diagnosis. Dermoscopy and reflectance confocal microscopy (RCM) have both demonstrated to increase the sensitivity and specificity of LM/LMM diagnosis compared to the visual inspection alone (see below). Lymph node palpation of the draining/regional nodes is also a crucial part in the evaluation of any invasive melanoma, and LMM is not an exception. Treatment for LMM is based on stage and similar to that of other cutaneous melanoma subtypes, as described elsewhere and in melanoma clinical practice guidelines, e.g., NCCN (National Comprehensive Cancer Network).

Dermoscopic Features of Lentigo Maligna The dermoscopic features of melanocytic lesions on facial skin are unique due to specific histologic characteristics of facial, chronically sun-damaged skin. There is flattening or even absence of the rete ridges, and the pigment is only interrupted by hair follicles and adnexal structures ostia, creating a pseudonetwork, contrary to the classic pigmented network seen elsewhere on the body. Moreover, in LM, melanocytes mainly proliferate along the hair follicles and adnexal structures, creating specific dermoscopic patterns that are folliculocentric (Lallas et al. 2014). The dermoscopic features of LM were first described by Schiffner et al. in 2000 (Schiffner

929

et al. 2000). The LM dermoscopic progression model shows the evolution of disease from an early atypical melanocytic proliferation to outright LMM (Fig. 2). Dermoscopic criteria associated with LM include initial pigmentation in and around hair follicles (circles, semicircles, and circles-within-circles; usually with a grayish hue; Fig. 2a); then, perifollicular gray dots and globules appear (annular-granular pattern; with bluegray color; Fig. 2b); next, rhomboidal structures/ angulated lines are formed in the interfollicular areas, creating rhomboids (Fig. 2c); and finally, pigmented blotches with obliteration of hair follicles (homogeneous areas) occur (Fig. 2d) (Schiffner et al. 2000; Pralong et al. 2012). As these features appear sequentially in LM and are associated with progression, this “4-step” development pathway has been termed “LM progression model” (Schiffner et al. 2000; Stolz et al. 2002; Tanaka et al. 2011). The sensitivity and specificity of this dermoscopic model for the diagnosis of LM is 89% and 96%, respectively (Schiffner et al. 2000). Theoretically, the dermoscopic features described on the LM progression model can help differentiate an early LM from an evolving LM, as the initial features (circles and annular-granular pattern) are more common on early lesions than the latter features (rhomboids and obliteration of hair follicles) (Stolz et al. 2002). The single most important dermoscopic feature for the early detection of LM is the asymmetric pigmentation of follicular openings (Fig. 3a–d) (Schiffner et al. 2000; Stolz et al. 2002; Tschandl et al. 2015). Tschandl et al. showed that more than two-thirds of lesions presenting circles as the main pattern were malignant (Tschandl et al. 2015). Particularly, a double layer of perifollicular circles “circle-within-circle” (also known as isobar sign) has a low sensitivity (4.2–5%) but high specificity (98.1%) for the diagnosis of LM (Fig. 4a–b) (Tschandl et al. 2015; Pralong et al. 2012). Pralong et al. described additional criteria including (1) “target-like pattern” defined as the presence of a dark dot in the center of a dark hyperpigmented follicle; (2) “darkening at dermoscopic examination,” defined as the observation of a darker hue of color invisible to the

930

C. Navarrete-Dechent et al.

Fig. 2 Melanoma dermoscopy progression model. (a) Asymmetric follicular openings and circle-within-a-circle. (b) Annular-granular pattern. (c) Rhomboidal structures. (d) Obliteration of follicular openings

naked eye under dermoscopic view; (3) “red rhomboidal structures,” defined as a rhomboidal vascular pattern occurring in the area separating the hair follicles from the others; and (4) “increased density of the vascular network” defined as a vascular network of higher density within the lesion than in peripheral skin. These latter two vascular criteria may be useful in the rare amelanotic variants of LM (Pralong et al. 2012). Dermoscopic features described for melanomas in other parts of the body that can also be seen on LM include “regression structures” such as peppering (35%) and “white scar-like depigmentation” (10%) (Pralong et al. 2012). The presence of 5 colors, marked pigmented rhomboidal structures, obliterated hair follicles, and red rhomboidal structures was associated with invasion in one study (Pralong et al. 2012). The sole presence of gray color in a pigmented macule should raise the hypothesis of a possible LM (Fig. 4c–d)

(Tiodorovic-Zivkovic et al. 2015; Tschandl et al. 2015). In fact, gray color/structures are present in 88–95% of LMs and should be considered as a high sensitivity but low specificity criteria (Tschandl et al. 2015; TiodorovicZivkovic et al. 2015). It is important to recognize dermoscopic features of benign pigmented lesions that are differential diagnosis of LM. The presence of fingerprint areas, moth-eaten borders, sharp demarcation, and milia-like cysts under dermoscopy is associated with solar lentigo and macular seborrheic keratosis; however, the presence of any of these dermoscopic structures is only relevant in the absence of any of the features described for LM (Fig. 5a) (Schiffner et al. 2000). Finally, pigmented actinic keratosis may show similar dermoscopic features to those of LM, and differentiation between them may not be possible without a biopsy (Akay et al. 2010). Despite not

Lentigo Maligna Melanoma

931

Fig. 3 Lentigo maligna, clinical, and dermoscopic features. (a) Clinical image. (b) Dermoscopic image showing asymmetric pigmentation of follicles (black arrows); asterisk represents the healing biopsy site (polarized light dermoscopy;

original magnification 10X). (c) Clinical image. (d) Dermoscopic image showing annular-granular pattern (black arrows) and rhomboidal structures (black lines) (polarized light dermoscopy; original magnification 10X)

always being helpful, the presence of an “inner gray halo,” shiny white rosettes (only visible with polarized light dermoscopy), or a pigmented network that somewhat spares the hair follicles (“widened” network) has been associated with pigmented actinic keratosis (Fig. 5b) (Nascimento et al. 2014; Tschandl et al. 2015). Finally, most LM criteria were defined in light-skinned individuals, and some ethnic differences are evident which can make diagnosis challenging.

magnification and has the ability to stitch images together creating mosaics that allow “navigation” within the lesional area (Rajadhyaksha et al. 2017). Though not yet widely available, RCM has emerged as a high-quality adjunctive tool for the diagnosis and potential pre- and intraoperative margin assessment of LM. It also aids in the diagnosis of benign pigmented lesions that can mimic LM, as discussed above. A score was developed by Guitera et al. that has shown an excellent sensitivity and specificity for the diagnosis of LM with a range of 93–85% and 61–76%, respectively, depending on the score cut point (Table 1) (Guitera et al. 2010). The hallmark of LM under RCM is the presence of dendritic or round, nucleated, hyperreflective large cells surrounding hair follicles (Fig. 6). These nucleated cells can have pagetoid spread or formation of dermal nests. When this early process evolves,

Reflectance Confocal Microscopy Diagnostic Features of Lentigo Maligna Reflectance confocal microscopy (RCM) is a noninvasive optical imaging modality that allows for the visualization of quasi histological features in vivo. It shows en face images with a 30X

932

C. Navarrete-Dechent et al.

Fig. 4 Lentigo maligna, clinical, and dermoscopic features. (a) Clinical image. (b) Dermoscopic features showing circle-within-a-circle (black arrows) and rhomboidal structures (black lines). The asterisk represents the biopsy site (polarized light dermoscopy; original magnification 10X). (c) Clinical image. (d) Dermoscopic features

showing the presence of gray color thorough the lesion and gray annular-granular pattern (black arrows) and rhomboidal structures (black lines); the asterisk represents suture. (Non-polarized light dermoscopy; original magnification 10X)

there is presence of heavy infiltration of hair follicles that radiates to the periphery forming the so-called caput medusae. Other common RCM findings in LM include an atypical epidermis (atypical honeycomb pattern) and the presence of non-edged papillae (Star and Guitera 2016). Since LM is characterized by the presence of subclinical extension and indistinct clinical borders, RCM has also been used in the surgical setting to delineate these ill-defined margins. This has the potential for more accurate assessment of surgical margins both pre- and intraoperatively as well as to guide the extent of noninvasive therapies such as imiquimod or radiotherapy (Yelamos et al. 2017; Star and Guitera 2016). Another role of RCM is for

surveillance following treatment (both surgical and nonsurgical) to detect early local recurrence (see “Potential Role of Reflectanct Confocal Microscopy for LM Management” below) (Alarcon et al. 2014; Guitera et al. 2014).

