Insights into the Pharmaceutical and Clinical Applications of Nanoparticles in Cancer Therapy 1527588548, 9781527588547

Healthcare has long been on a quest for a ‘magic bullet’ to cure the dreaded disease cancer. As this book shows, nanopar

239 80 38MB

English Pages 350 [351] Year 2022

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Table of Contents
Chapter 1
Chapter 2
Chapter 3
Chapter 4
Chapter 5
Chapter 6
Chapter 7
Chapter 8
Chapter 9
Chapter 10
Chapter 11
Chapter 12
Recommend Papers

Insights into the Pharmaceutical and Clinical Applications of Nanoparticles in Cancer Therapy
 1527588548, 9781527588547

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Insights into the Pharmaceutical and Clinical Applications of Nanoparticles in Cancer Therapy

Insights into the Pharmaceutical and Clinical Applications of Nanoparticles in Cancer Therapy Edited by

Sheba R. David and Rajan Rajabalaya

Insights into the Pharmaceutical and Clinical Applications of Nanoparticles in Cancer Therapy Edited by Sheba R. David and Rajan Rajabalaya This book first published 2022 Cambridge Scholars Publishing Lady Stephenson Library, Newcastle upon Tyne, NE6 2PA, UK British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Copyright © 2022 by Sheba R. David and Rajan Rajabalaya and contributors All rights for this book reserved. No part of this book may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of the copyright owner. ISBN (10): 1-5275-8854-8 ISBN (13): 978-1-5275-8854-7

TABLE OF CONTENTS

Chapter 1 .................................................................................................... 1 Current Nanodrug Delivery Systems Used for Colorectal Cancer Rajan Rajabalaya and Sheba Rani David Chapter 2 .................................................................................................. 32 Advances in Nanoparticles for Lung Cancer Therapy Rajan Rajabalaya and Stacy David Chapter 3 .................................................................................................. 52 Metal Nanoparticles for Cancer Therapy Rajan Rajabalaya and Sheba Rani David Chapter 4 .................................................................................................. 88 Nanotechnology in Gastroenterology Rajan Rajabalaya and Sheba Rani David Chapter 5 ................................................................................................ 125 Nanomedicines in Oral Cancer Therapies Sheba Rani David and Rajan Rajabalaya Chapter 6 ................................................................................................ 163 Emerging Nanotechnologies for Cancer Immunotherapy Sheba Rani David and Rajan Rajabalaya Chapter 7 ................................................................................................ 200 Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy Sheba Rani David and Rajan Rajabalaya Chapter 8 ................................................................................................ 220 Nanoparticles as Drug Delivery System in Melanoma Sheba Rani David and Rajan Rajabalaya

vi

Table of Contents

Chapter 9 ................................................................................................ 250 Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers Rajan Rajabalaya and Stacy David Chapter 10 .............................................................................................. 279 Advances in Nanomaterials for Enhanced Photodynamic Therapy Monosha Priyadarshini, Dhanashree Murugan, Loganathan Rangasamy, N. Arunai Nambi Raj Chapter 11 .............................................................................................. 310 Nanoparticle-Based Drug Delivery for Bone Disorders Sheba Rani David, Sanjoy Kumar Das and Soumalya Chakraborty Chapter 12 .............................................................................................. 327 Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control Sathish Kumar Gurupatham

CHAPTER 1 CURRENT NANODRUG DELIVERY SYSTEMS USED FOR COLORECTAL CANCER RAJAN RAJABALAYA1 AND SHEBA RANI DAVID*2 1

PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2 School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA *Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract Colorectal cancer is one of the five most widely diagnosed cancers among humans worldwide. This high ranking among different diseases presents an opportunity to treat colorectal cancer by maximising the therapeutic efficacy of anticancer drugs while reducing possible toxicities and side effects. Advancements in cancer drug delivery have led to nanosized particles as drug vehicles to improve therapeutic drug outcomes by translocating drugs to the cancerous targeted sites. This chapter presents different nanodrug delivery systems in cancer therapy, emphasising colorectal cancer. This chapter also summarises critical information on various current nanodrug delivery strategies developed to treat colon cancer, including examples of treatments and their preparation procedures.

Chapter 1

2

Abbreviations 5-FU anti-EGFR anti-HER2 anti-VEGF-A APC AVEX CEA CYP Doxo DR5-NP Dtxl FAP FDA HNPCC LPS mAb NLC OS PEG PFS PLGA PlGF PNP Ptxl RES RESOLV RESS ROS RR SLN

5-fluorouracil Anti-epidermal growth factor receptor Anti-human epidermal growth factor receptor2 Anti-vascular endothelial growth factor-A Adenomatous polyposis coli Avastin in the elderly with Xeloda Carcinoembryonic antigen Cytochrome Doxorubicin Death receptor 5–specific antibodies Docetaxel Familial adenomatous polyposis Food and Drug Administration Hereditary nonpolyposis colorectal cancer Lipopolysaccharides Monoclonal antibodies Nanostructured lipid carriers Overall survival Polyethylene glycol Progression-free survival Polylactic glycolic acid Placental growth factor Polymeric nanoparticles Paclitaxel Reticuloendothelial system Rapid expansion of a supercritical solution into a liquid solvent Rapid expansion of a supercritical solution Reactive oxygen species Resection rate Solid lipid nanoparticle

Current Nanodrug Delivery Systems Used for Colorectal Cancer

3

1. Introduction Cancer is defined as an uncontrolled replication or expansion (neoplasm) of cells that form an abnormal tissue mass known as a tumour [1]. Tumours have abnormal morphologies and/or functions compared to normal cells. They are classified into benign or malignant tumours. Notably, cancer results from a somatic mutation that alters the physiology of normal cells to form malignant tumour cells [2]. Tumour cells found at the outer edge of a mass have the best access to nutrients compared to inside cells. Cells in the inner region rely on diffusion to deliver nutrients and eliminate metabolic waste products. Eventually, cells in the inner region of a tumour mass will die because of an inadequate nutrient supply, resulting in a necrotic core within the tumour. A cancer cell that grows nearby a healthy tissue multiplies faster than other cells, necessitating a higher demand for nutrients from the bloodstream. Healthy tissues cannot compete with cancer cells for nutrient supply in the presence of a tumour. Oxygen, glucose and amino acids are examples of substrates required for the functioning of tumour cells. When a tumour grows in an environment with a limited supply of nutrients, the maximum size of tumour mass reaches approximately 2 mm. Angiogenesis must occur to expand beyond the size of 2 mm, where blood vessels form at the tumour growth site. Angiogenesis is an essential process for the continuous development of a tumour mass. The expansion of a tumour mass above 2 mm may take years to occur. A tumour will reach a steady state if the rate of proliferation of cells equals the rate of cell death [3].

2. Colorectal cancer Colorectal cancer is the fourth most diagnosed cancer among humans worldwide [4]. The incidence rate is significantly higher in the United States of America and Europe than in Africa and Asia [5]. Colorectal cancer is defined by the growth of a malignant tumour in the mucosa of the colon or rectum [4]. Most large bowel cancer cases occur within the pre-existing polyps. About 50% of cases occur in the rectum and 20% in the sigmoid colon. Signs and symptoms of colorectal cancer are the presence of bloody stool, irregular bowel habits, low appetite, reduced body weight, perforation or blockage in the colorectal region [6]. The staging of cancer depends on the size of the primary tumour (T stage), involvement of lymph nodes (N stage) and the incidence of metastases (M stage) [5]. It is categorised into five stages: Stage O, Stage II (IIA, IIB and IIC), Stage III (IIIA, IIIB and IIIC) and Stage IV [4]:

4

Chapter 1

x Stage 0: Stage 0 is the early stage of colorectal cancer [7]. Stage 0 indicates that a polyp or a benign tumour has grown on the mucosal layer [4]. x Stage I: Stage I indicates the invasion of a tumour into the submucosa and muscularis propria layers [4]. x Stage II: Stage II demonstrates the spread of cancer cells beyond the colon but not to the lymphatic system via metastasis [7]. Stage II comprises three sub-divisions (IIA, IIB and IIC). Stage IIA indicates the invasion of a tumour into the peri-colorectal tissues through the muscularis propria. Stage IIB indicates the invasion of a tumour into the visceral peritoneum. Finally, Stage IIC indicates the invasion and adherence of a tumour to other bodily structures or tissues [4]. x Stage III: Stage III depicts the spread of cancer cells throughout the colon wall and surrounding lymphatic nodes [7]. Similar to Stage II, Stage III of colorectal cancer has three sub-divisions, namely, IIIA, IIIB and IIIC. Stage IIIA indicates the invasion of a tumour into the muscularis propria layer and its dispersion in one to three lymphatic nodes or surrounding tissues. Another criterion for Stage IIIA is the tumour invasion into the submucosal layer and its dispersion in four to six lymphatic nodes. Stage IIIB is either indicated by the invasion of a tumour into the muscularis propria and its dispersion in more than six lymphatic nodes or the invasion of a tumour through the muscularis propria into the peri-colorectal tissues and its dispersion in four to six lymphatic nodes. Another criterion for Stage IIIB is the invasion of a tumour into the visceral peritoneum and its dispersion in one to three lymphatic nodes or surrounding tissues. The advanced Stage III cancer, Stage IIIC, is indicated by the invasion of a tumour into the visceral peritoneum and its dispersion in four to six lymphatic nodes. Another criterion for Stage IIIC is the invasion of a tumour into the peri-colorectal tissue and its dispersion in more than six lymphatic nodes or tumour adherence to other bodily structures with tumour dispersion in at least one lymphatic node [4]. x Stage IV: The final stage of colorectal cancer, Stage IV, is indicated by the dispersion of tumour cells at one site, such as in the liver, lungs, ovaries or a non-regional lymphatic node [4].

2.1 Pathophysiology and common risk factors of colorectal cancer Colorectal carcinogenesis manifests in the mucosal lining of the intestinal lumen. If it is left untreated, cancer cells can spread into the muscular layers

Current Nanodrug Delivery Systems Used for Colorectal Cancer

5

underlying the lining and the intestinal wall. At the early stage of colorectal cancer, a polyp can develop into a tumour and eventually enter the mucosal layer’s inner lining. Most colorectal cancer cells have overexpressed carcinoembryonic antigen (CEA) on the cell surface in comparison to normal cells in the colon and biliary epithelial layer. Risk factors of colorectal cancer are age, personal history, lifestyle, race and ethnic group [7]. For instance, an individual's lifestyle based on the Western lifestyle, frequent consumption of red meat from beef or pork and alcohol are often linked to an increased risk of colorectal cancer [5]. Further, environmental conditions, exposure to chemicals, infectious agents, and radiation can contribute to carcinogenesis. Genetic mutation, immune system dysregulation and hormonal factors can also trigger carcinogenesis in an individual. Hereditary predisposition syndrome involves the adenomatous polyposis coli (APC) gene mutation and can cause colorectal cancer. Patients with familial adenomatous polyposis (FAP) gene inherit a mutated APC gene that increases their risk of developing colorectal cancer [7]. The APC gene defect is caused by mismatched DNA formed during repair [5]. Additionally, hereditary nonpolyposis colorectal cancer (HNPCC) is linked with DNA gene mutation, including MLH1, MSH2 and MSH6 genes, accounting for 5% of colorectal cancer cases [5–7]. H. Pylori also shares a positive association with the risk of colon cancer [8]. The loss of regulation of COX-2 expression, a tumorigenesis rate-limiting step in tumorigenesis, occurs early in carcinogenesis [9]. Lipopolysaccharides (LPS) and gastrin are protumorigenic and stimulate the inflammatory pathways that contribute to neutrophil extracellular traps formation [10].

6

Chapter 1

Figure 1-1: Colon cancer pathogenesis and stages of benign and malignant carcinoma. Created with BioRender.com

Current Nanodrug Delivery Systems Used for Colorectal Cancer

7

2.2 Management and treatment of colorectal cancer The possibility of a cure and the survival rate of patients with colorectal cancer is determined by the stage of colorectal cancer [5]. There are various interventions for treating colorectal cancer, such as tumour-removal surgery, radiation therapy, chemotherapy and targeted therapies. A polypectomy procedure can eradicate polyps during colonoscopy in patients diagnosed with Stage 0 colorectal cancer. After the procedure, it would lead to a survival rate of above 90% [7]. After undergoing the primary surgery, a patient with cancer may receive additional therapy, known as adjuvant therapy [11]. Patients with Stage I and Stage II colorectal cancer are expected to have a five-year survival rate post-operation even without administering adjuvant chemotherapy [5]. There is an 80–95% chance of a five-year survival rate for patients with Stage I colorectal cancer. The chance of a five-year survival rate may decrease based on the patient’s prognostic factors, involvement of nodes and the extent of invasiveness of tumour cells [6]. Adjuvant therapy is usually administered to patients diagnosed with colorectal cancer from Stage III to Stage IV to lower the risk of cancer recurrence [5]. Adjuvant therapy with an oral administration of fluoropyrimidine has demonstrated improvement in survival rate after surgical intervention [6]. However, fluoropyrimidine and oxaliplatin or irinotecan co-administration is the standard treatment for patients with metastases. An increase in five-year survival rate by 2–3% by capecitabine or 5-fluorouracil (5-FU) has been shown in patients with Stage II who exhibit no risk factors for colorectal cancer. Additionally, the coadministration of oxaliplatin to capecitabine achieved a gain of 4% to a 3year disease-free survival rate. Patients with a poor prognosis may be administered bevacizumab and capecitabine or infusional 5-FU to increase their overall survival rate. An AVEX (Avastin in Elderly With Xeloda) study including the older population has shown an improvement in the overall survival rate of elderly patients administered with capecitabine (Xeloda) [5]. Capecitabine inhibits the formation of thymidine monophosphate, an essential substrate in DNA synthesis, by the de novo pathway [12]. Patients with Stage III colorectal cancer treated with 5-FU–based regimens have demonstrated an increase in their five-year survival rate by 10–15% [5]. Fluorouracil-based therapy helps improve the survival rate of patients with Stage III colorectal cancer. The administration of the FOLFOX regimen, comprising fluorouracil, folinic acid and oxaliplatin, can improve

8

Chapter 1

the disease-free survival rate without affecting the overall survival rate [6]. For patients at the final stage of colorectal cancer, Stage IV, their treatment options include tumour-removal surgery, chemotherapy and radiation therapy [7]. Patients with Stage IV colorectal cancer have less than a 10% chance in their five-year survival rate; they have a poor prognosis. In a second-line treatment for colorectal cancer, bevacizumab is added to the FOLFOX therapy to improve the survival rate of patients. An antiangiogenic drug, aflibercept, can target various growth factors, such as VEGF-A, VEGF-B and PlGF. The addition of aflibercept to the FOLFIRI therapy (5-FU and irinotecan) shows an overall survival rate after a failure in the initial therapy with oxaliplatin [5]. In a later-line treatment, the foremost objective is to prolong the survival of patients with regimens with minimal toxicity. Although more research is required, studies including new treatments with TS-102 and TS-114 compounds that disrupt thymidylate metabolism may benefit patients in later-line treatment [5]. Fluorouracil infusion is recommended in conventional palliative treatment. An alternative for fluorouracil is its prodrug: capecitabine [13]. Adding folinic acid to the therapy increases its therapeutic efficacy [14]. The main objective of treatment is to reduce the growth and replication of tumour cells. However, it produces a wide range of side effects (e.g., nausea, hair loss and vomiting) because of its poor specificity to target cells. Therefore, multiple cytotoxic drugs are simultaneously administered to patients to reduce side effects. The cancer therapy regimen depends on the patient’s general health condition and requirements [4]. Other essential factors to consider when choosing the proper regimen are the location or type of cancer, patients’ age, consent, preference and stage of cancer [3, 5]. The choice of therapy in a patient with cancer is a critical element to consider by the healthcare provider because each cancer cell responds to each treatment method differently [7]. Each patient’s regimen varies with the medicinal dose, medicinal dosage form, frequency of therapy and cycle duration [4]. In rectal cancer, capecitabine has demonstrated an improvement in the survival rate of patients compared to the 5-FU administration [5]. A significant challenge in cancer drug delivery is an efficient process to the desired cancer tissue without affecting the healthy tissues [15]. Therefore, developing a site-specific delivery system for chemotherapeutic agents would increase the concentration of drugs at the desired or target site. This action would increase the drug’s efficacy while reducing the side effects because only low doses are required [7]. Despite the globally high

Current Nanodrug Delivery Systems Used for Colorectal Cancer

9

prevalence of colorectal cancer, it could be prevented by maintaining a diet rich in minerals containing calcium, vitamin D, curcumin and quercetin [7]. Other primary prevention techniques are increasing the intake of whole grains, fruits and vegetables. Maintaining a proper weight through regular physical activity can help to reduce the colorectal cancer risk in patients with obesity [5].

2.3 Therapeutic approach in metastatic colorectal cancer Approximately 50% of patients with colorectal cancer will develop metastases, contributing to a high mortality rate for colorectal cancer. Fluorouracil treatment (5-FU) has been the first-line treatment for colorectal cancer. The co-administration of 5-FU with cytotoxic agents (e.g., irinotecan and oxaliplatin) has prolonged the survival of patients with metastatic cancer. Intravenous 5-FU can be substituted for capecitabine either as a single agent or in a combination regimen with oxaliplatin. The treatment goals for patients with metastatic colorectal cancer are increasing the survival length of patients, curing patients, alleviating symptoms related to malignancy, and stopping the progression of tumours. The priority established for first-line treatment is an immediate control of tumour growth in patients to decrease metastases before a surgical intervention. The medical intervention of metastatic colorectal cancer also aims to provide a higher resection rate (RR), a more prolonged progression-free survival (PFS) and better overall survival (OS) [16]. The OS rate is the period of survival from the date a patient is diagnosed with the disease [17].

3. Nanodrug delivery systems A drug delivery system delivers the necessary drug to the target site directly or via the systemic circulation in a controlled manner. An advanced drug delivery system functions in a controlled manner to create a personalised treatment for a specific disease while maintaining the drug level within the therapeutic window. Several strategies have been developed to regulate the essential parameters (e.g., rate of drug release, time of drug release and sitedirected drug delivery) that determine the treatment efficacy and mark the beginning of the drug delivery system [18]. Nanodrug delivery systems comprise submicron-sized particles with one or more therapeutic agents dispersed and adsorbed into vesicles, capsules or polymer matrices [15]. Submicron-sized particles are nanoparticles in the nanometre range of 10– 1,000 nm [19]. The diameter of entrapped drugs lies between 10 and 200 nm, ideal for drug uptake and efflux. The most common nanodrug delivery

10

Chapter 1

systems are inorganic nanodrug delivery, lipid-based and polymer nanodrug delivery systems [20]. Nanotechnology has opened the doors to advancing therapeutic and diagnostic strategies, such as selective drug delivery systems, molecular imaging techniques and potential theranostic agents in cancer therapeutics [15]. Theranostic agents can be used to diagnose and treat cancers. Theranostics involve monitoring and assessing drug outcomes in the tumour cells. Hence, the clinical applications of theranostic agents hold great potential for an individualised treatment [13]

4. Clinical application of nanoparticles in cancer therapy The nanoparticle drug delivery system applications are becoming popular because of their modified and site-specific drug release properties [20]. The development of site-specific or targeted drug delivery systems led to greater efficacy of cytotoxic drugs than non-targeted conventional drug delivery systems. Moreover, the nanoparticle drug delivery system enhances the bioavailability of drugs because only a low dose is required to produce an optimal therapeutic effect. For instance, oncology therapy's current available drug delivery systems are based on polymer nanoparticles, lipid nanoparticles, or liposomes. Drug loading into nanoparticles effectively targets tumour sites by targeting moieties or ligands. Moieties or ligands should be specific to overexpressed cancer cell receptors. Site-specific delivery increases the drug uptake by the tumour, resulting in improved chemotherapeutic efficacy. Examples of moieties with a targeting property are antibodies, peptides, oligonucleotides and other small molecules (e.g., folic acid), transferrin and integrin molecules. The first generation of drugs approved by the Food and Drug Administration (FDA) are liposomal doxorubicin (Doxil®) and albumin-bound (i.e., polymer-drug conjugate) paclitaxel (Abraxane®) [15].