Diagnostic Biopsy Techniques The ideal biopsy for melanoma is an excisional/ complete biopsy via elliptical, punch, or saucerization technique, with 1–3 mm margins around the concerning skin lesion, according to NCCN and American Academy of Dermatology guidelines (National Comprehensive Cancer Network, Bichakjian et al. 2011). However,

Lentigo Maligna Melanoma

933

Fig. 5 Dermoscopic features of lentigo maligna mimickers. (a) Solar lentigo showing fingerprint structures (black arrows), moth-eaten sharp demarcated border (red arrows), and symmetrically pigmented follicular openings (white arrows) (polarized light dermoscopy; original Table 1 Guitera et al. (2010) “lentigo maligna score” for reflectance confocal microscopy diagnosis of lentigo maligna

magnification 10X). (b) Pigmented actinic keratosis showing white scale (black arrows), “widened network” (white arrows) and rosettes in the center of the hair follicles (red arrows). (Polarized light dermoscopy; original magnification 10X)

RCM criteria Major criteria Non-edged papillae Round pagetoid cells >20 μm Minor criteria Three or more atypical cells at the DEJ in five 0.5  0.5 mm2 fields* Follicular localization of pagetoid cells and/or atypical junctional cells Nucleated cells within the dermal papillae Broadened honeycomb pattern

Points +2 +2 +1 +1 +1 1

Score of 1 or more points: sensitivity of 93% and specificity of 61% for LM Score of 2 or more points: sensitivity of 85% and specificity of 76% (OR 18.6; 95% CI: 9.3–37.1). (Modified from Guitera et al. 2010) * These atypical cells could be round or dendritic in shape

since LM lesions are usually large and ill-defined in cosmetically sensitive areas, an excisional biopsy is not always feasible. Biopsy techniques include punch biopsy (or multiple punch biopsies) and/or broad shave biopsy(ies) that allow for increased sampling of the lesion, versus a deeper saucerization technique that extends below any potential invasive component. In the setting of a large macular lesion concerning for LM, a broad shave biopsy (or multiple biopsies) may provide optimal histopathologic assessment and exclude microinvasive LMM. Palpation of a suspected LM is useful, as may determine the type of biopsy performed to assess for underlying

invasion (i.e., LMM) or a desmoplastic melanoma component that may be missed with a more superficial procedure. If a palpable lesion is evident, an incisional/partial elliptical/fusiform, punch, or deeper saucerization biopsy is recommended, and this may be accompanied by multiple punch or broad shave biopsies of any surrounding macular component to confirm surrounding melanoma in situ, LM type, and to guide surgical planning (Chen et al. 2013). Figure 7 depicts the different types of biopsies. Again, the key is to maximize histopathologic assessment and microstaging of the melanoma to allow for the most appropriate treatment.

934

C. Navarrete-Dechent et al.

Fig. 6 Reflectance confocal microscopy of lentigo maligna: presence of asymmetric, large, bright, nucleated dendritic cells (red arrows) surrounding a hair follicle (asterisk) (confocal mosaic measuring 1.0  1.2 mm). Insert showing an ill-defined pigmented macule on the right cheek

Fig. 7 Biopsy techniques for lentigo maligna. The red circle represents an excisional biopsy. The reddashed ellipse represents an incisional biopsy oriented along the relaxed skin tension lines. The bluedashed circles represents a broad shave/saucerized biopsy at two sites. The yellow-dashed circles represent multiple punch biopsies. (Image courtesy of Konstantinos Liopyris, M.D)

A study by Ng et al. showed an increased chance for misdiagnosis (OR 16.6; 95% CI [10–27]) with partial punch biopsies when compared to excisional biopsies. This misdiagnosis risk was lower (OR 2.6; 95% CI [1.2–5.7]) with broad shave biopsies due to a larger tissue sample. Punch biopsies were also associated with an incorrect diagnosis leading to negative clinical outcome (OR 20; 95% CI [10–41]). Both punch and shave biopsies from partially sampled LM were associated with inaccurate microstaging

(Ng et al. 2010). Overall, broad shave biopsies may be the best technique for diagnosis in patients with clinically large, macular lesions.

Histopathologic Diagnostic Challenges As described in Histopathology of Melanoma, LM is characterized by an increased density of atypical melanocytes at the dermal-epidermal

Lentigo Maligna Melanoma

junction, although junctional nests and histopathologic features simulating dysplastic nevus may also be evident (Farrahi et al. 2005). Differentiation between LM and melanocytic hyperplasia of sun-damaged skin (i.e., actinic melanocytic hyperplasia) may be difficult, as detailed below. As LM lesions clinically progress, the density, extent of adnexal extension, and nuclear atypia of melanocytes advance as well (Clark and Mihm 1969), with the additional appearance of “starburst giant cells” (multinucleated melanocytes with prominent dendritic processes (Cohen 1996)) and melanophages. An early invasive melanoma component arising within LM may be challenging to identify within the biopsy specimen; superficially invasive disease may be comprised of a few cells and/or histologically mimic an area of dermal nevus cells, particularly in the setting of a partial biopsy/incomplete biopsy. The invasive component may demonstrate a spindled-cell morphology as well, and up to two-thirds of desmoplastic melanomas are diagnosed in association with an overlying LM/LMM (Chen et al. 2013). Immunohistochemical melanocyte antigen markers, such as S100 (anti-s100 antibody), SOX10 (anti-SOX10 antibody), gp100 (HMB45 antibody), Melan-A or MART-1 (antiMelan-A, anti-MART1, or A103 antibodies), microphthalmia-associated transcription factor (MITF) (anti-MITF antibody), tyrosinase (antityrosinase), and nerve growth factor receptor (NGFR) (anti-NGFR antibody), may help identify a subtle dermal invasive component, but must be interpreted with caution to avoid overestimation of melanocytes and possible overinterpretation of melanocytic hyperplasia of sun-damaged skin (Prieto and Shea 2011). No specific markers exist to definitively distinguish melanocytic hyperplasia of sun-damaged skin from early LM. The performance of partial biopsies of large pigmented lesions, particularly on cosmetically sensitive areas, is clinically understandable, but it presents specific histopathologic diagnostic risks. For early lesions of LM, a region may be biopsied that demonstrates more histologically innocuous features supporting an incorrect

935

diagnosis of solar lentigo or actinic melanocytic hyperplasia. For LMM lesions, a region may be biopsied that misses the invasive melanoma component, possibly supporting a less aggressive therapeutic plan. Finally, LM has been demonstrated as collision tumors with basal cell carcinoma and squamous cell carcinoma, raising the possibility that two tumor phenotypes may be present in a single clinically appreciable lesion, and the LM component may be missed with a partial biopsy (Ahlgrimm-Siess et al. 2007; Belisle et al. 2005).

Treatment Wide excision (WE) is considered the standard therapy for cutaneous melanoma on the trunk and extremities, although most randomized controlled trials of surgical margins for invasive melanoma excluded head and neck and acral sites (see also chapter ▶ “Treatment of Primary Melanomas”). Nonetheless, WE remains standard therapy for LMM, with sentinel lymph node biopsy, if indicated. For LM, conventional WE has been associated with higher rates of incomplete excision and local recurrence in the head and neck region. This occurs due to irregular borders of larger clinical lesions and unpredictable subclinical extension and frequent histologic atypia within clinically normal surrounding skin (McLeod et al. 2011; Wildemore et al. 2001). To overcome these limitations, the use of peripheral margin-controlled techniques such as staged excision with permanent sections or Mohs micrographic surgery (MMS) has been developed for LM. However, histolopathologic analysis may be hampered by melanocytic atypia of chronically sun-damaged skin, regardless of the surgical technique employed. Furthermore, for patients who are not surgical candidates due to the anticipated extent of surgery, comorbidities, or patient/family preferences to forego surgery (particularly in very elderly patients), the off-label use of topical therapy with imiquimod cream or radiation therapy can be considered (McLeod et al. 2011). These surgical and nonsurgical treatment approaches are detailed below, including their associated

936

C. Navarrete-Dechent et al.

advantages, limitations, and impact on quality of life (Fosko et al. 2018). When evaluating the efficacy and recurrence rate of any LM treatment, the importance of longterm follow-up (>5 years) cannot be understated. This subtype of melanoma can recur years after initial treatment. Since follow-up time dictates the calculated local recurrence rate in different studies, most published recurrence rates are not comparable (Connolly et al. 2017b). Irrespective of the treatment modality, routine preoperative delineation of LM/LMM clinical margins may be aided by use of a Wood’s lamp, with surgical margins measured beyond the peripheral extent of the lesion as defined under Wood’s lamp examination (Reyes and Robins 1988; Paraskevas et al. 2005).

Surgical Modalities Treatment modalities are summarized in Table 2.

Standard Wide Excision Standard WE is the accepted surgical modality for the treatment of all subtypes of cutaneous melanoma with tissue processed with breadloaf technique (Fig. 8). Current NCCN and AAD guidelines recommend 0.5 cm margins for most melanomas in situ, although this is not based on randomized controlled trial data. However, they recognize the difficulty in achieving histologic clearance of LM and note that wider surgical margins (0.5–1.0 cm) may be necessary, as well as surgical techniques that allow for more

exhaustive peripheral margin assessment, i.e., MMS or staged excision with permanent sections (NCCN). This is due to the high rate of subclinical extension of LM and its tendency to recur if positive or narrow margins are found on histopathology. As noted above, LMM should be treated according to melanoma stage (NCCN). Standard WE with 2 mm from the free margin. The presence of LM (nested or confluent single melanocytes with significant cytological atypia) 1000 kV) photon or electron beam therapy, as well as brachytherapy using radioisotopes. Mechanistically, ionizing radiation causes DNA damage leading to cell injury and death, followed by an immune response, which clinically manifests as dermatitis. The various radiotherapy techniques deliver radiation to varying depths from the skin surface. Because LM and LMM are superficial cutaneous neoplasms, typically a deeply penetrating form of radiotherapy is unnecessary, though a depth of 5 mm has been proposed to ensure extension of radiotherapy below hair follicles (Fogarty et al. 2014). National melanoma treatment guidelines from around the world have indicated that RT for LM and LMM may be an appropriate second-line treatment option in certain situations (National Comprehensive Cancer Network, Marsden et al. 2010; Castro et al. 2015; Guo et al. 2015; Berrocal et al. 2015; Dummer et al. 2015; Pflugfelder et al. 2013; Bichakjian et al. 2011). These recommendations are based on several decades of observational analyses of over 1000 patients treated with a variety of RT techniques. In aggregate, the data suggest that local recurrence after RT occurs in approximately 10% of patients, and lymphatic or distant metastasis occurs in 1%. A major limitation of the data is the short duration of follow-up, which is a reflection of the elderly patient population selected for RT. While the quality of evidence characterizing the effectiveness of RT is limited to single institution observational analyses, a prospective randomized trial of RT versus imiquimod is in progress for unresectable LM and will provide valuable data in the future (NCT02394132). At least two studies have reported on the use of adjuvant RT after surgical excision, neither of

943

which clearly demonstrated benefit (Hedblad and Mallbris 2012; Schmid-Wendtner et al. 2000). Several studies have attempted to compare the results of RT to other forms of treatment for LM and LMM with mixed results (Lee et al. 2011; Tsang et al. 1994; Pitman et al. 1979; Panizzon and Guggisberg 1999; Zalaudek et al. 2003). While surgery is regarded as a superior treatment to RT for LM/LMM, no prospective randomized studies have been conducted to fully assess this issue. After the completion of RT, judging the efficacy of treatment can prove challenging. At least one report indicated that the lack of a strong inflammatory response to RT was associated with a higher chance of local recurrence (Hedblad and Mallbris 2012). Resolution of pigmentation is an imperfect indicator of efficacy, and investigational imaging modalities, such as RCM, may prove valuable (Richtig et al. 2015; Hibler et al. 2015a). RCM can be used to follow up this nonsurgically treated group of patients (see “Potential Role of Reflectanct Confocal Microscopy for LM Management” below).

Long-Term Follow-Up Because LM and LMM are associated with slowly progressive disease and delayed local recurrence, long-term follow-up is relevant. The 5-year follow-up threshold typically used for most cancers may not capture the biology of melanoma of the LM type (Connolly et al. 2017a). Recurrences may occur 5–10 years after the initial surgery; therefore patients may need to remain in surveillance for longer periods or even lifelong (Donigan et al. 2018; Connolly et al. 2017b). The advent of RCM may aid in the follow-up of this subgroup of patients (Longo et al. 2011).

Potential Role of Reflectance Confocal Microscopy for LM Management Currently, there is a need to improve margin evaluation in LM/LMM during initial surgical consultation. While various imaging modalities are

944

under investigation (e.g., optical coherence tomography and multiphoton microscopy), RCM is the most studied to date. RCM has been used to estimate and map LM/LMM margins presurgically. In a study, RCM estimated margins were then compared with postoperative final histopathological surgical margins. The RCM estimate significantly correlated with the postoperative assessment (Yelamos et al. 2017). Utilizing RCM presurgically offers the benefit of improved patient counseling and surgical planning, as it helps define the extent of subclinical spread prior to initiating surgery. This informs both the surgeon and the patient to assist in reconstructive design and patient expectations. It may also be useful in situations when precise histopathologic evaluation of margins is not immediately available. While RCM may indicate that surgical margins need to be increased due to subclinical extension, it may also confirm that narrower surgical margins may be adequate to clear the lesion in anatomically sensitive sites. Thus, RCM may provide valuable clinical information to potentially guide surgical management and lead to favorable cosmetic outcomes (Fig. 12). Shortcomings of RCM to date include cost, training, and relative lack of widespread clinical use, although this may change in the future. RCM is emerging as an imaging technology that is also proving useful in the assessment of nonsurgical treatments of LM (Nadiminti et al. 2010; Fogarty et al. 2014; Swetter et al. 2015). Surveillance has been classically performed by clinical examination, without adjunctive imaging. The benefit of RCM after topical off-label imiquimod use is that it represents a noninvasive modality to monitor response to treatment and detects if residual LM remains. Treatment with imiquimod may cause an alteration of the clinically apparent pigment, and it is therefore difficult to assess treatment success by clinical inspection alone (Kai et al. 2016). Examination with RCM my augment the ability to better define the radiation field pretreatment and has been shown to be capable of detecting areas concerning for residual or recurrent disease posttreatment, before clinical repigmentation (Hibler et al. 2015a). Detecting

C. Navarrete-Dechent et al.

LM/LMM recurrences after radiation therapy may be challenging due to the presence of postradiation skin changes. Lesions may recur as clinically amelanotic (not visibly pigmented) or can be further obscured by radiation-induced inflammation and postradiation pigment changes. As RCM allows for the same area of the skin to be re-examined over time, this technology can be used instead of repeat or “scouting” biopsies to monitor for LM recurrence (Erfan et al. 2011; Richtig et al. 2015). However, when using RCM to monitor for recurrence posttreatment for both imiquimod and RT, it is important to wait long enough to ensure any acute changes and inflammation have resolved and will not cause falsepositive interpretation (Vano-Galvan et al. 2013). Other uses of RCM have been described. RCM has been used intraoperatively to provide the surgeon with real-time assessment of tumor margins in vivo (Hibler et al. 2015b). This may be a valuable approach for large LM/LMM lesions that may require surgery under general anesthesia. Using RCM in this mapping fashion could improve clearance of LM and reduce need for reexcision while maximizing tissue conservation and lowering morbidity and recurrence rates. RCM may also be a useful tool in the evaluation of the repigmentation of scars after surgical treatment and can differentiate recurrent LM from other causes of repigmentation such as solar lentigo or pigmented actinic keratosis, hence, avoiding unnecessary biopsies and providing reassurance to patients.