5. Characteristics of nanoparticles Nanoparticle formulations should be stable and non-reactive to plasma components in the systemic circulation. The nanoparticle drug vehicle should preserve drug components from degradation in the blood [15]. Small nanoparticles can persist in the leaky defective blood vessels more than large nanoparticles [13]. The ideal size of a nanoparticle for targeted drug delivery is 100 nm or less [3]. Drug uptake into cancerous cells can be enhanced by optimising the size of nanoparticles. The morphology of nanoparticles can influence nanoparticle uptake into tissues because of changes in fluid dynamics. The polarity of nanoparticles can affect their

Current Nanodrug Delivery Systems Used for Colorectal Cancer

11

stability and dispersion in the blood [13]. Nanoparticles have a hydrophilic surface that can resist plasma proteins while allowing the adsorption of surfactants to their surface. Hydrophilic nanoparticles can be prepared from polyvinyl pyrrolidone [3]. The interactions between plasma proteins and nanoparticles can affect the biodistribution and biocompatibility of nanoparticles in the body. The creation of a protein-nanoparticle complex can alter the surface chemistry and size of the nanoparticle. This action will result in nanoparticle identification by macrophages, inducing phagocytosis and eliminating nanoparticles from the bloodstream [21]. Positively charged nanoparticles have demonstrated a targeting property to tumour vessels; however, the neutralisation of nanoparticles after extravasation enables rapid diffusion into the tumour tissue. Hydrophilic nanoparticles can increase the bioavailability of non-watersoluble drugs for effective drug delivery [13]. Additionally, particles with a hydrophobic surface will be subjected to uptake into the liver, spleen and lungs [3]. The reticuloendothelial system (RES) recognises hydrophobic compounds as foreign and removes them from the circulatory system [14]. Moreover, monocytes and macrophages will easily recognise hydrophobic substances coated with opsonin proteins, thus activating phagocytosis. Consequently, nanoparticles cannot penetrate the tumour cells to exert a therapeutic action [13]. The sustained release property of nanoparticles causes drug accumulation at the desired site, enhancing the therapeutic effect [20].

6. Modification of nanoparticles Modifications in nanoparticle formulation are the early steps for transforming (customising) nanoparticles for clinical use [15]. Targeted nanoparticles can be produced using passive or active mechanisms to enhance the chemotherapeutic efficacy of anticancer drugs. The passive targeting mechanism uses the increase in permeability and retention effects from nanoparticle accumulation without altering the surface of nanoparticles. The accumulation of nanoparticles will enable increased therapeutic action at the site of interest. The active targeting mechanism uses the binding of specific ligands to the nanoparticle surface. The surface of nanoparticles can be manipulated with specific moieties or ligands that can prolong retention time and enhance the uptake of nanoparticles into cancerous tissues. Ligands attached to the surface of nanoparticles can interact with specific receptors and antigens expressed on the surface of cancerous cells. Some ligands that can be used are transferrin, folic acid,

12

Chapter 1

antibodies and macromolecules (e.g., proteins and carbohydrates). The ligands’ density should be considered to prevent their elimination by the reticuloendothelial system and protein binding [13]. Nanoparticles conjugated with polyethylene glycol (PEG) have been used to prolong the retention time of nanoparticles in the circulatory system and reduce renal clearance rates [15]. The reduction in renal clearance enhances the bioavailability and increases the nanoparticle half-life to prevent drug degradation before reaching the tumour site. PEG is a hydrophilic and nonionic polymer that can conceal nanoparticles’ hydrophobicity and does not affect the function of charged molecules (e.g., DNA) [13]. Nanoparticles that can transport a large number of drugs are more desirable to produce an optimal therapeutic effect. Some examples of chemotherapeutics that can be conjugated to polymers are paclitaxel (Ptxl), docetaxel (Dtxl) and doxorubicin (Doxo) [22]. Stimuli-responsive nanoparticle drug delivery systems have also been developed to stimulate the drug release upon exposure to certain stimuli. pH level, redox reaction, ionic charges and stress are some examples of physiological and intracellular stimuli. Temperature, light exposure, ultrasound waves, and magnetic and electric fields are some examples of physical and external stimuli. For instance, blood vessels become more permeable at temperatures between 37ºC and 42°C, improving drug delivery by nanoparticles. Lipoproteindelivered benzoporphyrin derivative mono-acid (BPD) verteporfin is an example of light-responsive nanoparticles. Although BPD is biocompatible, it can form reactive oxygen species (ROS) that can destroy DNA, causing the death of cells [13]. Although there are several advantages of multifunctional nanoparticles, issues related to reproducibility and toxicity due to their complicated nature continue to remain a challenge. Therefore, before clinical testing, the pharmacodynamics of the nanoparticle system should be studied thoroughly [15].

7. Types of nanodrug delivery systems 7.1 Liposome A liposome-based nanodrug delivery system comprises vesicles consisting of eight phospholipid bilayers that envelop an inner aqueous phase [15]. The phospholipid bilayer has both hydrophilic and hydrophobic properties that support the encapsulation of water-soluble and lipid-soluble drugs [20]. Water-soluble cytotoxic agents dissolve in the aqueous phase of the liposomes, whereas lipid-soluble cytotoxic agents are entrapped in the lipid bilayer [15]. Liposomes have a special targeting property that can be presented by two

Current Nanodrug Delivery Systems Used for Colorectal Cancer

13

approaches: the passive or active targeting mechanism [2]. A wide range of applications of liposomes is present in gene delivery. The following are examples of lipid reagents used for gene delivery: 1,2bis[oleoyloxy]-3[trimethylammonio]propane; dioctadecylamido-glycylspermine, N-[1-(2,3dioleyloxy)propyl]-N,N,N-trimethylammonium chloride and 3ȕ[N-(N ',N'dimethylaminoethane)-carbamoyl]cholesterol [15]. The liposomes are classified based on the number of bilayers and size of liposomes. The terms used to describe the number of bilayers is unilamellar (i.e., one phospholipid bilayer) and multilamellar (i.e., more than one phospholipid bilayer). Liposomes may be formulated into the size range between 25 nm and 2.5 micrometres. However, drugs that loaded into liposomes with a diameter less than 400 nm can build up in tumour cells more efficiently. However, drugs loaded into liposomes with a diameter of less than 400 nm can build up in tumour cells more efficiently [2]. Because lipids have a high density of cations, they are combined with adjuvant lipids (e.g., cholesterol) to reduce the overall energy required to separate the ionically linked DNA and cationic lipid molecules [15]. Liposomes have a greater cellular retention capacity in the systemic circulation. This trait alters the pharmacokinetics and distribution of P-glycoprotein inhibitors, enabling the saturation of anticancer drugs in tumour cells and enhancing the chemotherapeutic effects. Liposome-based nanodrug delivery systems exhibit no specific targeting. However, this can be achieved by modifying the system using various ligands, including folic acid or anti-transferrin monoclonal antibodies (mAb). Specific cells may detect the infusion of ligands into the phospholipid bilayer of liposomes, rendering selective and modified nanoparticles. The composition of liposomes can be optimised by adding adjuvant lipids or PEG to reverse the incidence of multidrug resistance [20].

7.2 Treatments with liposome-based nanoparticle Doxorubicin restricts the synthesis of nucleic acids within cancer cells [3]. Drug-loaded liposomes are produced in a size that allows their penetration into tumour tissues through pores found in the tumour microvasculature [2]. The liposomes release the doxorubicin that has been entrapped in the cancer cells. Disadvantages of the doxorubicin therapy can be associated with undesirable side effects, such as cardiotoxicity and suppression of bone marrow (i.e., myelosuppression). These effects result in a low therapeutic window for doxorubicin [4]. Conjugating doxorubicin with dextran in chitosan nanoparticles is a strategy to diminish adverse effects by generating a targeted delivery [3]. Liposomal drug delivery can also enhance the

14

Chapter 1

bioavailability of platinum drugs. Examples of cisplatin drugs delivered via the liposome-based nanodrug delivery system for colorectal cancer therapy are LiPlaCis® and Aroplatin® [15]. Cisplatin restricts DNA synthesis by forming intrastrand covalent bonds with DNA bases, which eventually causes DNA damage [23].

7.3 Preparation liposome-based nanoparticle In water, the dispersion of amphiphilic lipids (such as phospholipids) creates liposomes. Liposomes are structurally stable and not bonded covalently compared to polymeric nanoparticles [2]. For example, Doxorubicin (Doxil®) has a lipid bilayer comprising a PEG-modified 1,2,distearoyl-sn-glycero-3-phosphoethanolamine sodium salt, hydrogenated soy phosphatidylcholine and cholesterol [15]. Myocet® comprises doxorubicin citrate surrounded by egg phosphatidylcholine and cholesterol. Doxorubicin is placed in an incubator with liposomes with an acidic core at a neutral pH. A non-ionic drug diffuses down its concentration gradient across the liposome and becomes protonated. Liposomes were initially created using a citric acid buffer (300 mM) to produce an acidic interior. Meanwhile, its exterior is coated with a buffer solution using sodium carbonate and sodium phosphate [24].

7.4 pH-sensitive liposome pH-sensitive liposomes provide another progressive liposomal nanodrug application with great potential in cancer therapy. They are stable at normal physiological pH of 7.4. However, after the internalisation of liposomes in the endosomes of cancer cells, the structure of liposomes becomes less stable because of the acidic pH (e.g., pH: 6.0–6.5). The pH-sensitive liposomes comprises zwitterionic oligopeptide lipid molecules and amino acid-based lipids (e.g., 1,5-dioctadecyl-l-glutamyl 2-histidyl-hexahydrobenzoic acid and 1,5-distearyl N-(N-alpha-(4-mPEG2000) butanedione)-histidyl-lglutamate)). Another advantage of pH-sensitive liposomes is that high concentrations of anticancer agent/cytotoxic agents (e.g., 5-FU) can be incorporated into pHsensitive liposomes compared to conventional liposomes [15]. Poly(styreneco-maleic acid) (SMA) is the main component in pH-responsive liposomes that promotes the delivery of 5-FU within the HT-29 cells of colon cancer [25]. 5-fluorouracil has been shown to interfere with DNA synthesis during the S phase by restricting the action of the thymidylate synthetase enzyme

Current Nanodrug Delivery Systems Used for Colorectal Cancer

15

[26]. Banerjee et al. demonstrated that polystyrene-co-maleic acid undergoes a conformational shift at low pH, causing the destabilisation of the liposomal structure and drug release. Drug-loaded pH-sensitive liposomes fuse with the endovascular membrane because of the acidic pH in cellular endosomes, resulting in a conformational change of the polymer chain. SMA is a copolymer that transforms from a random coil to a collapsed nonionic globular conformation at a low pH. The conformational change causes the destabilisation of liposomes following the formation of channels within the membrane [25]. This action stimulates the release of drug contents from the core of the liposomes. The extracellular environment of solid tumours has an acidic nature, signalling liposomal drug release [15]. The outcome is an increase in intracellular drug availability [25].

7.5 Antibody-coated liposomes Conjugating nanoparticles have developed a site-specific liposomal nanoparticle drug delivery system for cancer therapy with antibodies to produce immunoliposomes. The mechanism of immunoliposomes includes their localisation to the tumour site, where antibodies on the nanoparticle surface bind specifically to overexpressed receptors on the tumour cells. The interaction between tumour cells and immunoliposomes induces enhanced intracellular delivery of the nanoparticle. Examples of antibody moieties used to target specific tumour cells via surface antigen interaction are antihuman epidermal growth factor receptor 2 (anti-HER2), anti-epidermal growth factor receptor (anti-EGFR), anti-CD19 and GAH F(ab)2 goat antihuman monoclonal antibody [15]. Other targeted substances are anti-vascular endothelial growth factor-A (anti-VEGF-A), antiangiogenic multi-kinase inhibitor (e.g., Regorafenib), antiangiogenic compound (e.g., Aflibercept), antibodies including bevacizumab, cetuximab and panitumumab [5].

7.6 Preparation of temperature-sensitive liposome A COOH-PEG3400-PLGA copolymer (w/w) and camptothecin drug are mixed before forming an emulsion. The emulsion is combined with PLGA RG502H, followed by evaporation. The carboxyl groups of copolymer allow the conjugation of DR5-specific antibodies, conatumumab, via their amino groups [27]. An example of a temperature-sensitive liposomal nanodrug delivery system is ThermoDox®. It is formulated with liposomal doxorubicin, releasing the cytotoxic agent when the temperature reaches 39.5°C [15].

16

Chapter 1

7.7 Treatment antibody-coated liposomes For instance, Fay et al designed polylactic glycolic acid (PLGA) nanoparticles coated with conatumumab DR5-NP (death receptor 5–specific antibodies). The antibodies can target the colorectal HCT116 cancer cell model because cancer cells exhibiting DR5 receptors can be targeted by DR5-NP antibodies [7]. Once DR5 receptors become stimulated, they induce apoptosis, leading to the death of tumour cells . The interactions between antibodies and DR5-expressing tumour cells activate the caspase 8 enzyme, which enhances the cytotoxic activity of camptothecin drugs. Schmid et al. demonstrated that conatumumab antibodies stay on the nanoparticle surface for 96 hours while activating caspase 8 [27]. The clearance of antibodyliposome conjugates occurs more readily than Fab (antigen-binding fragment)?-liposome conjugates due to the inadequate Fc region of an antibody [15].

8. Lipid nanoparticle Lipid nanoparticles are assembled using synthetic lipids as a matrix or drug reservoir for cytotoxic drugs [20]. Stearic acid, lecithin and triglycerides are examples of the lipid materials used in this process [7]. Various routes of administration are available for nanodrug delivery systems that use lipid nanoparticles [20]. Solid lipid nanoparticles (SLN) and nanostructured lipid carriers (NLC) are two types of common lipid nanoparticles with a solid matrix [7–28]. The particle size of lipid nanoparticles ranges from 50 to 1000 nm. The advantages of lipid-based nanoparticles include good compatibility with biological substances and good stability. Moreover, the drug released from its matrix can be controlled to avoid leakage and degradation [20]. Another benefit of using lipid nanoparticles in vivo is their minimal toxicity, making them a suitable drug carrier. NLC is often referred to as the second generation of SLN . Similar to nanoemulsions and liposomes, SLNs are made up of lipids and fatty acids with good biocompatibility and non-toxicity. However, SLN has a solid core instead of a liquid core in nanoemulsions. In SLN, the motion of drug particles in the solid core is restricted, further enhancing the controlled release property of loaded drug particles [28]. Additionally, SLN has a solid matrix wherein the drug is distributed between fatty acids of glycerides that protect the drug against chemical degradation. The SLN’s stability can be improved by adding a coat of surfactant. NLCs comprise many lipid molecules in liquid and solid states. Although liquid lipid is present, the matrix of NLC is solid at room

Current Nanodrug Delivery Systems Used for Colorectal Cancer

17

temperature. The solid matrix of NLCs can restrict the mobility of drugs and prevent the formation of coalescence compared to nanoemulsions. The advantages of NLCs are good biocompatibility, biodegradability, controlledrelease properties and the absence of organic solvent in its formulation [28].

8.1 Preparation lipid nanoparticle A high-pressure homogenisation (100–2000 bar) is used for the large-scale production of SLN and NLC. The high-pressure application causes the fluid to levitate above a speed beyond 1000 km/h, resulting in the breakdown of particles to submicron size. The strong turbulence resulting from the mechanical shear coupled with a strong cavitation force and low pressure across the homogeniser valves are the essential steps in forming nanoparticles. The high-pressure homogenisation can be performed at a high temperature (hot homogenisation or low temperature (cold homogenisation)). The drug is added to the melted lipid at 5–10°C higher than its melting point [28].

8.2 Hot homogenisation procedure The procedure is conducted at temperatures higher than the melting point of the lipid used. Lipids and drugs can melt at this temperature and mix with a liquid surfactant. A high mechanical shear device can be used for preemulsion. Next, a piston-gap homogeniser or a jet-stream homogeniser transforms the pre-emulsion into a hot colloidal emulsion. Crystals of hot colloidal emulsion droplets (nanoemulsion) are formed at room temperature to produce SLNs or NLCs. The droplets should be a few micrometres in size. The low viscosity of the droplets’ inner region at higher temperatures gives rise to smaller particle sizes. About 3–5 homogenisation cycles at 500–1,500 bar are enough to complete the process. However, repeated cycles often produce a greater particle size because of a high particle kinetic energy [28].

8.3 Cold homogenisation procedure Issues associated with the hot homogenisation procedure have led to the development of the cold homogenisation procedure. Examples of these challenges are drug degradation, drug distribution into the liquid phase during the homogenisation process and complications that arise from the crystallisation step of the nanoemulsion and result in alterations and/or supercooled melts. The solid lipid is heated first, followed by the addition

18

Chapter 1

of drug molecules into the matrix of melted lipids. The drug-loaded lipid is converted into a solid form using dry ice or liquid nitrogen. The cooling process is done rapidly to stimulate a uniform distribution of the drug within the lipid. The solid is ground to form microparticles via milling process. Next, the microparticles are incorporated into a cold solution containing surfactant(s) with a high-pressure homogenisation to produce SLNs. In hot homogenisation, solid lipids are homogenised in cold homogenisation instead of homogenising molten lipid [28]. Camptothecin-based drugs, specifically irinotecan (Camptosar) and topotecan (hycamptin), are frequently used with 5-FU [3]. Irinotecan drugs loaded into lipid-based nanoparticles with a size between 100 and 375 nm have been produced using the SN-38 irinotecan analogue [3]. Irinotecan binds to topoisomerase-1 DNA complex reversibly to restrict the formation of double-stranded DNA after a single-strand break. Thus, the DNA becomes damaged, leading to cell apoptosis [29]. A study on mice injected with irinotecan demonstrated an increase in the survival length of mice (65 days) compared to mice treated with encapsulated irinotecan (48 days). However, some challenges still exist in drug delivery due to the hydrophobicity of camptothecin-based drugs [3]. A few limitations of SLNs are low drugloading capacity, risk of dose dumping after polymorphic shift during storage and high amount of water of the dispersions (e.g., 70–99.9%). Additionally, some methods can increase the oral bioavailability of drugs. These include increasing the solubility, preventing the formation of precipitates in the intestine, improving membrane permeability, inhibiting drug removal by protein efflux transporters, lowering cytochrome (CYP) liver enzymes level, and increasing chylomicron production and lymphatic drainage [28].