Quality of Life Considerations As discussed earlier, LM and LMM are gradually progressive and have a predilection to develop on sun-exposed areas such as facial skin. They also tend to occur in the older-aged population, with a mean age of 65 years and highest rates observed in the seventh and eighth decades of life (Swetter et al. 2005). LM typically presents with a slow, indolent radial expansion phase over a period of years (estimated 5–20 per some series) and is largely asymptomatic. LM can also have extensive subclinical extension beyond the visible

Lentigo Maligna Melanoma

945

Fig. 12 Presurgical mapping with reflectance confocal microscopy (RCM) in a 65-year-old healthy female with lentigo maligna (erythematous slightly depressed scar following broad shave biopsy). A paper rim is placed 3 mm from the Wood’s lamp clinical margin. RCM is moved

along the inner portion of the ring, showing positive margins in the peripheral margins (insert showing bright, nucleated dendritic cells surrounding hair follicles under RCM). An inside-outside approach is used to determine the subclinical extension. (From Yelamos et al. 2017)

lesion that when surgically removed may lead to a large surgical wound with significant cosmetic implications for reconstruction and, therefore, impact quality of life. Health-related quality of life (HRQOL) in cutaneous melanoma patients is heavily influenced by psychosocial factors. Although most patients cope well, distress from the possibility of disease recurrence and/or metastasis is a major stress factor and has been shown to be independent of disease stage (Loquai et al. 2013). Recurrence anxiety and psychological distress about progression may also increase at the time of dermatologic follow-up, indicating that emotional support is beneficial during visits (Boyle 2003). In a study of melanoma survivors, gender, age, and number of comorbidities were the strongest factors influencing HRQOL (Schlesinger-Raab et al. 2010). Psychological distress is also higher in patients with melanoma in visible areas, such as the face, relevant to many patients with LM/LMM (Kasparian et al. 2009). A systematic review of patient-reported outcome (PRO) measures for HRQOL related to melanoma showed the Short Form-36 (SF-36) and European Organisation for Research and Treatment of Cancer (EORTC QLQ-C30) are commonly used. The

Functional Assessment of Cancer Therapy-Melanoma (FACT-M) is a melanoma-specific PRO measure addressing general symptoms and surgical site symptoms and may be useful in clinical trials (Cornish et al. 2009). These measures do not address fear of recurrence or lifestyle changes deemed as important HRQOL domains (Cornish et al. 2009). More recently, disease-specific measures have been developed such as the Skin Cancer Index, a 15-item questionnaire validated in the nonmelanoma skin cancer population that addresses appearance concerns and worries related to metastatic spread and new skin cancers (Rhee et al. 2006). The FACE-Q Skin Cancer Module is also a new PRO measure developed for facial skin cancer patients. The FACE-Q Skin Cancer Module consists of independently functioning scales addressing facial appearance satisfaction and quality of life; the latter consists of a ten-item Cancer Worry Scale addressing concepts ranging from recurrence and metastatic worry, anxiety, and impact on daily activities and social relationships (Lee et al. 2018). Studies utilizing the newer PRO measures may further define the HRQOL impact of melanoma at diagnosis, treatment, and long-term follow-up.

946

C. Navarrete-Dechent et al.

Fig. 13 Shared decision-making. (a) 85-year-old male with multiple comorbidities with a 1.5 cm diameter biopsy-proven lentigo maligna of the lower eyelid. (b) Reflectance confocal microscopy (RCM) showed large, bright, nucleated dendritic cells in sheets (red arrows) surrounding hair follicles (asterisks) (0.75  0.75 mm). Given patient’s advanced age, comorbidities, and preferences, off-label 5% imiquimod five times per week over 12 weeks was used, resulting in clinical (panel c) and RCM

clearance (panel d) (0.75  0.75 mm). (e) 65-year-old healthy male with a 1.5 cm diameter biopsy-proven lentigo maligna of the lower eyelid. (f) Baseline RCM showed large, bright, nucleated dendritic cells (red arrows) surrounding hair follicles (asterisks) (2  1.5 mm). After discussion of treatment options, patient elected for surgery and reconstruction with an excellent clinical and cosmetic outcome (panel g)

In the elderly population, weighing the risks and benefits of treatment should also be considered given the tendency of LM to be slowly progressive. Therefore, it has been proposed that in the very elderly population (i.e., 85 years and older) given their limited life expectancy, treatment should be deferred in nonmetastatic and slowly progressive tumors (Linos et al. 2013). However, rather than chronological age alone, life expectancy estimates should be patient-based and include comorbidities, functional status, and factors such as patient-related costs, social and family support, duration of treatment, tolerability of therapy, and quality of life (Lee et al. 2015). A shared decision-making approach where the patient is informed of all their options while comprehending the disease and its course will support decisions that are patient-centered (Fig. 13). In the appropriate candidate, nonsurgical options such as off-label topical imiquimod can render excellent disease control and cosmetic outcome compared to surgery. Close observation alone for clinical change

utilizing adjuncts such as dermoscopy and RCM is also an option if the lesion is in a challenging anatomic location, of large size, and deemed that treatment would lead to a decrease in quality of life. However, in the cases of invasive disease (LMM), appropriate treatment should not be delayed in the elderly as older age has been shown to be an independent factor affecting overall survival and poor outcomes in patients with invasive melanoma (Garcovich et al. 2017).

Conclusion Many decades of experience have highlighted the specific challenges of diagnosis and treatment of LM/LMM. Recent data support the use of exhaustive peripheral margin assessment to reduce local recurrence rates following surgical excision of LM/LMM type on anatomically constrained sites, as well as the potential use of other nonsurgical modalities, which may have an adjunctive role to surgery or serve as primary

Lentigo Maligna Melanoma

treatment in patients who are poor surgical candidates. Imaging devices such as RCM can aid in defining clinical margin extent and local recurrence for both surgically and nonsurgically treated cases of LM/LMM. Future research directions for this melanoma subtype include a better understanding of the biology of the tumor and how to predict the risk of invasion and disease progression, as well as how to reduce local recurrence without compromising healthy tissue.

References Agarwal-Antal N, Bowen GM, Gerwels JW (2002) Histologic evaluation of lentigo maligna with permanent sections: implications regarding current guidelines. J Am Acad Dermatol 47:743–748 Ahlgrimm-Siess V, Hofmann-Wellenhof R, Zalaudek I, Cerroni L, Kerl H (2007) Collision of malignant melanoma (lentigo maligna type) with squamous cell carcinoma in solar-damaged skin of the face. Dermatol Surg 33:122–124 Akay BN, Kocyigit P, Heper AO, Erdem C (2010) Dermatoscopy of flat pigmented facial lesions: diagnostic challenge between pigmented actinic keratosis and lentigo maligna. Br J Dermatol 163:1212–1217 Alarcon I, Carrera C, Alos L, Palou J, Malvehy J, Puig S (2014) In vivo reflectance confocal microscopy to monitor the response of lentigo maligna to imiquimod. J Am Acad Dermatol 71:49–55 Belisle A, Gautier MS, Ghozali F, Plantier F, Wechsler J (2005) A collision tumor involving basal cell carcinoma and lentigo maligna melanoma. Am J Dermatopathol 27:319–321 Berrocal A, Arance A, Espinosa E, Castano AG, Cao MG, Larriba JL, Martin JA, Marquez I, Soria A, Algarra SM (2015) SEOM guidelines for the management of malignant melanoma 2015. Clin Transl Oncol 17:1030–1035 Bichakjian CK, Halpern AC, Johnson TM, Foote Hood A, Grichnik JM, Swetter SM, Tsao H, Barbosa VH, Chuang TY, Duvic M, Ho VC, Sober AJ, Beutner KR, Bhushan R, Smith Begolka W, American Academy of Dermatology (2011) Guidelines of care for the management of primary cutaneous melanoma. American Academy of Dermatology. J Am Acad Dermatol 65:1032–1047 Black WH, Thareja SK, Blake BP, Chen R, Cherpelis BS, Glass LF (2011) Distinction of melanoma in situ from solar lentigo on sun-damaged skin using morphometrics and MITF immunohistochemistry. Am J Dermatopathol 33:573–578 Bowen AR, Thacker BN, Goldgar DE, Bowen GM (2011) Immunohistochemical staining with Melan-A of uninvolved sun-damaged skin shows features characteristic of lentigo maligna. Dermatol Surg 37:657–663

947 Boyle DA (2003) Psychological adjustment to the melanoma experience. Semin Oncol Nurs 19:70–77 Bub JL, Berg D, Slee A, Odland PB (2004) Management of lentigo maligna and lentigo maligna melanoma with staged excision: a 5-year follow-up. Arch Dermatol 140:552–558 Castro LG, Messina MC, Loureiro W, Macarenco RS, Duprat Neto JP, di Giacomo TH, Bittencourt FV, Bakos RM, Serpa SS, Stolf HO, Gontijo G (2015) Guidelines of the Brazilian Dermatology Society for diagnosis, treatment and follow up of primary cutaneous melanoma – part I. An Bras Dermatol 90:851–861 Chang AE, Karnell LH, Menck HR (1998) The National Cancer Data Base report on cutaneous and noncutaneous melanoma: a summary of 84,836 cases from the past decade. The American College of Surgeons Commission on Cancer and the American Cancer Society. Cancer 83:1664–1678 Chen LL, Jaimes N, Barker CA, Busam KJ, Marghoob AA (2013) Desmoplastic melanoma: a review. J Am Acad Dermatol 68:825–833 Clark WH Jr, Mihm MC Jr (1969) Lentigo maligna and lentigo-maligna melanoma. Am J Pathol 55:39–67 Cohen LM (1995) Lentigo maligna and lentigo maligna melanoma. J Am Acad Dermatol 33:923–936; quiz 937–940 Cohen LM (1996) The starburst giant cell is useful for distinguishing lentigo maligna from photodamaged skin. J Am Acad Dermatol 35:962–968 Connolly KL, Dusza SW, Nehal KS, Lee EH (2017a) Follow up and recurrence. In: Nehal KS, Busam KJ (eds) Lentigo maligna melanoma. Springer, New York Connolly KL, Hibler BP, Lee EH, Rossi AM, Busam KJ, Nehal KS (2017b) Locally recurrent lentigo maligna and lentigo maligna melanoma: characteristics and time to recurrence after surgery. Dermatol Surg 43:792–797 Cornish D, Holterhues C, van de Poll-Franse LV, Coebergh JW, Nijsten T (2009) A systematic review of healthrelated quality of life in cutaneous melanoma. Ann Oncol 20(Suppl 6):vi51–vi58 Cox NH, Aitchison TC, Sirel JM, Mackie RM (1996) Comparison between lentigo maligna melanoma and other histogenetic types of malignant melanoma of the head and neck. Scottish Melanoma Group. Br J Cancer 73:940–944 Donigan JM, Hyde MA, Goldgar DE, Hadley ML, Bowling M, Bowen GM (2018) Rate of recurrence of lentigo maligna treated with off-label neoadjuvant topical imiquimod, 5%, cream prior to conservatively staged excision. JAMA Dermatol 154(8):885–889 Dummer R, Hauschild A, Lindenblatt N, Pentheroudakis G, Keilholz U, Committee EG (2015) Cutaneous melanoma: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann Oncol 26(Suppl 5): v126–v132 Erfan N, Kang HY, Cardot-Leccia N, Chignon-Sicard B, Passeron T, Ortonne JP, Lacour JP, Bahadoran P (2011) Reflectance confocal microscopy for recurrent lentigo maligna. Dermatol Surg 37:1519–1524

948 Farrahi F, Egbert BM, Swetter SM (2005) Histologic similarities between lentigo maligna and dysplastic nevus: importance of clinicopathologic distinction. J Cutan Pathol 32:405–412 Fleming CJ, Bryden AM, Evans A, Dawe RS, Ibbotson SH (2004) A pilot study of treatment of lentigo maligna with 5% imiquimod cream. Br J Dermatol 151:485–488 Florell SR, Boucher KM, Leachman SA, Azmi F, Harris RM, Malone JC, Martignoni G, Bowen GM, Gerwels JW, Hood AF (2003) Histopathologic recognition of involved margins of lentigo maligna excised by staged excision: an interobserver comparison study. Arch Dermatol 139:595–604 Fogarty GB, Hong A, Scolyer RA, Lin E, Haydu L, Guitera P, Thompson J (2014) Radiotherapy for lentigo maligna: a literature review and recommendations for treatment. Br J Dermatol 170:52–58 Fosko SW, Navarrete-Dechent CP, Nehal KS (2018) Lentigo maligna-challenges, observations, imiquimod, confocal microscopy, and personalized treatment. JAMA Dermatol 154(8):879–881 Garcovich S, Colloca G, Sollena P, Andrea B, Balducci L, Cho WC, Bernabei R, Peris K (2017) Skin cancer epidemics in the elderly as an emerging issue in geriatric oncology. Aging Dis 8:643–661 Giordano CN, Connolly KL, Busam KJ, Nehal KS (2017) Histologic processing and evaluation of surgical excisions specimens. In: Nehal KS, Busam KJ (eds) Lentigo maligna melanoma. Springer, New York Greveling K, Wakkee M, Nijsten T, van den Bos RR, Hollestein LM (2016) Epidemiology of lentigo maligna and lentigo maligna melanoma in the Netherlands, 1989–2013. J Invest Dermatol 136:1955–1960 Guitera P, Pellacani G, Crotty KA, Scolyer RA, Li LX, Bassoli S, Vinceti M, Rabinovitz H, Longo C, Menzies SW (2010) The impact of in vivo reflectance confocal microscopy on the diagnostic accuracy of lentigo maligna and equivocal pigmented and nonpigmented macules of the face. J Invest Dermatol 130(8):2080–2091 Guitera P, Haydu LE, Menzies SW, Scolyer RA, Hong A, Fogarty GB, Gallardo F, Segura S (2014) Surveillance for treatment failure of lentigo maligna with dermoscopy and in vivo confocal microscopy: new descriptors. Br J Dermatol 170:1305–1312 Guo J, Qin S, Liang J, Lin T, Si L, Chen X, Chi Z, Cui C, Du N, Fan Y, Gu K, Li F, Li J, Li Y, Liang H, Liu J, Lu M, Lu A, Nan K, Niu X, Pan H, Ren G, Ren X, Shu Y, Song X, Tao M, Wang B, Wei W, Wu D, Wu L, Wu A, Xu X, Zhang J, Zhang X, Zhang Y, Zhu H, Written & Chinese Society of Clinical Oncology Melanoma Panel (2015) Chinese guidelines on the diagnosis and treatment of melanoma (2015 Edition). Ann Transl Med 3:322 Hazan C, Dusza SW, Delgado R, Busam KJ, Halpern AC, Nehal KS (2008) Staged excision for lentigo maligna and lentigo maligna melanoma: a retrospective analysis of 117 cases. J Am Acad Dermatol 58:142–148 Hedblad MA, Mallbris L (2012) Grenz ray treatment of lentigo maligna and early lentigo maligna melanoma. J Am Acad Dermatol 67:60–68