9. Polymeric nanoparticle A polymeric nanoparticle is described as a spherical particle with a hydrophobic core and a hydrophilic shell. The assembly of amphiphilic block copolymers forms the hydrophilic envelope via the aqueous or microencapsulation approach [4]. Polymer nanoparticles is a term used to represent nanosized spheres and capsules [19]. Polymeric nanoparticles can be formulated into matrix-based nanoparticles (i.e., nanospheres) or nanocapsules, where therapeutic agents are entrapped, adsorbed or encapsulated [15]. Nanospheres are matrix particles wherein the drug molecules are adsorbed to the surface sphere or entrapped within the spherical particle. Nanocapsules act as a reservoir wherein drug molecules

Current Nanodrug Delivery Systems Used for Colorectal Cancer

19

become entrapped in a liquid phase comprising either oil or water. QA solid material envelops the liquid phase [19].

Figure 1-2: Schematic of nanocapsule containing water and oil Created with BioRender.com

There are two types of polymer nanoparticles: natural and synthetic polymers. Gelatin, albumin, chitosan and alginate are some examples of natural polymers. Polylactic acid, PLGA, polyhydroxyalkanoate and polymethyl methacrylate are some examples of synthetically produced polymers [15]. Flexibility is present in manipulating specific properties or the chemical composition of synthetic polymer for a desired biological application [18].

9.1 Natural polymers Natural polymers or biopolymers are proteins with a high molecular weight. A protein comprises building blocks called amino acids bonded together by peptide linkages. Collagen, gelatin and albumin are examples of natural polymers. Collagen plays a larger role in the composition of bone, cartilage and skin. Moreover, incomplete hydrolysis of collagen produces gelatin, a highly soluble protein with inadequate mechanical properties. Collagen is highly biocompatible, degradable and non-teratogenic in the body. Apart from the strengths of using natural polymers, they also present various limitations, such as antigenicity, an increased likelihood of viral infection, and a non-homogenous property between product batches [18].

20

Chapter 1

9.2 Synthetic polymers The synthetic polymer’s repeatability enables the production of polymeric drug delivery systems with homogenous properties. Synthetic polymers can be modified to produce polymers with specific interfacial, biological and mechanical properties. However, because many synthetic polymers cannot be broken down in the body, they are eliminated from the body by renal clearance. Hence, the molecular weight of synthetic polymers should be within the threshold of renal clearance. Acrylic polymers, such as polymethyl methacrylate and Polyhydroxyethylmethacrylate, are some examples of non-biodegradable polymers. Ester, ortho-ester, amide, urea or urethane are some examples of components found in a biodegradable polymer. Biodegradable polymers produce normal metabolites of the body that can be eliminated from the body. It is possible to mix the biodegradable and non-toxic properties of a natural polymer and the mechanical properties of a synthetic polymer during product formulation to obtain the desired properties [18]. The presence of functional groups on the surface of polymer nanoparticles allows modification with ligands to increase the nanoparticles’ specificity [15]. The polymeric-based drug delivery system is divided into four categories: diffusion-controlled system, chemically controlled system, solvent-activated system, osmotically controlled system, and magnetically controlled system. The diffusion-controlled system can be distinguished by the release of drugs via the diffusion process. It comprises two sub-systems: reservoir and matrix systems. The reservoir system consists of a polymeric membrane that coats the drug’s inner region, whereas the matrix system is based on a polymeric matrix where the drug is dispersed uniformly. Nevertheless, the poor polymeric membrane resistance of the reservoir system can lead to the sudden destruction of the system, which will result in dose dumping [18]. A chemically controlled system includes a polymer-drug conjugate, where drug molecules are connected to a polymer by a molecule. The linkage between the polymer and drug is cleaved by two processes: hydrolysis or by an enzymatic activity. Chain cleavage will result in either the biodegradation or bioerosion of the polymer-drug conjugate. Biodegradation refers to when the polymer’s weight is reduced after chain cleavage. Bioerosion is defined as reducing the polymer-drug conjugate system due to the erosion of the polymer surface. Polymeric matrix disruption is responsible for the drug’s release from the system. Hence, the control of the drug release depends on the matrix's degradation rate [18]. Polymer-drug conjugates should only

Current Nanodrug Delivery Systems Used for Colorectal Cancer

21

produce non-teratogenic and biocompatible metabolites. Ideally, polymers should be degradable in the body. Stimuli-responsive nanoparticles are produced to counteract barriers in the microenvironment of the tumour cells and enhance chemotherapeutic efficacy. Many polymer-drug conjugates have linkages between the drug molecule and polymer that control the drug release. Drugs will only be released at the tumour site upon exposure to physiological stimuli (e.g., light, heat or pH) [22]. For instance, in the pHsensitive polymeric nanoparticle system, nanoparticle uptake into cells is controlled by stimuli because the pH-sensitive nanoparticles are responsive in acidic and alkaline mediums [15]. The diffusion of water controls the solvent-activated system into the hydrophilic polymer chain, which does not dissolve the polymer[18]. Water entry into the polymeric system can cause the system to swell. The increased osmotic pressure inside the system pushes the drug into the external environment. Hence, drug release control depends on the amount of water in the matrix that determines the osmotic pressure [18]. Osmotically controlled systems use a device that comprises a semipermeable membrane across which a solvent flows to a chamber containing drugs. The increased osmotic pressure inside the chamber pushes the drug through the device’s orifice [18]. A magnetically controlled system involves the combination of a polymer with magnetic particles controlled by an externally applied magnetic field . The movement of drug particles depends on the final force from the combination of the magnetic and haemodynamic forces of the systemic circulation. However, the external magnetic force should be greater than the haemodynamic force to affect the motion of drug particles. Widely used magnetic nanoparticles include nickel, cobalt and iron, possessing a high magnetic force at room temperature [18].

22

Chapter 1

Figure 1-3: Schematic diagram of drug release that follows the swelling of a hydrophilic polymeric matrix [18] Created with BioRender.com

Current Nanodrug Delivery Systems Used for Colorectal Cancer

Figure 1-4: Schematic diagram of drug particles loaded into a magnetic particle with specific antibodies on the surface of the particle used in cancer therapy Created with BioRender.com

23

Chapter 1

24

9.3 Factors to be considered in polymer-drug conjugate preparation 9.3.1 Molecular weight The molecular weight of polymer impacts the bioavailability of hydrophilic polymers. Polymers with higher molecular weight will increase half-life and slower polymer clearance from the body. For example, polyHPMA-Doxo polymer-drug conjugate, with a molecular weight of 1230 kDa, has a longer half-life by 29 times. Its clearance rate from cancer cells was slower than single-agent Doxo by 25 times [22]. 9.3.2 Architecture The polymer architecture plays a pivotal role in the pharmacokinetic activity of the polymer-drug conjugate. Notably, the polymer architecture influences the renal clearance of polymers. Polymers with a high molecular weight have lower flexibility and more chains at the ends of polymers, resulting in reduced polymer removal and increased retention effect of the polymer in the bloodstream [22]. 9.3.3 Composition of block copolymer The ratio of copolymer composition influences the final spherical shape of nanoparticles. For instance, star micelles will be formed when the molecular weight of the hydrophilic block is greater than the hydrophobic block. The size of nanoparticles can also be affected by the concentration of copolymers in a solvent. Polymers that comprise zwitterions (e.g., cations and anions) have high water solubility to increase the circulation time of nanoparticles [22]. 9.3.4 Preparation of polymeric nanoparticles (PNP) Polymeric nanoparticles (PNP) can be produced by two procedures involving pre-formed polymers and monomers, respectively. The first procedure is the synthesis of nanoparticles from pre-formed polymers by dispersion methods, such as evaporation of solvent, salting-out, dialysis and supercritical fluid technology. The second procedure requires the polymerisation of monomers by various techniques, such as microemulsion, mini emulsion, additives-free emulsion and interfacial polymerisation methods. This chapter will discuss only two methods from each category. The PNP production method depends on the type of polymetric system, application site and nanoparticle size. Nanoparticles should not contain

Current Nanodrug Delivery Systems Used for Colorectal Cancer

25

additives (e.g., surfactants) or organic solvents, particularly nanoparticles developed for clinical use. Hence, RESS (rapid expansion of a supercritical solution) or RESOLV (rapid expansion of a supercritical solution into a liquid solvent) can be conducted to produce additive-free PNP [19]. Under a chemically controlled drug delivery system, the polymer-drug conjugate with paclitaxel demonstrates enhanced drug compatibility with other components in the blood. It prolongs the retention effect of the drug at the desired site than conventional paclitaxel administration [15]. Paclitaxel is a natural taxane substance that inhibits the synthesis of DNA, RNA and protein by interfering in the depolymerisation process of microtubules [30]. Nanoparticles comprising natural polymers, such as albumin-based nanoparticles, demonstrate high biocompatibility and accumulation of platinum drugs (e.g., paclitaxel) at the desired site of action. Thus, only a small dose is required to produce the chemotherapeutic effect, reducing the toxicity of paclitaxel to healthy tissues. Natural polymer reduces the toxicity of paclitaxel to healthy tissues by modifying the drug’s pharmacokinetics [15]. Similarly, paclitaxel-loaded nanoparticles comprising PLGA-PEG polymers can be produced by precipitation with CEA on the surface of nanoparticles. Nanoparticles with this formulation can target colorectal cancer cells. The nanoscale particles increase the efficacy of drugs in site-directed colon drug delivery because they can easily penetrate the inflammatory sites across the colon wall [31]. Cyclodextrin is polymeric amphiphiles with hydrophilic and hydrophobic phases that assist in micelle formation to compartmentalise various drugs. A conjugate of cyclodextrin-camptothecin (IT-101) demonstrated the modified release of an anticancer agent that significantly enhanced antitumour activity against LS174T and HT29 colorectal cancer cell line [15]. The use of hyaluronic acid as a ligand on the surface of nanoparticles has displayed an improvement in cellular uptake. Because an abnormal amount of folate receptors are found in colon cancer cells, using folate acid as a ligand on nanoparticles can also enhance cellular uptake [12]. Magnetic nanoparticles in Nanotherm® therapy (MagForce Nanotechnologies) are directly administered to tumour cells following an application of alternating currents. The nanoparticles have iron oxide cores coated with amino silane. Alternation in the magnetic field increases the nanoparticles’ polarity, leading to heat generation (i.e., intratumoral heat) at the tumour site [15]. The outcome is the death of tumour cells. Magnetic nanoparticles produced from magnetotactic bacteria exert an antineoplastic action with greater cell uptake on the human colon cancer cell line, HT-29 [31]. Moreover, cancer treatment with non-aqueous therapies is beneficial in the leaky vasculature

26

Chapter 1

due to the increased permeability of chemotherapeutics to tumour cells. Defective blood vessels can provide sufficient blood to cancerous tissue cells, and poor drainage of lymphatic fluid can increase the retention effect of cancer therapeutic agents [3]. Polymeric drug delivery systems are beneficial in lowering side effects and treatment costs while increasing patient’s compliance with their therapy [18] 9.3.5 Solvent evaporation In this procedure, a nanoparticle suspension is formed by evaporating a volatile solvent containing a polymer. The suspension may be formed from single-emulsions such as, water-in-oil (w/o) or double emulsions, such as (water-in-oil)-in-water [(w/o)/w]. For instance, a single emulsion can produce poly(ethylene oxide)-modified poly(ȕ-amino ester) nanoparticles between the size range of 100 and 150 nm. Before the evaporation process of the solvent, high-speed homogenisation or ultrasonication techniques are utilised. Nanoparticles are then collected by the ultracentrifugation method and rinsed with distilled water to eliminate any traces of additives. Finally, nanoparticles undergo a freeze-drying method known as lyophilisation from the final product. Despite the simplicity of the procedure, the procedure’s limitations include a long production time and the risk of agglutination of nanodroplets during solvent evaporation. Nanodroplets agglutination can modify nanoparticles' final size and shape [19]. 9.3.6 Salting-out Similar to the solvent evaporation technique, the emulsion is formed with a water-miscible polymer solvent (e.g., acetone). A high salt or sucrose sugar concentration is dissolved in the aqueous phase of the emulsion to give a strong salting-out effect. The emulsion is diluted with a large volume of water, which results in the polymer precipitation in the emulsion droplets to reverse the salting-out effect [19]. 9.3.7 Microemulsion polymerisation The procedure is a new approach for preparing polymeric nanoparticles. In microemulsion polymerisation, the particle size and the number of chains in one particle are relatively smaller when compared to miniemulsion polymerisation. The microemulsion polymerisation process is affected by different types and concentrations of surfactant, monomer, initiator and temperature [19].

Current Nanodrug Delivery Systems Used for Colorectal Cancer

27

9.3.8 Miniemulsion polymerisation The procedure requires water, a mixture of monomers, a co-stabilising agent, a surfactant and an initiator. Co-stabilisers and initiators can influence the nanoparticles' chemistry and their production. Critically stabilised miniemulsion requires a high mechanical shear device to achieve a steady state and a high interfacial tension. Hydrophobic SMA or DMA is used to stabilise the mini emulsion during styrene polymerisation to minimise the undesired changes in PNP properties due to the presence of surfactant in the polymer latex [19].

10. Conclusion Lipid nanoparticles are assembled using synthetic lipids as a matrix or drug reservoir for cytotoxic drugs. SLN and NLC are two types of common lipid nanoparticles with a solid matrix. The advantages of lipid nanoparticles are minimal toxicity, good biocompatibility and controlled release properties. A liposome-based nanodrug delivery system comprises a phospholipid bilayer that envelops an inner aqueous phase of vesicles. Liposomal doxorubicin (Doxil®) is the first generation of FDA-approved drugs for colorectal cancer. Advantages of lipid-based nanoparticles are good biocompatibility, good stability and minimal toxicity. It is also believed that the advanced therapeutic formulation strategies with nanoparticles can promise an individualised treatment for patients. While nanoparticles have numerous advantages in enhancing drug delivery further, comprehensive research is required in its clinical use, especially in chemotherapeutic. Colorectal cancer is defined by the growth of a malignant tumour in the mucosal lining of the colon or rectum. The possibility of cure and high survival rate of patients depend on their stage of colorectal cancer. Adjuvant therapy is normally administered to patients diagnosed with colorectal cancer from Stage III to Stage IV. Fluorouracil treatment (5-FU) has been the first-line treatment of colorectal cancer to control tumour growth and reduce metastases before surgical intervention immediately. However, efficacious drug delivery to the desired cancer tissue without affecting the healthy tissues remains to be a significant challenge. Nanomedicine in colon cancer therapy is an emerging field that presents various approaches in drug delivery to achieve optimal chemotherapeutic efficacy. A therapeutic approach using nanodrug delivery systems in cancer therapy is a promising strategy because of their modified and site-specific drug release properties. For instance, the currently available drug delivery systems are based on polymer nanoparticles, lipid nanoparticles or liposomes. A polymeric

28

Chapter 1

nanoparticle is described as a spherical particle with a hydrophobic core and a hydrophilic shell. There are two types of polymer nanoparticles: natural polymers and synthetic polymers. Polymeric drug delivery systems are beneficial in lowering side effects and treatment costs while increasing patients’ compliance with their therapy.

References 1.

2.

3. 4.

5. 6. 7. 8.

9. 10.

V. Jain, S. Jain, & S. C. Mahajan, Nanomedicines based drug delivery systems for anticancer targeting and treatment. Current Drug Delivery, (2015). https://doi.org/10.2174/1567201811666140822112516. M. H. Mehran Alavi, Passive and active targeting in cancer therapy by liposomes and lipid nanoparticles. Drug Metabolism and Personalized Therapy, (2019) 1–8. https://doi.org/10.1515/dmpt2018-0032. L. Brannon-peppas & J. O. Blanchette, Nanoparticle and targeted systems for cancer therapy. Advanced Drug Delivery Reviews, 64 (2012) 206–212. https://doi.org/10.1016/j.addr.2012.09.033. B. A. Cisterna, N. Kamaly, W. Il Choi, A. Tavakkoli, O. C. Farokhzad, & C. Vilos, Targeted nanoparticles for colorectal cancer. Nanomedicine, 11 (2016) 2443–2456. https://doi.org/10.2217/nnm2016-0194. S. Stintzing, Management of colorectal cancer. F1000Prime Reports, 6 (2014) 1–12. https://doi.org/10.12703/P6-108. D. Wisher, Martindale: The complete drug reference. 37th ed. Journal of the Medical Library Associationࣟ: JMLA, 100 (2012) 75– 76. https://doi.org/10.3163/1536-5050.100.1.018. A. Gulbake, A. Jain, A. Jain, A. Jain, & S. K. Jain, Insight to drug delivery aspects for colorectal cancer. World Journal of Gastroenterology, (2016). https://doi.org/10.3748/wjg.v22.i2.582. F. Teimoorian, M. Ranaei, K. Hajian Tilaki, J. Shokri Shirvani, & Z. Vosough, Association of helicobacter pylori infection with colon cancer and adenomatous polyps. Iranian Journal of Pathology, 13 (2018) 325–332. A. Ferrandez, S. Prescott, & R. Burt, COX-2 and colorectal cancer. Current Pharmaceutical Design, 9 (2003) 2229–2251. https://doi.org/10.2174/1381612033454036. W. Wang, L. Wu, W. Lu, W. Chen, W. Yan, C. Qi, S. Xuan, & A. Shang, Lipopolysaccharides increase the risk of colorectal cancer recurrence and metastasis due to the induction of neutrophil

Current Nanodrug Delivery Systems Used for Colorectal Cancer

11. 12. 13.

14.

15.

16.

17. 18.

19.

20.

29

extracellular traps after curative resection. Journal of Cancer Research and Clinical Oncology, 147 (2021) 2609–2619. https://doi.org/10.1007/s00432-021-03682-8. H. K. Chew, Medicine cabinet: Adjuvant therapy for breast cancer: Who should get what? Western Journal of Medicine, 174 (2001) 284–287. https://doi.org/10.1136/ewjm.174.4.284. X. You, Y. Kang, G. Hollett, X. Chen, W. Zhao, Z. Gu, & J. Wu, Overview and perspectives. Journal of Materials Chemistry, 4 (2016) 7779–7792. https://doi.org/10.1039/c6tb01925k. S. Tran, P. J. Degiovanni, B. Piel, & P. Rai, Cancer nanomedicine: A review of recent success in drug delivery. Clinical and Translational Medicine, (2017). https://doi.org/10.1186/s40169-017-0175-0. G. Pillai, Nanomedicines for cancer therapy: An update of FDA approved and those under various stages of development. SOJ Pharmacy & Pharmaceutical Sciences, (2014). https://doi.org/10.15226/2374-6866/1/2/00109. A. Babu, A. K. Templeton, A. Munshi, & R. Ramesh, Nanodrug delivery systems: A promising technology for detection, diagnosis, and treatment of cancer. AAPS PharmSciTech, (2014). https://doi.org/10.1208/s12249-014-0089-8. Julia Sánchez-Gundín, Ana María Fernández-Carballido, Lidia Martínez-Valdivieso, Dolores Barreda-Hernández, Ana Isabel Torres-Suárez, New trends in the therapeutic approach to metastatic colorectal cancer. International Journal of Medical Sciences, 15 (2018) 659–665. https://doi.org/10.7150/ijms.24453. National Cancer Institute. NIH public access, (n.d.). https://www.cancer.gov/publications/dictionaries/cancerterms/def/overall-survival-rate (accessed December 19, 2021). J. F. Coelho, P. C. Ferreira, P. Alves, R. Cordeiro, A. C. Fonseca, J. R. Góis, & M. H. Gil, Drug delivery systems: Advanced technologies potentially applicable in personalized treatments. EPMA Journal, 1 (2010) 164–209. https://doi.org/10.1007/s13167-010-0001-x. J. P. Rao & K. E. Geckeler, Polymer nanoparticles: Preparation techniques and size-control parameters. Progress in Polymer Science (Oxford), 36 (2011) 887–913. https://doi.org/10.1016/j.progpolymsci.2011.01.001. Y. Huang, S. P. C. Cole, T. Cai, & Y. Cai, Applications of nanoparticle drug delivery systems for the reversal of multidrug resistance in cancer. Oncology Letters, (2016).

30

21.

22. 23. 24. 25.

26. 27.

28.

29.

30.

Chapter 1

https://doi.org/10.3892/ol.2016.4596. P. Aggarwal, J. B. Hall, C. B. McLeland, M. A. Dobrovolskaia, & S. E. McNeil, Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efficacy. Advanced Drug Delivery Reviews, 61 (2009) 428–437. https://doi.org/10.1016/j.addr.2009.03.009. Q. Feng & R. Tong, Anticancer nanoparticulate polymer-drug conjugate. Bioengineering & Translational Medicine, 1 (2016) 277– 296. https://doi.org/10.1002/btm2.10033. C. Rocha, M. Silva, A. Quinet, J. Cabral-Neto, & C. Menck, DNA repair pathways and cisplatin resistance: An intimate relationship. Clinics, 73 (2018). https://doi.org/10.6061/clinics/2018/e478s. G. N. C. Hughes, Cancer nanotechnology, preparation and characterization of doxorubicin liposomes (2010). https://doi.org/10.1007/978-1-60761-609-2. S. Banerjee, K. Sen, T. K. Pal, & S. K. Guha, Poly(styrene-co-maleic acid)-based pH-sensitive liposomes mediate cytosolic delivery of drugs for enhanced cancer chemotherapy. International Journal of Pharmaceutics, 436 (2012) 786–797. https://doi.org/10.1016/j.ijpharm.2012.07.059. N. Zhang, Y. Yin, S.-J. Xu, & W.-S. Chen, 5-Fluorouracil: Mechanisms of resistance and reversal strategies. Molecules, 13 (2008) 1551–1569. https://doi.org/10.3390/molecules13081551. D. Schmid, F. Fay, D. M. Small, J. Jaworski, J. S. Riley, D. Tegazzini, C. Fenning, D. S. Jones, P. G. Johnston, D. B. Longley, & C. J. Scott, Efficient drug delivery and induction of apoptosis in colorectal tumours using a death receptor 5-targeted nanomedicine. Molecular Therapy, 22 (2014) 2083–2092. https://doi.org/10.1038/mt.2014.137. P. Ganesan & D. Narayanasamy, Lipid nanoparticles: Different preparation techniques, characterization, hurdles, and strategies for the production of solid lipid nanoparticles and nanostructured lipid carriers for oral drug delivery. Sustainable Chemistry and Pharmacy, 6 (2017) 37–56. https://doi.org/10.1016/j.scp.2017.07.002. M. Kciuk, B. Marciniak, & R. Kontek, Irinotecan—Still an important player in cancer chemotherapy: A comprehensive overview. International Journal of Molecular Sciences, 21 (2020) 4919. https://doi.org/10.3390/ijms21144919. B. A. Weaver, How taxol/paclitaxel kills cancer cells. Molecular Biology of the Cell, 25 (2014) 2677–2681. https://doi.org/10.1091/mbc.e14-04-0916.

Current Nanodrug Delivery Systems Used for Colorectal Cancer

31.

31

H. Nugrahani, Iskandarsyah, & Harmita, Stability study of azelaic acid proethosomes with lyoprotectant as stabilizer. Journal of Advanced Pharmaceutical Technology & Research, 9 (2018) 61. https://doi.org/10.4103/japtr.JAPTR_252_18.

CHAPTER 2 ADVANCES IN NANOPARTICLES FOR LUNG CANCER THERAPY RAJAN RAJABALAYA*1 AND STACY DAVID2 1

PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2 Department of Biology, Indiana University-Purdue University Indianapolis, IN 46202, USA. *Corresponding author: Dr Rajan Rajabalaya Senior Assistant Professor PAPRSB Institute of Health Sciences Universiti Brunei Darussalam Jalan Tungku Link BE1410 Bandar Seri Begawan Brunei Darussalam Email: [email protected] Phone: + 6732460922

Abstract Lung cancer is the most common type of cancer worldwide. It has been identified as a top life-threatening disease. The current treatments for lung cancer are surgery, chemotherapy and radiotherapy. However, some challenges, such as multidrug resistance (MDR), non-site-specific, high toxicity, and side effects, arise from these treatments. Recent years have witnessed advancements and developments of nanotechnology in medical applications (i.e., diagnosis, sensors, imaging tests and drug delivery). Because of these developments, nanoparticles have been used to deliver chemotherapeutics drugs. Different types of nanoparticles such as liposome, micelle, dendrimer, polymer-based and quantum dots are available. The use of nanoparticles in drug delivery has been increasing because of the stability, site-specific targeting and improved pharmacological effects of the drugs to

Advances in Nanoparticles for Lung Cancer Therapy

33

the desired location. Moreover, the properties exhibited by the nanoparticles can be modified, providing more useful applications for the nanoparticles. This chapter discusses the potential benefits of using nanoparticles in lung cancer therapy.

Abbreviations ADC AF AJCC AUC CdSe CMC CT Scan CUR DOX EGFR LCC LUV MDR MLV MRI NSCLC PAMAM PEG PET PLA RSV RSV-PEG-Lipo SCC SCLC SUV WHO ZnS

Adenocarcinoma Aminoflavone American Joint Committee on Cancer Area under the curve Cadmium selenide Critical micelle concentration Computed tomography scan Curcumin Doxorubicin Epidermal growth factor receptor Large cell undifferentiated carcinoma Large unilamellar vesicles Multidrug resistant Multilamellar vesicle Magnetic resonance imaging Non-small cell lung cancer polyamidoamine Polyethylene glycol Positron emission tomography Polylactic acid Trans-resveratrol RSV-loaded PEGylated liposome Squamous cell carcinoma Small cell lung cancer Small unilamellar vesicle World Health Organisation Zinc selenide

34

Chapter 2

1. Introduction According to the World Health Organisation (WHO), cancer was responsible for 9.6 million deaths in 2018, with lung cancer being the top life-threatening cancer with 1.76 million deaths [1]. Lung cancer is characterised by aberrant cell growth in the lung tissues [2]. Cancer cells that originate from the lungs can also metastasise to other parts or organs in the body (such as the brain and lymph nodes); this process is known as metastases. There are two types of lung cancer: small cell lung cancer (SCLC) and non-small cell lung cancer (NSCLC). A total of 15% of lung cancers occur due to SCLC; nevertheless, it can metastasise more rapidly than NSCLC. The most common type of lung cancer is NSCLC. It can be subdivided into three types: adenocarcinoma (ADC), squamous cell carcinoma (SCC) and large cell undifferentiated carcinoma (LCC). ADC is the most common lung cancer, and it starts in the epithelial cells that secrete substances such as mucus. SCC accounts for 25–30% of lung cancer cases. It usually begins in the squamous cells, which are flat cells that line the inside of the airways in the lung. Meanwhile, large cell carcinoma usually appears in any part of the lung (i.e., nearby lymph nodes) [3]. Tobacco smoking is the main aetiological factor of lung cancer [4]. The duration and number of cigarettes taken per day are considered the leading cause of lung cancer [5]. Involuntary smoking caused by prolonged exposure to tobacco smokes from the environment, such as homes and workplaces, can also cause lung cancer [6]. Other possible determinants of lung cancer are genetic factors, diet intake and exposure to toxic substances [7]. A few examples of toxic substances are asbestos and radon gas. An initial evaluation requires the examination of medical history, physical examination and blood tests from patients suspected of having lung cancer [8]. Imaging tests and biopsies will also be performed to diagnose cancer. Imaging tests such as chest x-ray, computed tomography scan (CT scan), magnetic resonance imaging (MRI) and positron emission tomography (PET) scan are also used for further confirmation. Bronchoscopy is an example of a biopsy. The first imaging test to evaluate lung cancer is a chest x-ray, wherein the appearance of lung tumours can be observed as a whitegrey mass [7]. However, negative results of lung cancer in the chest x-ray do not conclude the absence of lung cancer. Hence, more investigations are needed to confirm the diagnosis. Imaging tests aim to observe the suspected area regarding its size and shape, detect how far the cancer might have spread, check the cancer drug's effectiveness, and look for possible signs of recurrent lung cancer after post-treatment [8]. Thus, the procedure selection

Advances in Nanoparticles for Lung Cancer Therapy

35

must be appropriate and depend on the type, site and size of the tumour, the comorbidities, and the accessibility of metastases [9]. Following a cancer diagnosis, cancer staging and types of cancer should be established. Cancer staging for NSCLC is assessed through the degree of cancer from its original site. It is based on American Joint Committee on Cancer (AJCC) TNM system 8th Edition [7]. I In the TNM system, T stands for tumour (the size of the tumour), N refers to nodes (to check the spite of the cancer cells in lymph nodes), and M stands for the cancer cells' ability to metastasise [10]. Cancer staging of NSCLC ranges from Stage 0 to Stage IV. As for SCLC, there are only two stages of cancer that are limited and extensive [9]. The choice to treat lung cancer varies among patients and is determined by the type and stage of lung cancer and the patient’s physical conditions [7– 11]. Various treatments, such as surgery, radiotherapy, chemotherapy, immunotherapy and targeted therapy, are available to treat lung cancer. Other treatments include radiofrequency ablation, cryotherapy and photodynamic therapy. The treatment options available can be a local directed treatment or systemic treatment, depending on the patient’s condition. Surgery comprises lobectomy, pneumonectomy and wedge resection or segmentectomy. Radiotherapy comprises conventional external beam radiotherapy, stereotactic radiotherapy and internal radiotherapy. Chemotherapy treatments are administered in cycles where the number of cycles depends on the type and grade of lung cancer. Moreover, the treatment of lung cancer may also require combinatorial therapy. Generally, the treatment of all SCLC involves chemotherapy, radiotherapy or chemoradiotherapy [12]. However, surgery can also be performed as the initial treatment for SCLC and NSCLC. Surgery is done to remove the primary tumour with no evidence of spreading to lymph nodes [12–13]. For Stages 2 and 3 of NSCLC, chemotherapy or chemoradiotherapy, postsurgery or neoadjuvant chemotherapy before surgery is administered [11– 13]. Chemoradiotherapy is the only treatment available for Stage 4 NSCLC [11–12]. Despite these advances, there are challenges in current lung cancer treatments. First, the distribution of chemotherapy drugs in the body is nonsite-specific and has a higher chance of multidrug resistance (MDR) [14]. This results in treatment failure to inhibit tumour growth, metastasis and recurrence. Another significant limitation of current lung cancer treatment

36

Chapter 2

Figure 2-1: Lung cancer types and lung cancer treatment; small cell lung cancer and non-small cell lung cancer are the two major classifications of lung cancer; both local and systemic treatment options are available for respiratory carcinoma Created with BioRender.com

Advances in Nanoparticles for Lung Cancer Therapy

37

is its dose-limiting side effects. The side effects range from minor side effects (i.e., nausea and vomiting) to severe side effects (i.e., nephrotoxicity, cardiotoxicity and thrombocytopenia). Table 2-1 shows the cytotoxic effects of common lung cancer therapeutics. Thus, an approach to altering the antitumour drug vehicles via nanotechnology must be considered to maintain or improve the effectiveness of cancer treatment. Table 2-1 The cytotoxic effects on common lung cancer therapeutics. Drug

Classification of drug

Cisplatin

Alkylating agent

Paclitaxel

Carboplatin

Gemcitabine

Antimicrotubule agent

Alkylating agent

Antimetabolite

Mechanism of action in lung cancer Formation of DNA adducts Induces apoptosis Formation of DNA crosslink Bind to tubulin Form a stable, nonfunctional microtubule Inhibition of cell cycle at mitotic phase Form interstrand and intrastrand crosslinking of DNA molecules Inhibit DNA synthesis Inhibition of DNA synthesis Induction of apoptosis

Side effects

Reference(s)

Emesis Nephrotoxicity Neurotoxicity

[13]

Anaemia Febrile neutropenia Leukopenia

[14–15]

Thrombocytopenia Anaemia Leukopenia

[16]

Anaemia Neutropenia Thrombocytopenia

[17]

2. Nanotechnology and nanoparticles Nanotechnology can be defined as studying, processing and manipulating materials at a molecular level ranging in nanometres [15]. Nanotechnology mainly deals with nanoparticles having one of its dimensions in the scale range of 1–100 nm. There are three layers in nanoparticles: the surface layer, shell layer and core layer [16]. There are variations in the size and shape of nanotechnology products; hence, they can affect the nanoparticles circulation, target site, degradation and the drug release kinetics [17].

Chapter 2

38

Additionally, the colour also depends on the size and shape of the products [16]. Moreover, the biological effects of nanoparticles can also be affected by their surface properties (i.e., hydrophobicity and surface charge) and the presence or absence of targeting ligands [17]. Nanotechnology has been widely used and applied in modern industries, such as medicine, energy and industrial, food, devices, information technology and fabrics. One of the uses of nanotechnology in medicine is to improve medical imaging for disease diagnosis, providing imaging at the specific site of targeted tissue at resolutions that is not possible by the current technologies. Nanotechnology can also produce targeted drug delivery to the specific site, reducing patients’ side effects or toxicity problems [18]. Therefore, it is vital to understand the benefits of using nanotechnology in medicine. This review discusses how nanotechnology can help deliver the drug with minimal toxicity to treat patients with lung cancer and improve pharmacological effects.

2.1 Types of nanoparticles drug delivery systems for lung cancer There are two types of nanoparticles: organic and inorganic nanoparticles. The uses of every nanoparticle may vary because it can have more than one application (i.e., medical imaging and drug delivery) [17]. Liposome, micelle, dendrimers and polymer-based nanoparticles are some examples of organic nanoparticles. Fullerenes, quantum dots and metallic nanoparticles are some examples of inorganic nanoparticles. 2.1.1 Liposome The liposome is a spherical vesicle that comprises an aqueous cavity surrounded by a lipid bilayer [15]. It can be classified as follows: a multilamellar vesicle (MLV), large unilamellar vesicles (LUV) and small unilamellar vesicle (SUV) (Figure 2-2). The size of different classifications differs; MLV is about 0.5 μm, LUV is 100 nm to 1 μm, and SUV is about 20–100 nm [19]. Liposome nanoparticle has been widely used for drug delivery. For example, they deliver hydrophilic and hydrophobic drugs into a specific site in the body. Apart from drug delivery, liposome nanoparticles can also deliver genes. Hydrophobic drugs will have a higher affinity to the phospholipid bilayer in a conventional liposome. Meanwhile, hydrophilic drugs or genes will be surrounded by an aqueous medium of a liposome [20].

Advances in Nanoparticles for Lung Cancer Therapy

39

Figure 2-2: Classification of liposomes based on size and number of bilayers; multilamellar vesicle, large unilamellar vesicle and small unilamellar vesicle are few types of liposomes Created with BioRender.com

40

Chapter 2

Liposomal drug delivery systems can be either passive or active. Figure 2– 3 presents the examples of the passive conventional liposome. The structure of conventional liposomes exhibits a normal liposome behaviour in drug delivery. The alteration of the conventional liposome structure will be formed for active targeting. One example is the stealth liposome, wherein polyethylene glycol (PEG) is incorporated into the conventional liposome. Additional PEG helps decrease blood protein adsorption, thus increasing the circulation rate. A previous study aimed to prolong the half-life and rapid elimination of Trans-resveratrol (RSV) for treating glioma cells [21]. The results showed that the area under the curve, plasma half-life and mean residence time increased by using RSV-loaded PEGylated liposome (RSVPEG-Lipo). Moreover, a decrease in the volume of distribution was also reported, which benefited the treatment of RSV for glioma cells. Another active targeting liposome is known as ethosome, which assists in skin penetrations by adding more ethanol content. Figure 2-3 shows the four different structures of liposomes, which behave differently because of additional materials to alter the properties. Different methods, such as thin hydration, reverse-phase evaporation, freeze-drying and ethanol injection, are available to prepare liposomes. However, most liposomes are prepared using thin hydration [20]. Membrane extrusion, sonication, homogenisation and/or freeze-thawing are used to adjust the size and size distribution. Additionally, the charge and lamellarity of liposomes produced can also be formulated. Because of the advancement of liposome technology, more types of liposomal drug delivery were formulated in cancer therapy, fungal diseases, analgesics and photodynamic therapy [20]. The increased usage of these nanoparticles for drug delivery systems is attributed to their minor toxicity, biocompatibility and controlled drug release. In 1995, the first drug formulated using liposome drug delivery known as Doxil® was introduced to treat ovarian cancer. Additionally, Doxil® was also used as an alternative treatment for AIDS-related Kaposi’s sarcoma after failure to respond to chemotherapy drugs. The result demonstrated that out of 15 patients who received Doxil® as their treatment, a partial response rate in 11 patients was reported, with disease stabilisation in the remaining patients [23].