C. Navarrete-Dechent et al. Hibler BP, Connolly KL, Cordova M, Nehal KS, Rossi AM, Barker CA (2015a) Radiation therapy for synchronous basal cell carcinoma and lentigo maligna of the nose: response assessment by clinical examination and reflectance confocal microscopy. Pract Radiat Oncol 5(5):e543–e547 Hibler BP, Cordova M, Wong RJ, Rossi AM (2015b) Intraoperative real-time reflectance confocal microscopy for guiding surgical margins of lentigo maligna melanoma. Dermatol Surg 41:980–983 Higgins HW 2nd, Lee KC, Galan A, Leffell DJ (2015) Melanoma in situ: part II. Histopathology, treatment, and clinical management. J Am Acad Dermatol 73:193–203; quiz 203–4 Holman CD, Mulroney CD, Armstrong BK (1980) Epidemiology of pre-invasive and invasive malignant melanoma in Western Australia. Int J Cancer 25:317–323 Hou JL, Reed KB, Knudson RM, Mirzoyev SA, Lohse CM, Frohm ML, Brewer JD, Otley CC, Roenigk RK (2015) Five-year outcomes of wide excision and Mohs micrographic surgery for primary lentigo maligna in an academic practice cohort. Dermatol Surg 41:211–218 Huilgol SC, Selva D, Chen C, Hill DC, James CL, Gramp A, Malhotra R (2004) Surgical margins for lentigo maligna and lentigo maligna melanoma: the technique of mapped serial excision. Arch Dermatol 140:1087–1092 Hutchinson J (1894) Lentigo-melanosis. Arch Surg 5:252 Hyde MA, Hadley ML, Tristani-Firouzi P, Goldgar D, Bowen GM (2012) A randomized trial of the off-label use of imiquimod, 5%, cream with vs without tazarotene, 0.1%, gel for the treatment of lentigo maligna, followed by conservative staged excisions. Arch Dermatol 148:592–596 Johnson TM, Headington JT, Baker SR, Lowe L (1997) Usefulness of the staged excision for lentigo maligna and lentigo maligna melanoma: the “square” procedure. J Am Acad Dermatol 37:758–764 Kai AC, Richards T, Coleman A, Mallipeddi R, Barlow R, Craythorne EE (2016) 5 year recurrence rate of lentigo maligna after treatment with imiquimod. Br J Dermatol 174:165–168 Kallini JR, Jain SK, Khachemoune A (2013) Lentigo maligna: review of salient characteristics and management. Am J Clin Dermatol 14:473–480 Kasparian NA, Mcloone JK, Butow PN (2009) Psychological responses and coping strategies among patients with malignant melanoma: a systematic review of the literature. Arch Dermatol 145:1415–1427 Kelley LC, Starkus L (2002) Immunohistochemical staining of lentigo maligna during Mohs micrographic surgery using MART-1. J Am Acad Dermatol 46:78–84 Kopf AW, Bart RS, Gladstein AH (1976) Treatment of melanotic freckle with x-rays. Arch Dermatol 112:801–807 Kunishige JH, Brodland DG, Zitelli JA (2012) Surgical margins for melanoma in situ. J Am Acad Dermatol 66:438–444 Kvaskoff M, Siskind V, Green AC (2012) Risk factors for lentigo maligna melanoma compared with superficial spreading melanoma: a case-control study in Australia. Arch Dermatol 148:164–170

Lentigo Maligna Melanoma Lallas A, Argenziano G, Moscarella E, Longo C, Simonetti V, Zalaudek I (2014) Diagnosis and management of facial pigmented macules. Clin Dermatol 32:94–100 Lawrence CM, Rahim R, Charlton F, Husain A (2014) Prospective study of formalin-fixed Mohs surgery and haematoxylin and eosin stains with control contralateral biopsies for lentigo maligna: 5-year follow-up results. Br J Dermatol 171:298–303 Lee H, Sowerby LJ, Temple CL, Yu E, Moore CC (2011) Carbon dioxide laser treatment for lentigo maligna: a retrospective review comparing 3 different treatment modalities. Arch Facial Plast Surg 13:398–403 Lee EH, Brewer JD, Macfarlane DF (2015) Optimizing informed decision making for basal cell carcinoma in patients 85 years or older. JAMA Dermatol 151:817–818 Lee EH, Klassen AF, Cano SJ, Nehal KS, Pusic AL (2018) FACE-Q Skin Cancer Module for measuring patientreported outcomes following facial skin cancer surgery. Br J Dermatol 179(1):88–94 Lesage C, Barbe C, Le Clainche A, Lesage FX, Bernard P, Grange F (2013) Sex-related location of head and neck melanoma strongly argues for a major role of sun exposure in cars and photoprotection by hair. J Invest Dermatol 133:1205–1211 Linos E, Parvataneni R, Stuart SE, Boscardin WJ, Landefeld CS, Chren MM (2013) Treatment of nonfatal conditions at the end of life: nonmelanoma skin cancer. JAMA Intern Med 173:1006–1012 Longo C, Moscarella E, Pepe P, Cesinaro AM, Casari A, Manfredini M, Stanganelli I, Gardini S, Cota C, Argenziano G, Pellacani G, Zalaudek I (2011) Confocal microscopy of recurrent naevi and recurrent melanomas: a retrospective morphological study. Br J Dermatol 165:61–68 Loquai C, Scheurich V, Syring N, Schmidtmann I, Rietz S, Werner A, Grabbe S, Beutel ME (2013) Screening for distress in routine oncological care-a survey in 520 melanoma patients. PLoS One 8:e66800 Marsden JR, Newton-Bishop JA, Burrows L, Cook M, Corrie PG, Cox NH, Gore ME, Lorigan P, Mackie R, Nathan P, Peach H, Powell B, Walker C, British Association of Dermatologists Clinical Standards Unit (2010) Revised U.K. guidelines for the management of cutaneous melanoma 2010. Br J Dermatol 163: 238–256 Marsden JR, Fox R, Boota NM, Cook M, Wheatley K, Billingham LJ, Steven NM, NCRI Skin Cancer Clinical Studies Group, The U. K. Dermatology, Clinical Trials Network, The Limit-1 Collaborative Group (2017) Effect of topical imiquimod as primary treatment for lentigo maligna: the LIMIT-1 study. Br J Dermatol 176:1148–1154 Mayo C, Johnson TM, Harms KL (2017) Staged excision techniques. In: Nehal KS, Busam KJ (eds) Lentigo maligna melanoma. Springer, New York Mcleod M, Choudhary S, Giannakakis G, Nouri K (2011) Surgical treatments for lentigo maligna: a review. Dermatol Surg 37:1210–1228 Micali G, Lacarrubba F, Nasca MR, Schwartz RA (2014) Topical pharmacotherapy for skin cancer: part I.

949 Pharmacology. J Am Acad Dermatol 70(965):e1–e12; quiz 977–8 Michalopoulos P, Yawalkar N, Bronnimann M, Kappeler A, Braathen LR (2004) Characterization of the cellular infiltrate during successful topical treatment of lentigo maligna with imiquimod. Br J Dermatol 151:903–906 Dubreuilh MW (1894) Lentigo malin des viellards. Ann Dermatol Syphiligr 5:1092 Nadiminti H, Scope A, Marghoob AA, Busam K, Nehal KS (2010) Use of reflectance confocal microscopy to monitor response of lentigo maligna to nonsurgical treatment. Dermatol Surg 36:177–184 Nascimento MM, Shitara D, Enokihara MM, Yamada S, Pellacani G, Rezze GG (2014) Inner gray halo, a novel dermoscopic feature for the diagnosis of pigmented actinic keratosis: clues for the differential diagnosis with lentigo maligna. J Am Acad Dermatol 71:708–715 National Comprehensive Cancer Network. Melanoma (Version 2.2018) [Online]. Available: https://www.nc cn.org/professionals/physician_gls/pdf/melanoma.pdf. Accessed 4 Apr 2018 Newell GR, Sider JG, Bergfelt L, Kripke ML (1988) Incidence of cutaneous melanoma in the United States by histology with special reference to the face. Cancer Res 48:5036–5041 Ng JC, Swain S, Dowling JP, Wolfe R, Simpson P, Kelly JW (2010) The impact of partial biopsy on histopathologic diagnosis of cutaneous melanoma: experience of an Australian tertiary referral service. Arch Dermatol 146:234–239 Osborne JE, Hutchinson PE (2002) A follow-up study to investigate the efficacy of initial treatment of lentigo maligna with surgical excision. Br J Plast Surg 55:611–615 Panizzon RG, Guggisberg D (1999) Clinical aspects and pathology of melanoma. Ther Umsch 56:302–308 Paraskevas LR, Halpern AC, Marghoob AA (2005) Utility of the Wood’s light: five cases from a pigmented lesion clinic. Br J Dermatol 152:1039–1044 Patel AN, Perkins W, Leach IH, Varma S (2014) Johnson square procedure for lentigo maligna and lentigo maligna melanoma. Clin Exp Dermatol 39:570–576 Petratos MA, Kopf AW, Bart RS, Grisewood EN, Gladstein AH (1972) Treatment of melanotic freckle with x-rays. Arch Dermatol 106:189–194 Pflugfelder A, Kochs C, Blum A, Capellaro M, Czeschik C, Dettenborn T, Dill D, Dippel E, Eigentler T, Feyer P, Follmann M, Frerich B, Ganten MK, Gartner J, Gutzmer R, Hassel J, Hauschild A, Hohenberger P, Hubner J, Kaatz M, Kleeberg UR, Kolbl O, Kortmann RD, KrauseBergmann A, Kurschat P, Leiter U, Link H, Loquai C, Loser C, Mackensen A, Meier F, Mohr P, Mohrle M, Nashan D, Reske S, Rose C, Sander C, Satzger I, Schiller M, Schlemmer HP, Strittmatter G, Sunderkotter C, Swoboda L, Trefzer U, Voltz R, Vordermark D, Weichenthal M, Werner A, Wesselmann S, Weyergraf AJ, Wick W, Garbe C, Schadendorf D, German Dermatological Society, DermatologicCooperative Oncology Group (2013) Malignant melanoma S3-guideline “diagnosis, therapy and follow-up of melanoma”. J Dtsch Dermatol Ges 11(Suppl 6):1–116, 1–126

950 Pitman GH, Kopf AW, Bart RS, Casson PR (1979) Treatment of lentigo maligna and lentigo maligna melanoma. J Dermatol Surg Oncol 5:727–737 Powell AM, Robson AM, Russell-Jones R, Barlow RJ (2009) Imiquimod and lentigo maligna: a search for prognostic features in a clinicopathological study with long-term follow-up. Br J Dermatol 160:994–998 Pralong P, Bathelier E, Dalle S, Poulalhon N, Debarbieux S, Thomas L (2012) Dermoscopy of lentigo maligna melanoma: report of 125 cases. Br J Dermatol 167:280–287 Prieto VG, Shea CR (2011) Immunohistochemistry of melanocytic proliferations. Arch Pathol Lab Med 135:853–859 Rajadhyaksha M, Marghoob A, Rossi A, Halpern AC, Nehal KS (2017) Reflectance confocal microscopy of skin in vivo: from bench to bedside. Lasers Surg Med 49:7–19 Rajpar SF, Marsden JR (2006) Imiquimod in the treatment of lentigo maligna. Br J Dermatol 155:653–656 Read T, Noonan C, David M, Wagels M, Foote M, Schaider H, Soyer HP, Smithers BM (2016) A systematic review of non-surgical treatments for lentigo maligna. J Eur Acad Dermatol Venereol 30:748–753 Reyes BA, Robins P (1988) Wood’s lamp and surgical margins in malignant melanoma in situ. J Dermatol Surg Oncol 14:22 Rhee JS, Matthews BA, Neuburg M, Logan BR, Burzynski M, Nattinger AB (2006) Validation of a quality-of-life instrument for patients with nonmelanoma skin cancer. Arch Facial Plast Surg 8:314–318 Richtig E, Arzberger E, Hofmann-Wellenhof R, FinkPuches R (2015) Assessment of changes in lentigo maligna during radiotherapy by in-vivo reflectance confocal microscopy: a pilot study. Br J Dermatol 172:81–87 Schiffner R, Schiffner-Rohe J, Vogt T, Landthaler M, Wlotzke U, Cognetta AB, Stolz W (2000) Improvement of early recognition of lentigo maligna using dermatoscopy. J Am Acad Dermatol 42:25–32 Schlesinger-Raab A, Schubert-Fritschle G, Hein R, Stolz W, Volkenandt M, Holzel D, Engel J (2010) Quality of life in localised malignant melanoma. Ann Oncol 21:2428–2435 Schmid-Wendtner MH, Brunner B, Konz B, Kaudewitz P, Wendtner CM, Peter RU, Plewig G, Volkenandt M (2000) Fractionated radiotherapy of lentigo maligna and lentigo maligna melanoma in 64 patients. J Am Acad Dermatol 43:477–482 Schon MP, Schon M (2006) The small-molecule immune response modifier imiquimod – its mode of action and clinical use in the treatment of skin cancer. Expert Opin Ther Targets 10:69–76 Shah HC (1931) Lentigo maligna: report of one case treated with radium. Am J Cancer 15:1557–1569 Shin TM, Etzkorn JR, Sobanko JF, Margolis DJ, Gelfand JM, Chu EY, Elenitsas R, Shaikh WR, Miller CJ (2017) Clinical factors associated with subclinical spread of in situ melanoma. J Am Acad Dermatol 76:707–713 Spatz A, Giglia-Mari G, Benhamou S, Sarasin A (2001) Association between DNA repair-deficiency and high

C. Navarrete-Dechent et al. level of p53 mutations in melanoma of Xeroderma pigmentosum. Cancer Res 61:2480–2486 Star P, Guitera P (2016) Lentigo maligna, macules of the face, and lesions on sun-damaged skin: confocal makes the difference. Dermatol Clin 34:421–429 Stigall LE, Brodland DG, Zitelli JA (2016) The use of Mohs micrographic surgery (MMS) for melanoma in situ (MIS) of the trunk and proximal extremities. J Am Acad Dermatol 75:1015–1021 Stoll DB, Ruschak P, Kauh Y, Martin J, Luscombe H (1981) Lentigo maligna in a woman with oculocutaneous albinism. Arch Dermatol 117:360–361 Stolz W, Schiffner R, Burgdorf WH (2002) Dermatoscopy for facial pigmented skin lesions. Clin Dermatol 20:276–278 Swetter SM, Boldrick JC, Jung SY, Egbert BM, Harvell JD (2005) Increasing incidence of lentigo maligna melanoma subtypes: northern California and national trends 1990–2000. J Invest Dermatol 125:685–691 Swetter SM, chen FW, Kim DD, Egbert BM (2015) Imiquimod 5% cream as primary or adjuvant therapy for melanoma in situ, lentigo maligna type. J Am Acad Dermatol 72:1047–1053 Tanaka M, Sawada M, Kobayashi K (2011) Key points in dermoscopic differentiation between lentigo maligna and solar lentigo. J Dermatol 38:53–58 Thomas L, Tranchand P, Berard F, Secchi T, Colin C, Moulin G (1998) Semiological value of ABCDE criteria in the diagnosis of cutaneous pigmented tumors. Dermatology 197:11–17 Tiodorovic-Zivkovic D, Argenziano G, Lallas A, Thomas L, Ignjatovic A, Rabinovitz H, Moscarella E, Longo C, Hofmann-Wellenhof R, Zalaudek I (2015) Age, gender, and topography influence the clinical and dermoscopic appearance of lentigo maligna. J Am Acad Dermatol 72:801–808 Tsang RW, Liu FF, Wells W, Payne DG (1994) Lentigo maligna of the head and neck. Results of treatment by radiotherapy. Arch Dermatol 130:1008–1012 Tschandl P, Rosendahl C, Kittler H (2015) Dermatoscopy of flat pigmented facial lesions. J Eur Acad Dermatol Venereol 29:120–127 Vano-Galvan S, Fernandez-Lizarbe E, Truchuelo M, DiazLey B, Grillo E, Sanchez V, Rios-Buceta L, Paoli J, Sancho S, Montero A, Hernanz R, Ramos A, Jaen P, Gonzalez S (2013) Dynamic skin changes of acute radiation dermatitis revealed by in vivo reflectance confocal microscopy. J Eur Acad Dermatol Venereol 27:1143–1150 Weinstock MA, Sober AJ (1987) The risk of progression of lentigo maligna to lentigo maligna melanoma. Br J Dermatol 116:303–310 Wildemore JKT, Schuchter L, Mick R, Synnestvedt M, Elenitsas R, Bedrosian I, Czerniecki BJ, Guerry DT, Lessin SR, Elder DE, Bucky LP (2001) Locally recurrent malignant melanoma characteristics and outcomes: a single-institution study. Ann Plast Surg 46:488–494 Wolf IH, Kodama K, Cerroni L, Kerl H (2007) Nature of inflammatory infiltrate in superficial cutaneous malignancies during topical imiquimod treatment. Am J Dermatopathol 29:237–241

Lentigo Maligna Melanoma Yelamos O, Cordova M, Blank N, Kose K, Dusza SW, Lee E, Rajadhyaksha M, Nehal KS, Rossi AM (2017) Correlation of handheld reflectance confocal microscopy with radial video mosaicing for margin mapping of lentigo maligna and lentigo maligna melanoma. JAMA Dermatol 153:1278–1284 Zalaudek I, Horn M, Richtig E, Hodl S, Kerl H, Smolle J (2003) Local recurrence in melanoma in situ: influence