Advances in Nanoparticles for Lung Cancer Therapy

41

Figure 2-3: (A) Liposomes: Conventional liposomes are made of phospholipids; (B) PEGylated/stealth liposomes contain a layer of polyethylene glycol (PEG) at the surface of liposomes; (C) targeted liposomes contain a specific targeting ligand to target a cancer site; and (D) multifunctional such as theranostic liposomes, which can be used for diagnosis and treatment of solid tumours [22]. Created with BioRender.com

42

Chapter 2

2.1.2 Micelle Micelles are made up of amphiphiles molecules such as polymers or lipids. The size of the micelle nanoparticle ranges from 5 to 100 nm. Micelle comprises a core and a shell; the core consists of a hydrophobic group. Further, the shell can be adjusted with hydrophobic or hydrophilic domains [24]. PEG is used to produce the hydrophilic shell. Materials such as polylactic acid (PLA), polystyrene and polycaprolactone can produce the hydrophobic shell. Micelles are classified into normal micelle, reverse micelle, and unimolecular micelle [25]. Because of the different structural orientations of micelles, hydrophobic or hydrophilic drugs for delivery can be formulated. First, the individual amphiphilic chains will be dissolved in an aqueous medium to produce micelle nanoparticles [12]. When the surfactant in the water reaches beyond the critical micelle concentration (CMC) and is triggered by stimuli such as temperature or pH change, the amphiphilic block copolymers will self-assemble themselves and surround the aqueous medium to form a micelle [12,24]. Because of the involvement of stimuli sensitivity that forms micelle, these nanoparticles were also used to deliver chemotherapy drugs [26]. Notably, the selected materials used to prepare a micelle can affect the size, shape, stability, and drug retention rate of a micelle. Some advantages of using micelle nanoparticles are decreased toxicity, ease of preparation of the micelle drug delivery systems, and high drug solubility and circulation rate [12]. A study was conducted to investigate the effectiveness of micelle nanoparticles to reduce tumour growth and toxicity in triplet negative breast cancer [27]. The study demonstrated that both amentoflavone (AF) and GE-11, a 12-amino acid peptide targeting epidermal growth factor receptor (EGFR), can increase the cellular uptake and stop tumour growth. However, AF can cause toxicity; therefore, AF-loaded GE-11 conjugated unicellular micelle nanoparticle was used. The study results showed significant tumour growth inhibition in using AF-loaded GE-11 conjugated unicellular micelle nanoparticles compared to free-AF or AF-loaded GE-11. 2.1.3 Dendrimers Dendrimers formed a spherical shape comprising repetitive tree-like branched molecules. The structures of dendrimers contain a core, interior layers of branches and are surrounded by an external surface of terminal functionalised branches [28]. A different generation of the dendrimer will be obtained from the addition of repeating branching units to form a

Advances in Nanoparticles for Lung Cancer Therapy

43

globular structure [29]. Thus, this action leads to variation in the size and shapes of the dendrimers and the length and density of the branches. Additionally, the alteration of the external surface can also form different chemical functional groups. This outcome improves the molecular targeting groups, imaging agents and binding sites to achieve the maximum therapeutic outcomes. Because of the possible adjustment of dendrimers’ structure, it has been used as sensors for drug delivery and as a carrier to deliver genes. The synthesis of dendrimers can be produced by using divergent or convergent approaches [30]. In the divergent synthesis, a new layer forming a new dendrimer’s generation will be added to the final molecules, leading to the expansion of the final molecule from its core. The previous process must be completed before adding a new generation to avoid branch defects. By using divergent synthesis, the fast production of dendrimers can be made, and the addition of a substance to the external surface can be formulated to obtain the desired function of the dendrimer. However, divergent synthesis requires a long period of purification. As for convergent methods, the expansion begins at the terminal chain of the branches, whereby it merges the branching point with the other monomer [29]. After the branches expand, they will join the core to form a dendrimer. Advantages of using convergent synthesis are ease of purification, and it does not require additional dendrimer generation, reducing the possibility of branch defects. Despite the advantages, convergent synthesis may also cause problems such as steric hindrance and low production of dendrimers. Polyamidoamine (PAMAM) is a widely used dendrimer with high branches and numerous functional groups on its surface, which assist in drug delivery via encapsulation or conjugation [31]. An example of PAMAM drug delivery is delivering chemotherapy drugs to the targeted site. Although dendrimers have many advantages, dendrimers also exhibit some consequences that lead to cell toxicity and cause haemolysis. Moreover, dendrimers can possess positively charged polymers, which cause cells to be destabilised, protein to leak out of the membrane and cell lysis [29]. 2.1.4 Polymer-based nanoparticles Polymer-based nanoparticles are colloidal particles that range from 10 nm to 1000 nm. The incorporation of drugs or therapeutic agents into the colloidal particles can be achieved by dissolving the drugs into the colloidal solution. Then, the drugs will be entrapped or encapsulated within the polymer. Additionally, the drugs can also be adsorbed to the surface of the

44

Chapter 2

colloidal particles. Materials used to prepare these nanoparticles can include synthetic (i.e., poly-lactide and polyacrylates) or natural polymers (i.e., chitosan, albumin and alginate). There are two types of polymer-based nanoparticles: nanospheres and nanocapsules. The type of nanoparticles produced will be based on the selected preparation methods. While formulating these nanoparticles, the modification of the materials used may be needed to achieve the required properties of the nanoparticles for the desired application [32]. There are two methods to prepare polymer-based nanoparticles: the dispersion of preformed polymer and the polymerisation of monomers. Notably, the polymers used in a preformed polymer are synthetic polymers. There are three mechanisms of drugs release by using polymeric-based drugs to the desired location. The first mechanism is the nanoparticles' hydration, which causes the nanoparticles to become swollen. Then, the drugs will be released via diffusion. The second mechanism involves polymer degradation due to enzymatic reactions. Then, the drugs will be released from the inner core. The third mechanism is the swelling of nanoparticles that will cause the dissociation of drugs from the surface of the polymer-based nanoparticles. In medicine, these nanoparticles have been applied as diagnostic tools and for drug, protein or DNA delivery to the targeted site in the body. One advantage of using polymer-based nanoparticles for drug delivery is the easy modification of the structure to obtain the required functions [32]. Hence, this action increases drug safety. Despite that advantage, drug delivery by these nanoparticles also improves in its effectiveness compared to the free drug [33]. 2.1.5 Inorganic nanoparticles – Quantum dots Quantum dots are semiconductor particles that consist of a core, shell and ligands. The core is made up of cadmium selenide (CdSe) and a zinc selenide (ZnS) shell. The diameter of quantum dots nanoparticles is less than 10 nm. Some applications of quantum dots nanoparticles include diagnostic tools (i.e., cell labelling and biomolecule tracking), imaging and drug delivery. Because of the ability of quantum dots nanoparticles to produce fluorescence colour, they have been used for a theranostic purpose [34]. The size particle of a drug-loaded nanoparticle is small; thus, the drugs will remain at the targeted site for a longer time, increasing the drugs’ solubility, absorption, distribution and metabolism. Moreover, quantum dots nanoparticles can also control the release of drugs, causing the reduction of doses and side effects [35].

45

Doxorubicin (DOX) + Curcumin (CUR)

Tumourbearing Mice

Phase 1 and 2 Human

Human

Paclitaxel

Micelle Paclitaxel

Subcutaneous, 806 mcg

Phase 3 Human

1 μg/mL or higher (Higher than the DOXresistant A549/Adr cells)

230 mg/m2

Intravenous, 60 mg/m2

Dose

Subject

Drug Liposome Tecemotide (LBLP25)

Not recorded

Similar with free drug

Lower serious adverse events compared to the placebo + saline group Anaemia Leukopenia

Adverse effects

Decreased side effects of DOX

Longer circulation of DOX or CUR Strongly inhibited the growth of tumours

Slightly higher response rate than free drug

Significant antitumour effect to SCC with manageable toxicities

Higher survival rate

Pharmacological effects

[39]

[38]

[37]

[36]

Reference

Table 2-2: Evidence supporting the use of nanoparticles as drug delivery systems for the treatment of lung cancer

Advances in Nanoparticles for Lung Cancer Therapy

Etoposide

Mice

Human (Stage IV NCLC)

Rats

Doxorubicin

Polymer Paclitaxel

Mice

Dendrimer Doxorubicin

46

Intratumorally, 36 mg/kg

Intravenous, 100 mg/m2

Intratracheal instillation / 3 mg/kg

Pharyngeal aspiration or intravenous, 1 mg/kg

Not reported

No severe hypersensitivity reactions No episodes of treatment-related grade 4 adverse effects

Respiratory distress

Low/ no toxicity

Chapter 2

Suppressed tumour growth

Sustained the drug release

Increase survival rate

One-year survival rate: 30%

Disease control rate: 40%

More than 95% reduction of tumour burden when compared to intravenous administration.

Slow clearance of the drug

Reduced metastatic lung tumour

Reduces tumour size and its mass than free drug

[43]

[42]

[41]

[40]

Advances in Nanoparticles for Lung Cancer Therapy

47

3. Conclusion and future perspectives The use of nanoparticles has been shown as an excellent approach in medical applications and is not only limited to drug delivery. It can also improve the disease diagnosis and the results of the imaging tests performed. Because of the properties exhibited by the nanoparticles, more applications of the nanoparticles can be used. Moreover, these nanoparticles have benefited the patients in many ways, such as reducing the side effects toxicity and improving the drug’s pharmacology and patients’ compliance. Additionally, the effectiveness of using nanoparticles, particularly for lung cancer therapy, has been strengthened by supporting evidence and clinical trials performed in vivo and in vitro. However, more ongoing research and clinical trials in vivo are expected to investigate the effectiveness of using nanoparticles for lung cancer therapy.

References 1. 2.

3. 4. 5.

6. 7.

K. Chaitanya Thandra, A. Barsouk, K. Saginala, J. Sukumar Aluru, & A. Barsouk, Epidemiology of lung cancer. :VSyáF]HVQD Onkologia, 25 (2021) 45–52. https://doi.org/10.5114/wo.2021.103829. J. Ferlay, M. Colombet, I. Soerjomataram, D. M. Parkin, M. Piñeros, A. Znaor, & F. Bray, Cancer statistics for the year 2020: An overview. International Journal of Cancer, 149 (2021) 778–789. https://doi.org/10.1002/ijc.33588. C. Zappa & S. A. Mousa, Non-small cell lung cancer: Current treatment and future advances. Translational Lung Cancer Research, 5 (2016) 288–300. https://doi.org/10.21037/tlcr.2016.06.07. C. S. Dela Cruz, L. T. Tanoue, & R. A. Matthay, Lung cancer: Epidemiology, etiology, and prevention. Clinics in Chest Medicine, 32 (2011) 605–644. https://doi.org/10.1016/j.ccm.2011.09.001. J. Malhotra, M. Malvezzi, E. Negri, C. La Vecchia, & P. Boffetta, Risk factors for lung cancer worldwide. European Respiratory Journal, 48 (2016) 889–902. https://doi.org/10.1183/13993003.00359-2016. A. Cassidy & J. K. Field, Environmental and genetic risk factors of lung cancer. Outcome Predict. Cancer (2007), 67–100. https://doi.org/10.1016/B978-044452855-1/50006-4. M. Mustafa, A. J. Azizi, E. IIIzam, A. Nazirah, S. Sharifa, & S. Abbas, Lung cancer: Risk factors, management, and prognosis. IOSR Journal of Dental and Medical Sciences, 15 (2016) 94–101. https://doi.org/10.9790/0853-15100494101.

48

8.

9. 10. 11. 12.

13. 14. 15. 16. 17. 18.

19.

Chapter 2

C. Goebel, C. L. Louden, R. Mckenna, O. Onugha, A. Wachtel, & T. Long, Diagnosis of non-small cell lung cancer for early stage asymptomatic patients. Cancer Genomics - Proteomics, 16 (2019) 229–244. https://doi.org/10.21873/cgp.20128. G. P. Kalemkerian, Staging and imaging of small cell lung cancer. Cancer Imaging, 11 (2011) 253–258. https://doi.org/10.1102/14707330.2011.0036. O. Lababede & M. A. Meziane, The eighth edition of TNM staging of lung cancer: Reference chart and diagrams. The Oncologist, 23 (2018) 844–848. https://doi.org/10.1634/theoncologist.2017-0659. T. Yano, Therapeutic strategy for postoperative recurrence in patients with non-small cell lung cancer. World Journal of Clinical Oncology, 5 (2014) 1048. https://doi.org/10.5306/wjco.v5.i5.1048. M. Amararathna, K. Goralski, D. Hoskin, & H. P. Rupasinghe, Pulmonary nano-drug delivery systems for lung cancer: Current knowledge and prospects. Journal of Lung Health and Diseases, 3 (2019) 11–28. https://doi.org/10.29245/2689-999X/2019/2.1148. D. Landesman-Milo & D. Peer, Altering the immune response with lipid-based nanoparticles. Journal of Controlled Release, 161 (2012) 600–608. https://doi.org/10.1016/j.jconrel.2011.12.034. C.-Y. Zhao, R. Cheng, Z. Yang, & Z.-M. Tian, Nanotechnology for cancer therapy based on chemotherapy. Molecules, 23 (2018) 826. https://doi.org/10.3390/molecules23040826. P.Heeraand & S.Shanmugam, Nanoparticle characterization and application: An overview. International Journal of Current Microbiology and Applied Sciences, 4 (2015) 379–386. I. Khan, K. Saeed, & I. Khan, Nanoparticles: Properties, applications and toxicities. Arabian Journal of Chemistry, 12 (2019) 908–931. https://doi.org/10.1016/j.arabjc.2017.05.011. R. Singh & J. W. Lillard, Nanoparticle-based targeted drug delivery. Experimental and Molecular Pathology, 86 (2009) 215–223. https://doi.org/10.1016/j.yexmp.2008.12.004. Anderson DS, Sydor MJ, Fletcher P and Holian A, Nanotechnology: The risks and benefits for medical diagnosis and treatment. Journal of Nanomedicine & Nanotechnology, 7 (2016). https://doi.org/10.4172/2157-7439.1000e143. Y. Y. Khan & V. Suvarna, Liposomes containing phytochemicals for cancer treatment-An update. International Journal of Current Pharmaceutical Research, 9 (2016) 20. https://doi.org/10.22159/ijcpr.2017v9i1.16629.

Advances in Nanoparticles for Lung Cancer Therapy

20. 21.

22.

23.

24. 25.

26.

27.

28. 29.

49

U. Bulbake, S. Doppalapudi, N. Kommineni, & W. Khan, Liposomal formulations in clinical use: An updated review. Pharmaceutics, 9 (2017) 12. https://doi.org/10.3390/pharmaceutics9020012. M. R. Vijayakumar, R. Kosuru, P. R. Vuddanda, S. K. Singh, & S. Singh, Trans resveratrol loaded DSPE PEG 2000 coated liposomes: An evidence for prolonged systemic circulation and passive brain targeting. Journal of Drug Delivery Science and Technology, 33 (2016) 125–135. https://doi.org/10.1016/j.jddst.2016.02.009. M. Riaz, M. Riaz, X. Zhang, C. Lin, K. Wong, X. Chen, G. Zhang, A. Lu, & Z. Yang, Surface functionalization and targeting strategies of liposomes in solid tumor therapy: A review. International Journal of Molecular Sciences, 19 (2018) 195. https://doi.org/10.3390/ijms19010195. N. D. James, R. J. Coker, D. Tomlinson, J. R. W. Harris, M. Gompels, A. J. Pinching, & J. S. W. Stewart, Liposomal doxorubicin (Doxil): An effective new treatment for Kaposi’s sarcoma in AIDS. Clinical Oncology, 6 (1994) 294–296. https://doi.org/10.1016/S09366555(05)80269-9. Z. Li, S. Tan, S. Li, Q. Shen, & K. Wang, Cancer drug delivery in the nano era: An overview and perspectives. Oncology Reports, 38 (2017) 611–624. https://doi.org/10.3892/or.2017.5718. R. Trivedi & U. B. Kompella, Nanomicellar formulations for sustained drug delivery: Strategies and underlying principles. Nanomedicine, 5 (2010) 485–505. https://doi.org/10.2217/nnm.10.10. Y. Li, A. Yu, L. Li, & G. Zhai, The development of stimuliresponsive polymeric micelles for effective delivery of chemotherapeutic agents. Journal of Drug Targeting, 26 (2018) 753–765. https://doi.org/10.1080/1061186X.2017.1419477. A. M. Brinkman, G. Chen, Y. Wang, C. J. Hedman, N. M. Sherer, T. C. Havighurst, S. Gong, & W. Xu, Aminoflavone-loaded EGFRtargeted unimolecular micelle nanoparticles exhibit anti-cancer effects in triple negative breast cancer. Biomaterials, 101 (2016) 20– 31. https://doi.org/10.1016/j.biomaterials.2016.05.041. H. Hemeg, Nanomaterials for alternative antibacterial therapy. International Journal of Nanomedicine, 12 (2017) 8211–8225. https://doi.org/10.2147/IJN.S132163. L. Palmerston Mendes, J. Pan, & V. Torchilin, Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules, 22 (2017) 1401. https://doi.org/10.3390/molecules22091401.

50

30.

31.

32.

33. 34.

35.

36.

37.

38.

Chapter 2

E. Abbasi, S. F. Aval, A. Akbarzadeh, M. Milani, H. T. Nasrabadi, S. W. Joo, Y. Hanifehpour, K. Nejati-Koshki, & R. Pashaei-Asl, Dendrimers: Synthesis, applications, and properties. Nanoscale Research Letters, 9 (2014) 247. https://doi.org/10.1186/1556-276X9-247. D. Kong, J. Liu, Liu, Chu, Wang, Duan, Feng, Wang, & Yang, Novel peptide–dendrimer conjugates as drug carriers for targeting nonsmall cell lung cancer. International Journal of Nanomedicine, (2010) 59. https://doi.org/10.2147/IJN.S14601. C. I. C. Crucho & M. T. Barros, Polymeric nanoparticles: A study on the preparation variables and characterization methods. Materials Science and (QJLQHHULQJ C, 80 (2017) 771–784. https://doi.org/10.1016/j.msec.2017.06.004. Z. Zeng, Recent advances of chitosan nanoparticles as drug carriers. International Journal of Nanomedicine, (2011) 765. https://doi.org/10.2147/IJN.S17296. C. Matea, T. Mocan, F. Tabaran, T. Pop, O. Mosteanu, C. Puia, C. Iancu, & L. Mocan, Quantum dots in imaging, drug delivery and sensor applications. International Journal of Nanomedicine, 12 (2017) 5421–5431. https://doi.org/10.2147/IJN.S138624. M.-X. Zhao & B.-J. Zhu, The research and applications of quantum dots as nano-carriers for targeted drug delivery and cancer therapy. Nanoscale Research Letters, 11 (2016) 207. https://doi.org/10.1186/s11671-016-1394-9. P. Mitchell, N. Thatcher, M. A. Socinski, E. Wasilewska-Tesluk, K. Horwood, A. Szczesna, C. Martín, Y. Ragulin, M. Zukin, C. Helwig, M. Falk, C. Butts, & F. A. Shepherd, Tecemotide in unresectable stage III non-small-cell lung cancer in the phase III START study: Updated overall survival and biomarker analyses. Annals of Oncology, 26 (2015) 1134–1142. https://doi.org/10.1093/annonc/mdv104. G. Chen, L. Sheng, & X. Du, Efficacy and safety of liposomepaclitaxel and carboplatin based concurrent chemoradiotherapy for locally advanced lung squamous cell carcinoma. Cancer Chemotherapy and Pharmacology, 82 (2018) 505–510. https://doi.org/10.1007/s00280-018-3640-6. S. Y. Lee, H. S. Park, K. Y. Lee, H. J. Kim, Y. J. Jeon, T. W. Jang, K. H. Lee, Y. C. Kim, K. S. Kim, in J. Oh, & S. Y. Kim, Paclitaxelloaded polymeric micelle (230 mg/m2) and cisplatin (60 mg/m2) vs. paclitaxel (175 mg/m2) and cisplatin (60 mg/m2) in advanced non– small-cell lung cancer: A multicenter randomized phase IIB trial.