951 of sex, age, site of involvement and therapeutic modalities. Br J Dermatol 148:703–708 Zitelli JA, Mohs FE, Larson P, Snow S (1989) Mohs micrographic surgery for melanoma. Dermatol Clin 7:833–843 Zitelli JA, Moy RL, Abell E (1991) The reliability of frozen sections in the evaluation of surgical margins for melanoma. J Am Acad Dermatol 24:102–106

Mucosal Melanoma Michael A. Henderson, Charles M. Balch, Claus Garbe, Alexander N. Shoushtari, Bin Lian, Chuanliang Cui, and Jun Guo

Contents Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Epidemiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pathological Features and Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Staging and Prognosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

954 955 956 956

Mucosal Melanoma of the Head and Neck . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Differential Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Staging and Prognosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Treatment Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

957 957 958 958 959

M. A. Henderson (*) Division of Cancer Surgery, Peter MacCallum Cancer Center, Melbourne, VIC, Australia e-mail: [email protected] C. M. Balch Department of Surgical Oncology, The University of Texas MD Anderson Cancer Center, Houston, TX, USA e-mail: [email protected]; [email protected] C. Garbe Centre for Dermatooncology, Department of Dermatology, Eberhard Karls University, Tuebingen, Germany e-mail: [email protected] A. N. Shoushtari Memorial Sloan Kettering Cancer Center, Weill Cornell Medical College, New York, NY, USA e-mail: [email protected] B. Lian · C. Cui · J. Guo Department of Renal Cancer and Melanoma, Key Laboratory of Carcinogenesis and Translational Research, Ministry of Education, Peking University Cancer Hospital and Institute, Beijing, China e-mail: [email protected]; [email protected]; [email protected] © Springer Nature Switzerland AG 2020 C. M. Balch et al. (eds.), Cutaneous Melanoma, https://doi.org/10.1007/978-3-030-05070-2_15

953

954

M. A. Henderson et al. Female Genital Tract Mucosal Melanomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vulvar Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vulvar Melanoma Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vaginal Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vaginal Melanoma Treatment Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cervix and Urethra Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

959 960 961 961 962 962

Mucosal Melanoma of the Penis and Scrotum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 962 Anorectal Mucosal Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 962 Anorectal Melanoma Treatment Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964 Gastrointestinal Tract Melanoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964 Mucosal Melanoma Adjuvant Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964 Mucosal Melanoma Systemic Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 965 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 966

Abstract

Mucosal melanomas are rare tumors which most commonly arise in the upper aerodigestive tract (oral cavity, nasal cavity, and sinuses), anorectum, and female genital tract. Compared to cutaneous melanomas, far less is known about the pathogenesis, natural history, and management of mucosal melanomas. At presentation, mucosal melanomas are characteristically more advanced and associated with poorer outcomes than cutaneous melanomas. The primary treatment modality is complete surgical excision, but due to the anatomical location and advanced stage at presentation, complete removal may not be possible. As the overwhelming majority of patients with locally advanced tumors will die of metastatic disease, highly morbid and/or extensive procedures with major impact on quality of life may not be justified. For patients with advanced disease, immunotherapy with anti PD-1 therapy should be considered and similarly for the small proportion of patients with a c-kit mutation targeted therapy with imatinib may be worthwhile. Unfortunately, the results of treatment for advanced disease do not match those seen for cutaneous melanoma. In this chapter, we review the clinical and pathologic features of mucosal melanomas in general and provide a more detailed discussion concerning the presentation and management of tumors originating in specific anatomical locations. Management of advanced disease is considered separately.

Introduction Primary melanomas arising from the mucosal epithelium lining the respiratory, alimentary, and genitourinary tracts have been well documented but are relatively rare. Unlike their cutaneous counterparts, for which large databases have been established, most reports of mucosal melanoma outcomes are small, retrospective studies. The rarity of these tumors is partially responsible for the fact that insights into the pathogenesis, natural history, and treatment of mucosal melanomas have not kept pace with the advances made in the understanding and treatment of cutaneous melanoma. Guidelines for the management of mucosal melanoma have been published but given the paucity of high-level data, they are consensus-based recommendations only (Ano-uro-genital Mucosal Melanoma Guideline Development Group; Cancer Council Australia Melanoma Guidelines Working Party). By comparison with cutaneous melanomas, mucosal melanomas lack a clear association with ultraviolet exposure, commonly present at a more advanced stage, behave more aggressively, and overall have a much worse prognosis. There has been considerable debate whether these features are the result of an intrinsic biologic aggressiveness (as even small and thin mucosal melanomas can be fatal), the advanced stage at diagnosis due to the clinically occult location of these tumors, or other factors including lack of a dermal/epidermal junction and the richness of the adjacent vascular and lymphatic supply.

Mucosal Melanoma

Recent developments in melanoma tumor biology have highlighted differences between cutaneous and mucosal melanomas (see also chapter ▶ “Molecular Pathology and Genomics of Melanoma”). Unlike cutaneous melanomas, BRAF mutations are uncommonly seen in mucosal melanomas whereas c-KIT amplification or mutations which are rarely seen in cutaneous melanomas may occur in up to half of mucosal melanomas (Curtin et al. 2005, 2006). A large Chinese series found a 10% incidence of c-KIT mutations, somewhat lower than reported from North American populations, but a higher rate of BRAF mutations (12%) (Lian et al. 2017b). Not surprisingly, the UVrelated mutational burden frequently seen in cutaneous melanomas is not prominent in mucosal melanoma. Rather, increased copy number and structural variations predominate (c-KIT, CCND1, and TERT) (Merkel and Gerami 2017). Variations in the frequency of c-KIT mutations by primary site have been noted; they are uncommon in head and neck melanomas but more common in vulvar melanomas (approximately one-third). Another significant feature that separates mucosal melanoma from cutaneous melanoma is the lack of a validated clinico-pathologic staging system. Unlike the current eighth edition AJCC staging system for cutaneous melanoma which is based on 47,000 patients, the staging systems for sinonasal melanoma and vulvar melanoma are not evidence-based and have not been proven to be as useful (Gershenwald et al. 2017b; Verschraegen et al. 2001; Michel et al. 2014; Chae et al. 2016). In many reports, the standard TNM criteria of cutaneous melanoma staging are employed; however, the lack of a dermal-epidermal junction or a suitable identifiable layer deep to mucosa to define tumor thickness is a major issue. In this chapter, we review the clinical and pathologic features of mucosal melanomas in general and provide a more detailed discussion concerning the presentation and management of tumors originating in specific anatomical locations. Adjuvant therapy and management of the patient with recurrent disease will be considered separately.

955

Epidemiology The annual age-adjusted incidence of noncutaneous melanomas was reported by the Third U.S. National Cancer Survey to be 0.7 per 100,000 persons in 1976 (Scotto et al. 1976). In a large population-based study of over 84,000 cases in the US National Cancer Data Base, melanomas arising from mucosal surfaces accounted for 1.3% of all melanomas along with occult primary melanoma 2.2% and ocular melanoma 5.3%, while 91.2% were cutaneous. The majority of mucosal melanomas arose in head and neck sites (55%), followed by female genital (18%), anorectal (24%), and urinary sites (2.8%) (Chang et al. 1998). The incidence of mucosal melanoma from a longitudinal review of the SEER database (1990–2010) was 2.3 per million persons per year, and unlike cutaneous and ocular melanoma, there was no evidence of any change in the incidence over that period (Bishop and Olszewski 2014). The male to female ratio was 0.4:1 predominantly related to the excess of female genitourinary melanomas. Head and neck mucosal melanoma was seen equally frequently in both sexes. The median age at presentation of mucosal melanoma for all sites is in the seventh decade, considerably older than for cutaneous melanoma (67 years vs. 55 years) (Chang et al. 1998). It is most uncommon to see mucosal melanoma presenting in younger persons. The incidence of mucosal melanoma is similar for white and Hispanic persons (Chang et al. 1998). Among black persons, the incidence of mucosal melanoma was approximately two-thirds that seen in white persons, however, because of the much lower incidence of cutaneous melanoma in Blacks, the proportion of persons with AfricanAmerican or Hispanic ethnicity with mucosal melanoma was 8.8% compared to