Advances in Nanoparticles for Lung Cancer Therapy

39.

40.

41.

42.

43.

51

Clinical Lung Cancer, 14 (2013) 275–282. https://doi.org/10.1016/j.cllc.2012.11.005. Y. Gu, J. Li, Y. Li, L. Song, D. Li, L. Peng, Y. Wan, & S. Hua, Nanomicelles loaded with doxorubicin and curcumin for alleviating multidrug resistance in lung cancer. International Journal of Nanomedicine, 11 (2016) 5757–5770. https://doi.org/10.2147/IJN.S118568. Q. Zhong, E. R. Bielski, L. S. Rodrigues, M. R. Brown, J. J. Reineke, & S. R. P. da Rocha, Conjugation to poly(amidoamine) dendrimers and pulmonary delivery reduce cardiac accumulation and enhance antitumor activity of doxorubicin in lung metastasis. Molecular Pharmaceutics, 13 (2016) 2363–2375. https://doi.org/10.1021/acs.molpharmaceut.6b00126. L. M. Kaminskas, V. M. McLeod, G. M. Ryan, B. D. Kelly, J. M. Haynes, M. Williamson, N. Thienthong, D. J. Owen, & C. J. H. Porter, Pulmonary administration of a doxorubicin-conjugated dendrimer enhances drug exposure to lung metastases and improves cancer therapy. Journal of Controlled Release, 183 (2014) 18–26. https://doi.org/10.1016/j.jconrel.2014.03.012. Zheng, Weekly intravenous nanoparticle albumin-bound paclitaxel for elderly patients with stage IV non-small-cell lung cancer: a series of 20 cases. Journal of Biomedical Research, 26 (2012). https://doi.org/10.7555/JBR.26.20110106. B. C. Tang, J. Fu, D. N. Watkins, & J. Hanes, Enhanced efficacy of local etoposide delivery by poly(ether-anhydride) particles against small cell lung cancer in vivo. Biomaterials, 31 (2010) 339–344. https://doi.org/10.1016/j.biomaterials.2009.09.033.

CHAPTER 3 METAL NANOPARTICLES FOR CANCER THERAPY RAJAN RAJABALAYA*1 AND SHEBA RANI DAVID2 1

PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2 School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA *Corresponding author: Dr Rajan Rajabalaya Senior Assistant Professor PAPRSB Institute of Health Sciences Universiti Brunei Darussalam Jalan Tungku Link BE1410 Bandar Seri Begawan Brunei Darussalam Email: [email protected] Phone: + 6732460922

Abstract Cancer is the most common non-communicable disease and a leading cause of death worldwide. Various treatments, such as chemotherapy, radiotherapy and surgical interventions, are available for cancer therapeutics. However, the effectiveness of the current cancer therapeutics is still limited because they cannot reach the target site specifically, hence affecting the normal tissues and cells. Moreover, they may result in toxicity and severe side effects. With the use of nanotechnologies, nanoparticles, specifically metal nanoparticles, have been used widely in cancer treatment because they offer numerous advantages to improving patients' health. Moreover, they are used to overcome the issues associated with conventional cancer therapeutics. This chapter aims to review the applications of metal nanoparticles for cancer therapy. Several articles have shown that metal nanoparticles are effective for cancer therapy, hence improving patient compliance. However,

Metal Nanoparticles for Cancer Therapy

53

these nanoparticles also have toxicological issues. Therefore, future research is required to improve the applications of metal nanoparticles for future cancer therapeutics.

Abbreviations 5-FU bFGF EGF GSH nm PDGF siRNAs TGA TGF-ȕ VEGF

5-fluorouracil Basic fibroblast growth factor Epidermal growth factor Glutathione Nanometre Platelet-derived growth factor Small interfering RNAs Thioglycolic acid Transforming growth factor-beta Vascular endothelial growth factor

1. Introduction Cancer is widely known as one of the leading causes of mortality and morbidity worldwide. Moreover, the burden of cancer is expected to rise to 18.1 million cases, with 9.6 million deaths from cancer globally. Cancer is a non-communicable disease occurring because of genetic and environmental factors. Thus, it changes the properties of the normal cells, leading to tumour cells [1]. Early detection of cancer may help increase its five-year survival rate. For instance, the early detection of breast cancer may result in a fiveyear survival rate between 93 and 100%. However, the later detection decreases the survival rate between 22 and 77% [2]. Patients with cancer may experience several symptoms. Thus, these symptoms might affect their quality of life as an individual. Hence, alleviating the symptoms with treatment may help them relieve the suffering and improve their quality of life [3]. Numerous treatments, such as surgery, radiation, hormone therapy and chemotherapy, are available managing for cancer depending on its type and stages [4]. However, nowadays, the effectiveness of cancer therapeutics is limited because they cannot reach the target site with sufficient concentrations that exert a pharmacological effect. Moreover, it may affect normal tissues and cells, leading to toxicity and severe side effects [1].

54

Chapter 3

For instance, cisplatin is one of the chemotherapy drugs used to treat patients with breast, lung and ovarian cancer [5]. Cisplatin binds covalently to DNA and interferes with DNA functions. When cisplatin enters the cells, the chloride ligands are substituted by water molecules. Consequently, positively charged platinum complexes are formed. Cisplatin then reacts with the nucleophilic sites on DNA through intra-stand and interstrand cross-links, which form cisplatin-DNA adducts and prevent the formation of DNA, RNA and protein. About 20–30% of patients were reported to experience nephrotoxicity due to cisplatin [5]. Additionally, 3–20% of patients were administered doxorubicin, a drug commonly administered for breast cancer [6]. Doxorubicin binds directly to the DNA through intercalation between base pairs of DNAs. Further, it inhibits the enzyme topoisomerase II, thus restricting the DNA repair. Because of this action, it also blocks the formation of DNA and RNA [7]. Doxorubicin was reported to be developed with side effects of cardiotoxicity [6]. Table 3-1 shows some side effects of standard cancer therapeutics.

6-mercaptopurine

Gemcitabine

Methotrexate

Cyclophosphamide

Drug Cisplatin

Leukaemia

Type of cancer Bladder Head and neck Lung Ovarian Testicular Breast cancer Lung cancer Ovarian Leukaemia Lymphoma Bladder Breast Gastric Head and neck Leukaemia Lymphoma Lung Pancreatic Bladder Incorporation of gemcitabine diphosphate into DNA with the assistance of gemcitabine triphosphate Inhibition of DNA synthesis Induction of apoptosis Inhibition of purine synthesis

Inhibition of enzyme dihydrofolate reductase and thymidylate synthase Inhibition of cell division

Bind to DNA Crosslink to DNA and RNA strands Inhibition of protein synthesis

Mechanism of action Binds to DNA Interfere function of DNA Induction of apoptosis

Anaemia Leukopenia Thrombocytopenia Stomatitis Hepatotoxicity

Anaemia Neutropenia Thrombocytopenia

Neutropenia Thrombocytopenia

Myelosuppression Cardiac dysfunction Haemorrhagic cystitis

Side effects Nephrotoxicity Myelosuppression Ototoxicity

Table 3-1: Current cancer therapeutics, its mechanism of action and side effects

Metal Nanoparticles for Cancer Therapy

[12]

[11]

[10]

[9]

Reference [8]

55

Bladder Brain Cervical Head and neck Lung Lymphoma Ovarian Prostate Testicular

Breast Lung Ovarian

Brain Breast Cervical Colorectal Leukaemia Lung Lymphoma Melanoma Ovarian

Etoposide

Paclitaxel

Vincristine

56

Prevents the polymerisation of tubulin to form microtubules Induction of depolymerisation of formed tubules

Binds to tubulin to form stable, nonfunctional microtubules Microtubules block cells in the M phase Inhibition of cell division Induction of apoptosis

Inhibition of DNA topoisomerase II Inhibition of DNA synthesis

Chapter 3 Type 1 hypersensitivity Myelosuppression Fatigue Alopecia Anorexia Constipation Diarrhoea Mucositis Nausea and vomiting Stomatitis Taste alteration Acute leukaemia Anaemia Febrile neutropenia Leukopenia Neutropenia Cardiovascular events Nausea Vomiting Hypersensitivity Peripheral neuropathy [15]

[14]

[13]

Breast

Doxorubicin

Tamoxifen

EGFR = Epidermal growth factor receptor VEGF = Vascular endothelial growth factor

Cetuximab

Cervical Larynx and Para larynx Lymphoma Renal Testicular Vulva Lung Head and neck Leukaemia Lymphoma Lung Gastric Pancreatic Breast Colorectal Head and neck

Bleomycin

Binds to EGFR Inhibition of cell growth Induction of apoptosis Decreased VEGF production Binds to oestrogen receptors Decreases DNA synthesis Induction of apoptosis

Binds to DNA Inhibition of DNA repair Blockade of DNA and RNA synthesis and fragmentation of DNA

Breakage of DNA strand Inhibition of RNA synthesis Inhibition of protein synthesis

Metal Nanoparticles for Cancer Therapy

Infusion reactions Cardiopulmonary arrest Malaise Fever Thromboembolic events Hot flashes Vaginal bleeding Hypercalcaemia Cataracts Endometrial cancer

Cardiotoxicity

Rash Pneumonitis

[18]

[17]

[7]

[16]

57

Chapter 3

58

2. Nanoparticles Nanoparticles are particles with a size range between 10 and 1000 nm [19]. The size of nanoparticles depends on their use. For instance, the size of nanoparticles synthesised for the delivery of drugs is usually 100 nm or higher to allow enough supply of drugs. Likewise, the ideal size of nanoparticles for cancer therapy is between 70 and 200 nm because the endothelial fenestrations in a growing tumour are approximately 200 to 780 nm. The shape of nanoparticles can have the shapes of rods, triangles, round, octahedral and polyhedral. The preparations of nanoparticles can be either engineered in a laboratory or nature comprising organic and inorganic compounds. Nanoparticles can be prepared with different techniques, such as physical, chemical and biological, in the laboratory. Various nanoparticles, such as quantum dots, dendrimer, liposomes, carbon nanotubes, magnetic, metallic and polymer, are available [20]. Nanoparticles have been widely used in many fields nowadays, especially in medicine. They offer various advantages in improving human health [1,22]. It also offers tools for diagnosing and treating cancer. The use of nanotechnologies for cancer overcomes the limitations of conventional cancer therapeutics because they are more specific and may deliver molecular therapy in the bloodstream undetected by the immune system, hence minimising the side effects [1].

2.1 Synthesis of metal nanoparticles Metal nanoparticles are nanoparticles used for treating cancer. Gold, silver, iron, platinum, titanium and zinc can be used as metal nanoparticles. There are two methods to group the synthesis of metal nanoparticles: bottom-up and top-down methods. The bottom-up method is defined as the method in which nanoparticles are built from molecules or atoms. In contrast, the topdown method involves disintegrating the bulk materials into smaller pieces, resulting in nanoparticles. Nonetheless, it is recommended to classify the procedures to synthesise metal nanoparticles into physical, chemical and biological forms [1]. 2.1.1 Physical methods There are two physical methods to synthesise metal nanoparticles: evaporation-condensation and laser ablation. The advantages of using physical methods to prepare metal nanoparticles are the absence of contamination and a uniform distribution of nanoparticles compared to chemical synthesis [23]. First, evaporation-condensation is a gas phase procedure using a horizontal tube furnace at atmospheric pressure

Metal Nanoparticles for Cancer Therapy

59

conditions to produce nanoparticles. The furnace is a vessel that helps carry the synthesising metal source material vaporised into the carrier air. For this process, a change or modification in the reactor system controls the generated nanoparticles’ yield, concentration and size [20]. However, the tube furnace has a large capacity that uses more energy to increase the environmental temperature around the source material. Moreover, it takes time to reach a stable temperature [23]. Laser ablation is a process of laser irradiation to a region of the solid target material that contributes to its removal. The laser ablation of a solid placed in a liquid medium can synthesise nanoparticles. After the laser irradiation, the liquid will not contain other ions, compounds and reducing agents, except for the intended solid nanoparticles. Several factors, such as the nature of liquid media, laser fluence, target solid and duration of irradiation, affect the ablation mechanism and characteristics of metal nanoparticles produced. Chemical reagents are not present in physical methods. Therefore, the basic liquid medium is used with mild surfactants. Hence, the production of nanoparticles using laser ablation is also pure and not contaminated as chemical synthesis. 2.1.2 Chemical methods A reduction in the metal salts is the chemical method of metal nanoparticles synthesis. In this method, the stabilising agent is added to prevent the nanoparticles from aggregating. The type and strength of the reducing agent and the stabiliser may influence the finished product’s shape, size and other properties. The number of metal salts can be adjusted depending on the required metal nanoparticles. Such salts can be adjusted by adding chemical agents such as amino boranes, hydrazine, oleylamine, polyols, oxalic acid, sugar and citrate [20]. However, some absorbed toxic chemicals are present on the surface and lead to adverse effects [24]. 2.1.3 Biological methods Finally, this method prevents costly and undesirable techniques that use poisonous chemicals such as reducing agents, surfactants and stabilisers. Hence, it is a safe and better approach for synthesising nanoparticles. The use of green products as biological methods is considered as inexpensive, biocompatible, safer and non-toxic. It is mainly a one-step synthesis using biomolecules; carbohydrates, polymers and proteins are extracted from whole organisms, such as algae, fungi, bacteria and plants, as reduction agents for metal salts. The green synthesis also produces pure nanoparticles

Chapter 3

60

compared to chemical methods, wherein the composition of the generated nanoparticles is contaminated with the chemicals used throughout the procedures. The metal salt solution is incubated with the biological agent in a suitable medium to synthesise nanoparticles according to their shape, size and surface properties. Overall, algal or plant systems have been used commonly to generate nanoparticles concerning efficiency, safety and time. In contrast, the less preferable would be fungal, bacterial and viral-mediated synthesis because of the culturing of these microorganisms and the limited size and shapes of nanoparticles produced by them in addition to a slower rate [20]. For example, silver and gold nanoparticles were synthesised from extracting a plant known as Acalypha indica. Both silver and gold nanoparticles were used to investigate cytotoxicity effects on MDA-MB-231, human breast cancer cells at different concentrations. The concentrations ranged from 1 to 100 ȝg/ml. The results reported that the synthesised metal nanoparticles at 100 ȝg/ml had shown a significant cytotoxic effect and apoptotic features. Hence, the results indicated that the biologically synthesised silver and gold nanoparticles might be used to treat breast cancer. However, a lot of research may be needed to ensure their potential as an anticancer agent [25]. Table 3-2 highlights the salient examples of biological agents synthesising metal nanoparticles. Table 3-2: Use of biological agents to synthesise metal nanoparticles for cancer therapy Biological agent Bacteria Bacillus funiculus Enterococcus sp. Bacillus licheniformis Escherichia fergusoni Bacillus clausii

Metal used

Shape

Size (nm)

Application

Reference

Silver

Spherical

20

Breast cancer

[26]

Gold

Spherical

6–12

[27]

Silver

Spherical

50

Silver

Spherical

10–50

Lung cancer Liver cancer Dalton’s lymphoma Breast cancer

Silver

Spherical

16-20

Ovarian cancer

[30]

[28] [29]

Metal Nanoparticles for Cancer Therapy Plants Nigella sativa Nyctanthes arbortristis Couroupita guianensis

Silver Silver

Spherical -

15 -

Gold

61

7–48

Cytotoxicity Antibacterial and cytotoxic Leukaemia

[31] [32] [33]

30

Breast cancer

[25]

Acalypha Indica Acalypha Indica Dysosma pletantha rhizome Morinda citrifolia L root

Silver

Spherical, triangular, tetragonal and pentagonal Spherical

Gold

Spherical

30

Breast cancer

[25]

Gold

Spherical

127

Fibrosarcoma

[34]

Gold

12.17– 38.26

Anticancer

[35]

Citrullus colocynthis Euphorbia nivulia Curcumin

Silver

Triangle and spherical Spherical

31

[36]

Copper

Spherical

5–15

Laryngeal cancer Lung cancer

Copper

Irregular

Breast cancer

[38]

Dioscorea bulbifera Mentha piperita Sesbania grandiflora Melia azedarach

Platinumpalladium Platinum

Rodshaped Irregular

20–25

Anticancer

[39]

Spherical

54.3

Colon cancer

[40]

Silver

Spherical

22

Breast cancer

[41]

Silver

78

HeLa cells

[42]

Melia azedarach

Silver

78

Lymphoma mice model

[42]

Ficus religiosa

Silver

Cubical and spherical Cubical and spherical Spherical

5–35

[43]

Pimpinella anisum seeds

Silver

Spherical

3.2

Moringa oleifera flower

Gold

Spherical Hexagonal Triangular

3–5

Dalton’s lymphoma Human neonatal skin stromal cells and colon cancer cells A549 human lung cancer

[37]

[44]

[45]

Chapter 3

62 Fungi Agaricus bisporus Saccharomyces boulardii Saccharomyces boulardii

Silver

Spherical

8–20

Breast cancer

[46]

Platinum

Spherical

90

[47]

Platinum

Spherical

90

A431 cell lines Breast cancer

[47]

2.2 Characteristics metal nanoparticles The characteristics of metal nanoparticles are expected to result in better penetration for the treatment used within the body. Hence, it provides an efficient therapy with minimal risk compared to conventional drug therapy. The characteristics of metal nanoparticles are a high surface-to-volume ratio, broad optical properties, ease of synthesis, achievable surface chemistry and functional properties. Metal nanoparticles have broad optical properties that allow them to be easily adjusted based on the desired wavelengths per the nanoparticles’ shape (nanoparticles, nanoshells, nanorods etc.), size (e.g., 1–100 nm) and composition (e.g., core/shell or noble alloy metals). Therefore, these properties enable photothermal applications under native tissues. Further, metal nanoparticles can be easily functionalised with several moieties, such as peptides, antibodies and/or DNA/RNA, to target dissimilar cells specifically and with biocompatible polymers such as polyethylene glycol. Therefore, metal nanoparticles can prolong the applications of drugs and genes in vivo. Further, metal nanoparticles can convert light or radio frequencies to heat, allowing target cancer cells to be thermally killed [1]. Table 3-3 shows some of the available characteristics of metal nanoparticles in biomedical applications. Table 3-3: Characteristics of metal nanoparticles Metal nanoparticles Gold

Gold

Gold

Characteristics

Reference

Biocompatible. A soft acid able to bind strongly with soft bases such as thiols and amine functionalities Its functionalisation with organic molecules allows for conjugation with antibodies, drug molecules or ligands, whether active or even passive drug delivery. The gold core of gold nanoparticles is inert and non-toxic Easy to synthesise

[48]

[49]

[50]

Metal Nanoparticles for Cancer Therapy

Gold Silver

Adjustable diameter (ranging from 1 to 150 nm) Various shapes are available. Adjustable optical properties by changing the core diameter and shell wall thickness Free electron system that contains an even number of positive ions and fee-moving electrons Exhibits the highest efficiency of plasmon excitation Tunable plasmon resonance at any wavelength Ability to absorb light for thermal killing Free electron system that contains an even number of positive ions and fee-moving electrons

63

[51] [51]

2.3 Advantages of metal nanoparticles Compared to conventional cancer therapeutics, metal nanoparticles offer several anticancer drug delivery features. These features enhance nanoparticles’ ability to deliver drugs with poor solubility in water. These drugs have better penetration into the tumour cells with a more significant dose. As pH-sensitive or temperature-sensitive agents, they can deliver the drug in the acidic environment of the cancer cells withstanding temperature changes, respectively [52]. Moreover, nanoparticles’ high specificity can target the specific tissue or cell with minimal side effects. Other advantages are that they can control the released drugs over a period, and they can deliver multiple types of drugs as combination therapy [53]. However, the use of metal nanoparticles in drug delivery has become a concern. Indeed, once the drug is administered, the metal nanoparticles can stay in the body even though they are inert and biocompatible [54].

2.4 Applications of metal nanoparticles for cancer therapy Metal nanoparticles' unique physical and chemical characteristics facilitate their use in cancer therapeutics such as drug delivery, antiangiogenic agent, radiotherapy, gene silencing, and hyperthermia. The evidence of using metal nanoparticles in cancer therapy and some of the applications of metal nanoparticles in cancer therapy are presented in Tables 3-4, 3-5, 3-6, 3-7 and 3-8.

Chapter 3

Carriers

Gold

Gold

Gold Gold

Gold

Gold

Gold

Silver

Silver

Silver

Gold

Drug

5-FU

5-FU

Topical 5-FU Doxorubicin

Doxorubicin

Doxorubicin

Doxorubicin

Doxorubicin

Doxorubicin

Paclitaxel

Paclitaxel

Binds to tubulin to form stable, nonfunctional microtubules

Binds to DNA Inhibition of DNA repair Blockade of DNA and RNA synthesis and fragmentation of DNA

Irreversible inhibition of thymidylate synthase Inhibition on the synthesis of DNA or RNA synthesis

Mechanism of action

Liver tumour (HepG2) cells Liver tumour (HepG2) cells

MCF7 cells

Human Caucasian hepatocyte cells T47D cells

A549, H460 and H520 lung cancer cells

M139 and M213 cells HaCaT Human breast cancer cells Pc-3 cancer cell

Biological target site Colorectal cancer

Higher permeability Inhibits the proliferation of breast cancer cells Inhibits the proliferation of PC-3 cancer cell Inhibition of lung cancer cells growth Upregulates the expression of tumour-suppressor genes Induction of apoptosis in lung cancer cells Gives cytotoxic effect based on the concentration Inhibits the proliferation of tumour cells Inhibits the proliferation of tumour cells Induction of apoptosis Enhanced cytotoxic effect Inhibition of tumour growth and reduction of tumour size

Induction of apoptosis and stopped the cell cycle progression Increased cytotoxic effect

Pharmacological effects

[64]

[63]

[62]

[62]

[61]

[60]

[59]

[57] [58]

[56]

[55]

References

Table 3-4 shows the evidence supporting the use of metal nanoparticles as drug delivery for cancer treatment

64

Carriers

Gold

Gold Gold

Gold

Gold

Gold Silver

Gold

Gold Platinum

Platinum

Drug

Paclitaxel

Paclitaxel Paclitaxel

Paclitaxel

Paclitaxel

Paclitaxel Etoposide

Etoposide

Cisplatin Cisplatin

Cisplatin

Inhibition of DNA topoisomerase II Inhibition of DNA synthesis Binds to DNA Interfere function of DNA Induction of apoptosis

Microtubules block cells in the M phase Inhibition of cell division Induction of apoptosis

Mechanism of action

U87 glioma cells

Lung cancer Prostate cancer

HeLa cells

HepG2 HeLa cells

B16F10 cells

H460 and H460PTX

Biological target site MDA-MB-231 cells Lung cancer MDA-MB-231

Metal Nanoparticles for Cancer Therapy

Decreases the size of tumour Induction of apoptosis Improve reactive oxygen species generation Altered mitochondrial membrane potential Greater cytotoxic effect Maintain a greater cytotoxic effect in drug resistant tumour cells Ability to deliver paclitaxel with the improvement of anticancer efficiency Inhibition of tumour cells A significant decline in cell viability A significant decline in cell viability Decreases the size of tumour Provides a greater cytotoxic effect in comparison to free cisplatin Provides pro-apoptotic effect on cancer cells

Greater cytotoxic effect

Pharmacological effects

[74]

[72] [73]

[71]

[70] [71]

[69]

[68]

[66] [67]

[65]

References

65

Chapter 3

10-25

10-25

10-25

15 ± 2

5-35

16.5

Silver

Gold

Silver

Gold

Silver

Silver

Gold

Size (nm) 10-25

Metal used

In vivo

In vitro

In vitro

In vivo

In vivo

In vitro

Type of study In vitro

Chicken Chorioallantoic Membrane

Breast cancer cell lines (MCF-7 and MDA-MB-231) Rat aortic ring model

Mouse Matrigel Model

Chicken chorioallantoic membrane Chicken chorioallantoic membrane Mouse Matrigel Model

Model

50% reduction in the number of formation of new blood vessels. Hence, it shows therapeutic benefits against angiogenesis. The number and length of blood vessels formation had reduced significantly.

Enhance the antiangiogenic effects Enhance the antiangiogenic effects Enhance the antiangiogenic effects Enhance the antiangiogenic effects Inhibition of angiogenesis

Results

[78]

[77]

[76]

[75]

[75]

[75]

[75]

References

Table 3-5: Evidence supporting the use of metal nanoparticles as an antiangiogenic agent for cancer treatment

66

In vitro

14, 50 and 74

1.9

20, 50 and 100

20

10.8

11

14-74

29, 36, 42 and 52

Gold

Silver

Silver

Gold

Gold

Gold

Platinum

In vitro

In vitro

In vivo

In vitro

In vitro

In vitro

In vivo

Type of study

Size (nm)

Type of metal nanoparticles Gold

HeLa cells

Mouse fibroblast cell lines 3T3 Breast cancer cells Brain tumours in mice HeLa cells

Tumour-bearing mice Glioma cells

HeLa cells

Biological target site

Gold nanoparticles increase the rate of killing cancer cells It was demonstrated that the mice had 50% long-term tumour-free survival Gold nanoparticles with 50 nm have shown better radiosensitisation enhancement factors than gold nanoparticles of 14 and 74 nm. It has shown radiosensitisation for all sizes of platinum nanoparticles.

50 nm of gold nanoparticles had better radio sensitising ability in comparison to 14 and 74 nm gold nanoparticles. It shows 86% of one-year survival in comparison to radiotherapy alone. Enhance the effect of radiation on glioma cells 20 and 50 nm of silver nanoparticles have shown better radiation sensitivity in comparison to 100nm of silver nanoparticles Enhance the efficacy of radiotherapy

Results

Table 3-6: Evidence supporting the use of metal nanoparticles as radiotherapy for cancer treatment

Metal Nanoparticles for Cancer Therapy

[86]

[85]

[84]

[83]

[82]

[81]

[80]

[79]

References

67

Chapter 3

40

40

15

90

Gold

Gold

Gold

Size (nm)

Type of metal nanoparticles Gold

In vitro

In vitro

In vitro

Type of studies In vitro

Breast cancer cells

Human breast cancer cell lines Human prostate cancer cell lines HeLa cells

Model siRNA was released gradually and provided an extended gene silencing effect siRNA was released gradually and provided an extended gene silencing effect Improves the efficiency in delivering siRNA. Improves the delivery of siRNA.

Results

[89]

[88]

[87]

[87]

References

Table 3-7: Evidence supporting the use of metal nanoparticles as gene silencing for cancer treatment

68

20.6 ± 2.7 53.8 ± 7.6 137.3 ± 43.0

76.1 10-30

-

Silver Silver

Gold

-

Size (nm)

Silver

Type of metal nanoparticles Gold

In vitro

In vitro In vitro

In vitro

Type of studies In vitro

Breast cancer cells Human breast cancer MCF-7 cell lines Human prostate cancer cell lines

Glioma cells

Kb cells

Model The gold nanoparticles transduced NIR light into heat to target the tumour cells. Increased the rate of apoptosis depending upon the sizes of silver nanoparticles Induction of cell death Induction of hyperthermia Increases the rate of cell death

Results

[94]

[92] [93]

[91]

[90]

References

Table 3-8: Evidence supporting the use of metal nanoparticles as hyperthermia for cancer treatment

Metal Nanoparticles for Cancer Therapy 69

70

Chapter 3

3. Tumour targeting The essential aspects of metal nanoparticles in tumour targeting therapy are selectivity and specificity. Consequently, the metal nanoparticles decrease the chemotherapeutic toxicity and adverse effects by delivering numerous therapeutic agents to the targeted tumour site, circumventing the biological barriers. The applications of metal nanoparticles with antitumour agents may help synergise the antitumour activity. There are two ways for the metal nanoparticles to target the tumour cells: passive targeting and active targeting. In passive targeting, an increase in permeability and retention effect may cause the metal nanoparticles to accumulate more in the tumour tissues compared to normal tissues. Tumour tissues have leaky blood vessels, fenestrations and poor lymphatic drainage. Consequently, nanoparticles can permeate the tumour cells to kill them. In addition to the unique structures and enhanced permeability of metal nanoparticles, they easily target cancerous cells. Further, metal nanoparticles are protected from macrophages-mediated uptake because of hydrophilic molecules on the nanoparticles’ surface. Hydrophilic molecules increase nanoparticles’ solubility and half-life, thus preventing them from being removed from systemic circulation. Meanwhile, metal nanoparticles are functionalised with biomolecules or antibodies, DNA/RNA and ligands in active targeting per the complimentary cell surface protein or receptors on cancerous cells. Compared to passive targeting, active targeting has better selectivity and specificity in tumour cells with minimal (or absent) damage to healthy cells. Figure 3-1 represents active and passive targeting in metal nanoparticles.

Metal Nanoparticles for Cancer Therapy

71

Figure 3-1: Active targeting of metal nanoparticles functionalised with ligands and have greater selectivity and specificity to cancer cells and passive targeting of metal nanoparticles that target inflamed and tumour cells, resulting in a lack of specificity [51]. Created with BioRender.com

72

Chapter 3

3.1 Drug delivery carriers In this era, the difficulties in targeting drugs into specific tissues and organs have been known as one of the challenges in cancer therapy because the conventional drugs used in cancer therapeutics have a low molecular weight. Consequently, these drugs diffuse into the tumour tissues and cells and the healthy tissues and cells. Moreover, they have a short half-life and a high clearance rate. Hence, fewer drugs reach their target site, decreasing therapeutic efficacy. However, there is an increase in the side effects. Metal nanoparticles have been used for drug delivery in cancer therapy to address this concern. Factors affecting the performance of metal nanoparticles are their size, drug release rate and rate of particle disintegration. Metal nanoparticles also can be used as targeted drug delivery to increase the permeability of the drug and overcome the first-pass metabolism [95]. In preclinical studies, nanoparticles have been used as drug carriers of chemotherapeutic drugs, such as hydroxycamptothecin, 5-fluorouracil (5FU), docetaxel and gemcitabine for lung, colon, squamous and pancreatic cancers. Moreover, other preclinical studies have shown that nanoparticles loaded with TNF-Į have helped enhance the malignant cells. Gold nanoparticles have shown to be a better type to use as a potential drug delivery carrier due to their unusual physical and chemical properties compared to other metal nanoparticles [96]. For instance, it has been used as a drug delivery carrier for paclitaxel or platinum-based drugs such as cisplatin and oxaliplatin [1]. However, gold nanoparticles are not biodegradable. Because of their properties related to surface modification, it can affect the properties of the transported drugs, such as toxicity, biodistribution and pharmacokinetics [96]. A previous study used gold nanoparticles in the drug delivery of 5-FU to treat colorectal cancer cells. 5-FU is an antimetabolite drug used for treating several cancers of different regions (breast, colon, rectum and stomach). However, 5-FU demonstrated numerous side effects that could influence its pharmacological effect, including myelosuppression, cardiotoxicity and chest pain [55]. Hence, gold nanoparticles were used to deliver 5-FU to increase the anticancer effect and reduce the side effects. The 5-FU was loaded into the gold nanoparticles using ligands containing two thiols, namely thioglycolic acid (TGA) and glutathione (GSH). The gold nanoparticles were made at different 5-FU/ligand molar ratios and assessed with different techniques. Flow cytometry was used to study the efficacy of the anticancer effect in the cancerous tissues of patients with colorectal cancer. The shape and size of the gold nanoparticles were spherical and between 9 and 17 nm, respectively. The salt concentration and pH solution

Metal Nanoparticles for Cancer Therapy

73

were used to stabilise the gold nanoparticles and drug release. The maximum 5-FU loading was achieved at a molar ratio of 1:1 and 2:1 for TGA-GNPs and GSH-GNPs of 5-FU/ligand, respectively. The results showed a slow release of 5-FU from gold nanoparticles. Additionally, the slow release depends on the pH of the environment. Hence, 5-FU/GSHGNPs induce cell death, known as apoptosis. Moreover, the cell cycle progression was inhibited in colorectal cancer cells. Compared to free 5-FU, gold nanoparticles showed a two-fold higher anticancer effect. Therefore, gold nanoparticles have been shown to increase the anticancer efficacy of 5-FU [55].

3.2 Antiangiogenic agent Angiogenesis is one of the factors that may cause the growth and spread of tumour cells. It is a process of forming new blood vessels and capillaries from the preexisting ones. Oxygen and other nutrients are being supplied by blood vessels to the tumour cells, resulting in the growth, migration and spreading of tumours to other parts of the organs. This process is known as metastasis. Angiogenesis in the tumour is triggered by endothelial-specific mitogens, such as the vascular endothelial growth factor (VEGF), basic fibroblast growth factor (bFGF), platelet-derived growth factor (PDGF) and transforming growth factor-beta (TGF-ȕ  The characteristics of tumour blood vessels are leaky and can permeate to plasma and heterogeneous, be it in structure or function. Therefore, cancer therapy focusing on antiangiogenic agents should be implemented. Many antiangiogenic agents have been approved. However, some of the agents were reported to cause toxicity. Metal nanoparticles, particularly gold nanoparticles, inhibit VPF/VEGF165–mediated endothelial cell proliferation in vitro and in vivo. Gold nanoparticles bind to the heparin domain and inhibit the binding of heparin growth factors such as VEFG-165 and bFGF. However, they do not inhibit the function of non-heparin growth factors such as VEGF-121 and epidermal growth factor (EGF). In vitro, the inhibition of gold nanoparticles to VEGF-165 reduces phosphorylation, intracellular calcium release and RhoA activation. Meanwhile, in vivo, gold nanoparticles were administered to a male mouse ear, initially treated with adenovirus, producing VEGF. The administration of gold nanoparticles demonstrated a reduction of angiogenesis and the formation of oedema [4].

74

Chapter 3

3.3 Radiotherapy Apart from chemotherapy and surgery, radiotherapy has been used widely to treat various cancer. Radiotherapy uses ionising radiation as cancer therapy to control the proliferation of tumour cells [1]. The radiation dose is delivered to the tumour cells without damaging the healthy tissues. Metal nanoparticles can improve the efficiency of radiotherapy by targeting the tumour cells with lower radiation doses, reducing its side effects [96]. A study was conducted to determine the potential effects of radiotherapy in killing glioma cells. It was mediated by 10, 20 and 40 nm of gold nanoparticles and 20, 50 and 100 nm of silver nanoparticles. It was all modified with foetal bovine serum proteins. The results showed that cytotoxicity depends on the radiation dose. The particles with small sizes exhibited the highest cytotoxic effect [1].

3.4 Gene silencing Small interfering RNAs (siRNAs) have been used to block gene function and sequence-specific post-transcriptional gene silencing. In other words, siRNAs play an important role in the downregulation of particular gene expressions in tumour cells. N are available to transfect siRNAs into mammalian cells, which may affect the strength and duration of the silencing response. This is affected by the amount of delivered siRNA and the ability of each siRNA to suppress its target. However, some problems may arise because siRNAs have short half-lives, weak protection against action and poor chemical stability. Moreover, it is still doubtful to deliver RNAs therapy with a high therapeutic impact. Therefore, nanotechnology has been introduced to address these challenges. Due to the small size of nanoparticles, they can interact with biomolecules on the surface and inside the cells longer. For example, gold nanoparticles have shown potential as intracellular delivery vehicles for antisense oligonucleotides and therapeutic siRNA to protect against RNAs and the ease of functionalisation for selective targeting [1].

3.5 Hyperthermia Using heat for cancer treatment is not a new concept. It was popular in the late 1800s. The use of hyperthermia for cancer treatment has been proven successful [97]. Hyperthermia necessitates a higher temperature on living cells. Hyperthermia can affect the denaturation of extracellular proteins and blood to induce apoptosis. A temperature of 42ºC may help reduce the cell

Metal Nanoparticles for Cancer Therapy

75

viability, and a temperature of 50ºC may result in cell death and tissue ablation. Nanoparticles can be applied in hyperthermia to increase the temperature of tumour cells until it reaches the limits of temperature tolerance to kill them selectively. The procedure for this treatment is that either the patient or the targeted area is exposed to an alternating current magnetic field. Therefore, it causes the nanoparticles to heat up and induce thermal ablation of the tumour. In 2005, the first in vivo Phase II of magnetic nanoparticles hyperthermia was conducted in Germany. It was done by injecting biocompatible metal nanoparticles into patients with prostate cancer. After numerous sessions, the results demonstrated success with a minimally invasive ablation of the tumour in an alternating current magnetic field [1].

3.6 Toxicity However, metal nanoparticles also have toxicological issues. Hence, it should be considered before using them for medical applications. It is essential to understand the toxicity of metal nanoparticles to identify the potential health risks [52]. The behaviour of nanomedicines in the body can be summarised as follows: 1. There are six ways for the nanomaterials to enter the body (intravenous, dermal, inhalation, oral, subcutaneous and intraperitoneal); 2. Absorption may occur when the nanomaterials interact with the biological components in the body; 3. Distribution can occur in numerous organs, and it may remain the same structurally or be modified or metabolised; 4. Finally, the nanomaterials will enter the cells of the organ and stay in the cells for some time before they move or leave to other organs or be eliminated. The factors that may affect the toxicity of nanoparticles include the dose, biodegradability, route of administration, solubility, pharmacokinetics, shape and structure [98]. For instance, the small size of nanoparticles in nanoparticles may be useful in cancer therapy to carry large doses; however, it is also one reason that makes it dangerous to human health [99]. Usually, the toxicity of metal nanoparticles depends on their size. A study was conducted to investigate the toxicity of silver nanoparticles based on different sizes on various microbial and mammalian cells. The result shows that silver nanoparticles of 10 nm are more toxic than silver nanoparticles of 20–80 nm. The outcomes showed that smaller metal nanoparticles have

76

Chapter 3

higher surface areas, resulting in an efficient contact between cells and particles that may lead to a higher intracellular bioavailability than bigger nanoparticles [51]. In contrast, nanoparticles with a large size may also be a disadvantage because they cannot pass through the small capillaries for drug delivery. Hence, controlling the size of metal nanoparticles could be an alternative approach to minimise toxicity [98]. However, a preclinical study showed that gold nanoparticles are known as non-toxic. No toxicity was observed when a high dose of 1.9 nm of gold nanoparticles was injected into the mice’s tail vein [4]. Moreover, a study conducted in vitro reported an interaction between nanoparticles and dye (such as the MTT assay) during the determination of cell viability, leading to invalid results. Consequently, numerous assays may be considered to confirm the toxicity of nanoparticles [98]. Overall, the toxicity of metal nanoparticles can be minimised by modifying the nanoparticles [52].

4. Challenges Although the use of metal nanoparticles helps overcome the problem associated with conventional cancer therapy concerning specificity, there are also challenges when using this strategy. For example, during the treatment course, there is a possibility of drug resistance in some cancer cells, which reduces the effectiveness of drugs released from metal nanoparticles. Hence, combinational therapies such as chemotherapy and gene therapeutics for metal nanoparticle drug delivery may effectively help deliver and target the specific cancer cells to solve this drug resistance and inhibit tumour growth. Another challenge is that the targeted nanoparticles may influence the transported drugs’ stability, solubility, and pharmacokinetics. However, there is still insufficient research on this aspect [52]. Therefore, it is essential to consider the selection of targeted metal nanoparticles, size and other properties to achieve a better targeting system [4].

5. Conclusion and future perspectives This review reaffirms the urgent need for future research to investigate the applications of metal nanoparticles in cancer therapy before their successful use as therapeutic agents. The studies should be in vitro and in vivo. Future research is also needed to show evidence of using different metal nanoparticles for cancer therapy. Further, the toxicity of metal nanoparticles is known as a significant barrier to having better metal nanoparticles in cancer therapeutics. Therefore, it should also be considered.

Metal Nanoparticles for Cancer Therapy

77

In conclusion, metal nanoparticles are known to be effective targeted nanoparticles for medical applications, specifically cancer therapeutics. It may benefit the patients since it can minimise the side effects, improving the patient’s compliance. Gold nanoparticles are the most commonly used metal nanoparticles for cancer therapy because they are non-toxic in both in vitro and in vivo environments. Moreover, many articles have supported gold nanoparticles compared to other metal nanoparticles. Hence, other metal nanoparticles can be modified so that they can be used to address future cancer therapeutics issues.

References 1. 2.

3.

4. 5. 6.

7.

8.

J. Conde, G. Doria, & P. Baptista, Noble metal nanoparticles applications in cancer. Journal of Drug Delivery, 2012 (2012) 1–12. https://doi.org/10.1155/2012/751075. M. Akhtari-Zavare, M. H. Juni, S. M. Said, I. Z. Ismail, L. A. Latiff, & S. Ataollahi Eshkoor, Result of randomized control trial to increase breast health awareness among young females in Malaysia. BMC Public Health, 16 (2016) 1–11. https://doi.org/10.1186/s12889-0163414-1. M. G. Nayak, A. George, M. S. Vidyasagar, S. Mathew, S. Nayak, B. S. Nayak, Y. N. Shashidhara, & A. Kamath, Quality of life among cancer patients. Indian Journal of Palliative Care, 23 (2017) 445– 450. https://doi.org/10.4103/IJPC.IJPC_82_17. S. Bhattacharyya, R. A. Kudgus, R. Bhattacharya, & P. Mukherjee, Inorganic nanoparticles in cancer therapy. Pharmaceutical Research, 28 (2011) 237–259. https://doi.org/10.1007/s11095-010-0318-0. R. P. Miller, R. K. Tadagavadi, G. Ramesh, & W. B. Reeves, Mechanisms of cisplatin nephrotoxicity. Toxins, 2 (2010) 2490– 2518. https://doi.org/10.3390/toxins2112490. H. Greenlee, J. Shaw, Y.-K. I. Lau, A. Naini, & M. Maurer, Lack of effect of coenzyme Q10 on doxorubicin cytotoxicity in breast cancer cell cultures. Integrative Cancer Therapies, 11 (2012) 243–250. https://doi.org/10.1177/1534735412439749. F. Yang, S. S. Teves, C. J. Kemp, & S. Henikoff, Doxorubicin, DNA torsion, and chromatin dynamics. Biochimica et Biophysica Acta (BBA) - Reviews on Cancer, 1845 (2014) 84–89. https://doi.org/10.1016/j.bbcan.2013.12.002. L. Galluzzi, L. Senovilla, I. Vitale, J. Michels, I. Martins, O. Kepp, M. Castedo, & G. Kroemer, Molecular mechanisms of cisplatin resistance. Oncogene, 31 (2012) 1869–1883.

78

9.

10.

11.

12.

13.

14. 15. 16. 17.

18.

Chapter 3

https://doi.org/10.1038/onc.2011.384. G. Voelcker, The mechanism of action of cyclophosphamide and its consequences for the development of a new generation of oxazaphosphorine cytostatics. Scientia Pharmaceutica, 88 (2020) 42. https://doi.org/10.3390/scipharm88040042. H. Tian & B. N. Cronstein, Understanding the mechanisms of action of methotrexate: implications for the treatment of rheumatoid arthritis. Bulletin of the NYU Hospital for Joint Diseases, 65 (2007) 168–73. L. de Sousa Cavalcante & G. Monteiro, Gemcitabine: Metabolism and molecular mechanisms of action, sensitivity and chemoresistance in pancreatic cancer. European Journal of Pharmacology, 741 (2014) 8–16. https://doi.org/10.1016/j.ejphar.2014.07.041. A. A. Fernández-Ramos, C. Marchetti-Laurent, V. Poindessous, S. Antonio, P. Laurent-Puig, S. Bortoli, M.-A. Loriot, & N. Pallet, 6mercaptopurine promotes energetic failure in proliferating T cells. Oncotarget, 8 (2017) 43048–43060. https://doi.org/10.18632/oncotarget.17889. J. A. Sinkule, Etoposide: A Semisynthetic Epipodophyllotoxin Chemistry, Pharmacology, Pharmacokinetics, Adverse effects and use as an antineoplastic agent. 3KDUPDFRWKHUDS\ The Journal of Human Pharmacology and Drug Therapy, 4 (1984) 61–71. https://doi.org/10.1002/j.1875-9114.1984.tb03318.x. B. A. Weaver, How taxol/paclitaxel kills cancer cells. Molecular Biology of the Cell, 25 (2014) 2677–2681. https://doi.org/10.1091/mbc.e14-04-0916. J. Škubník, V. S. 3DYOtþNRYi T. Ruml, & S. Rimpelová, Vincristine in combination therapy of cancer: emerging trends in clinics. Biology, 10 (2021) 849. https://doi.org/10.3390/biology10090849. S. M. Hecht, Bleomycin: New perspectives on the mechanism of action. Journal of Natural Products, 63 (2000) 158–168. https://doi.org/10.1021/np990549f. A. A. Argyriou & H. P. Kalofonos, Recent advances relating to the clinical application of naked monoclonal antibodies in solid tumours. Molecular medicine (Cambridge, Mass.), 15 (2009) 183–91. https://doi.org/10.2119/molmed.2009.00007. F. Yu & W. Bender, The mechanism of tamoxifen in breast cancer prevention. Breast Cancer Research, 3 (2001) A74. https://doi.org/10.1186/bcr404.

Metal Nanoparticles for Cancer Therapy

19. 20.

21.

22. 23. 24.

25.

26.

27.

28.

79

S. Bharathala & P. Sharma, Biomedical Applications of nanoparticles. Nanotechnol. Mod. Anim. Biotechnol. (Elsevier, 2019), 113–132. https://doi.org/10.1016/B978-0-12-818823-1.00008-9. H. Chugh, D. Sood, I. Chandra, V. Tomar, G. Dhawan, & R. Chandra, Role of gold and silver nanoparticles in cancer nanomedicine. Artificial Cells, Nanomedicine and Biotechnology, 46 (2018) 1210–1220. https://doi.org/10.1080/21691401.2018.1449118. G. M. 9OăVFHDQX ù Marin, R. E. ğLSOHD I. R. Bucur, M. Lemnaru, M. M. Marin, A. M. Grumezescu, & E. Andronescu, Silver nanoparticles in cancer therapy. Nanobiomaterials Cancer Ther. Appl. Nanobiomaterials (2016), 29–56. https://doi.org/10.1016/B978-0-323-42863-7.00002-5. K. McNamara & S. Tofail, Nanoparticles in biomedical applications. (2017). S. Iravani, H. Korbekandi, S. V Mirmohammadi, & B. Zolfaghari, Synthesis of silver nanoparticles: Chemical, physical and biological methods. Research in Pharmaceutical Sciences, 9 (2014) 385–406. N. Baig, I. Kammakakam, & W. Falath, Nanomaterials: A review of synthesis methods, properties, recent progress, and challenges. Materials Advances, 2 (2021) 1821–1871. https://doi.org/10.1039/D0MA00807A. C. Krishnaraj, P. Muthukumaran, R. Ramachandran, M. D. Balakumaran, & P. T. Kalaichelvan, Acalypha indica Linn: Biogenic synthesis of silver and gold nanoparticles and their cytotoxic effects against MDA-MB-231, human breast cancer cells. Biotechnology Reports, 4 (2014) 42–49. https://doi.org/10.1016/J.BTRE.2014.08.002. S. Gurunathan, J. W. Han, V. Eppakayala, M. Jeyaraj, & J. H. Kim, Cytotoxicity of biologically synthesised silver nanoparticles in MDA-MB-231 human breast cancer cells. BioMed Research International, 2013 (2013). https://doi.org/10.1155/2013/535796. S. Rajeshkumar, Anticancer activity of eco-friendly gold nanoparticles against lung and liver cancer cells. Journal of Genetic Engineering and Biotechnology, 14 (2016) 195–202. https://doi.org/10.1016/j.jgeb.2016.05.007. M. Irulappan, S. Selvaraj, B. Mani, K. Kalimuthu, & K. Sangiliyandi, Antitumour activity of silver nanoparticles in Dalton’s lymphoma ascites tumour model. International Journal of Nanomedicine, (2010) 5–753. https://doi.org/10.2147/IJN.S11727.

80

29.

30.

31.

32.

33.

34.

35.

36.

Chapter 3

S. Gurunathan, J. W. Han, A. A. Dayem, V. Eppakayala, J. H. Park, S.-G. Cho, K. J. Lee, & J.-H. Kim, Green synthesis of anisotropic silver nanoparticles and its potential cytotoxicity in human breast cancer cells (MCF-7). Journal of Industrial and Engineering Chemistry, 19 (2013) 1600–1605. https://doi.org/10.1016/j.jiec.2013.01.029. S. Gurunathan & X.-F. Zhang, Combination of salinomycin and silver nanoparticles enhances apoptosis and autophagy in human ovarian cancer cells: An effective anticancer therapy. International Journal of Nanomedicine, 11 (2016) 3655–3675. https://doi.org/10.2147/IJN.S111279. R. Amooaghaie, M. R. Saeri, & M. Azizi, Synthesis, characterization and biocompatibility of silver nanoparticles synthesised from Nigella sativa leaf extract in comparison with chemical silver nanoparticles. Ecotoxicology and Environmental Safety, 120 (2015) 400–408. https://doi.org/10.1016/j.ecoenv.2015.06.025. N. Gogoi, P. J. Babu, C. Mahanta, & U. Bora, Green synthesis and characterization of silver nanoparticles using alcoholic flower extract of Nyctanthes arbortristis and in vitro investigation of their antibacterial and cytotoxic activities. Materials Science and (QJLQHHULQJ C, 46 (2015) 463–469. https://doi.org/10.1016/j.msec.2014.10.069. R. Geetha, T. Ashokkumar, S. Tamilselvan, K. Govindaraju, M. Sadiq, & G. Singaravelu, Green synthesis of gold nanoparticles and their anticancer activity. Cancer Nanotechnology, 4 (2013) 91–98. https://doi.org/10.1007/s12645-013-0040-9. P. Karuppaiya, E. Satheeshkumar, W.-T. Chao, L.-Y. Kao, E. C.-F. Chen, & H.-S. Tsay, Anti-metastatic activity of biologically synthesised gold nanoparticles on human fibrosarcoma cell line HT1080. Colloids and Surfaces % Biointerfaces, 110 (2013) 163–170. https://doi.org/10.1016/j.colsurfb.2013.04.037. T. Y. Suman, S. R. Radhika Rajasree, R. Ramkumar, C. Rajthilak, & P. Perumal, The Green synthesis of gold nanoparticles using an aqueous root extract of Morinda citrifolia L. Spectrochimica Acta Part $ Molecular and Biomolecular Spectroscopy, 118 (2014) 11– 16. https://doi.org/10.1016/j.saa.2013.08.066. K Satyavani, S Gurudeeban, T Ramanathan, T Balasubramanian Biomedical potential of silver nanoparticles synthesised from calli cells of Citrullus colocynthis (L.) Schrad. Journal of Nanobiotechnology, 9 (2011) 43. https://doi.org/10.1186/1477-3155-9-43.

Metal Nanoparticles for Cancer Therapy

37.

38.

39.

40.

41.

42.

43.

44.

81

M. Valodkar, R. N. Jadeja, M. C. Thounaojam, R. V. Devkar, & S. Thakore, Biocompatible synthesis of peptide capped copper nanoparticles and their biological effect on tumour cells. Materials Chemistry and Physics, 128 (2011) 83–89. https://doi.org/10.1016/j.matchemphys.2011.02.039. S. Kamble, B. Utage, P. Mogle, R. Kamble, S. Hese, B. Dawane, & R. Gacche, Evaluation of curcumin capped copper nanoparticles as possible inhibitors of human breast cancer cells and angiogenesis: A comparative study with native curcumin. AAPS PharmSciTech, 17 (2016) 1030–1041. https://doi.org/10.1208/s12249-015-0435-5. S. Ghosh, R. Nitnavare, A. Dewle, G. B. Tomar, R. Chippalkatti, P. More, R. Kitture, S. Kale, J. Bellare, & B. A. Chopade, Novel platinum-palladium bimetallic nanoparticles synthesised by Dioscorea bulbifera: Anticancer and antioxidant activities. International Journal of Nanomedicine, 10 (2015) 7477–90. https://doi.org/10.2147/IJN.S91579. C. Yang, M. Wang, J. Zhou, & Q. Chi, Bio-synthesis of peppermint leaf extract polyphenols capped nano-platinum and their in vitro cytotoxicity towards colon cancer cell lines (HCT 116). Materials Science and (QJLQHHULQJ C, 77 (2017) 1012–1016. https://doi.org/10.1016/j.msec.2017.04.020. M. Jeyaraj, G. Sathishkumar, G. Sivanandhan, D. MubarakAli, M. Rajesh, R. Arun, G. Kapildev, M. Manickavasagam, N. Thajuddin, K. Premkumar, & A. Ganapathi, Biogenic silver nanoparticles for cancer treatment: An experimental report. Colloids and Surfaces % Biointerfaces, 106 (2013) 86–92. https://doi.org/10.1016/j.colsurfb.2013.01.027. R. Sukirtha, K. Manasa Priyanka, J. J. Antony, S. Kamalakkannan, R. Thangam, P. Gunasekaran, M. Krishnan, & S. Achiraman, Cytotoxic effect of green synthesised silver nanoparticles using Melia azedarach against in vitro HeLa cell lines and lymphoma mice model. Process Biochemistry, 47 (2011) 273–279. https://doi.org/10.1016/j.procbio.2011.11.003. J. J. Antony, M. Ali, A. Sithika, T. A. Joseph, U. Suriyakalaa, A. Sankarganesh, D. Siva, S. Kalaiselvi, & S. Achiraman, In vivo antitumour activity of biosynthesized silver nanoparticles using Ficus religiosa as a nanofactory in DAL induced mice model. Colloids and Surfaces % Biointerfaces, 108 (2013) 185–190. https://doi.org/10.1016/j.colsurfb.2013.02.041. M. AlSalhi, S. Devanesan, A. Alfuraydi, R. Vishnubalaji, M. A. Munusamy, K. Murugan, M. Nicoletti, & G. Benelli, Green synthesis

82

45.

46. 47.

48. 49. 50. 51.

52. 53.

54.

Chapter 3

of silver nanoparticles using Pimpinella anisum seeds: Antimicrobial activity and cytotoxicity on human neonatal skin stromal cells and colon cancer cells. International Journal of Nanomedicine, 11 (2016) 4439–4449. https://doi.org/10.2147/IJN.S113193. K. Anand, R. M. Gengan, A. Phulukdaree, & A. Chuturgoon, Agroforestry waste moringa oleifera petals mediated green synthesis of gold nanoparticles and their anticancer and catalytic activity. Journal of Industrial and Engineering Chemistry, 21 (2015) 1105– 1111. https://doi.org/10.1016/j.jiec.2014.05.021. S. M. El-Sonbaty, Fungus-mediated synthesis of silver nanoparticles and evaluation of antitumour activity. Cancer nanotechnology, 4 (2013) 73–79. https://doi.org/10.1007/s12645-013-0038-3. V. Borse, U. Chand Banerjee, A. Kaler, & M. Tech student, Microbial synthesis of platinum nanoparticles and evaluation of their anticancer activity synthesis and characterization of nanoparticles view project microbial synthesis of platinum nanoparticles and evaluation of their anticancer activity. International Journal of Emerging Trends in Electrical and Electronics, 11 (2015) 2320– 9569. https://doi.org/10.13140/RG.2.1.3132.5283. R. Bhattacharya & P. Mukherjee, Biological properties of ‘naked’ metal nanoparticles. Advanced Drug Delivery Reviews, 60 (2008) 1289–1306. https://doi.org/10.1016/j.addr.2008.03.013. K. McNamara & S. A. M Tofail, Nanoparticles in biomedical applications. Advances in 3K\VLFV X, 2 (2017) 54–88. https://doi.org/10.1080/23746149.2016.1254570. J. Peng & X. Liang, Progress in research on gold nanoparticles in cancer management. Medicine, 98 (2019) e15311. https://doi.org/10.1097/MD.0000000000015311. H. Sharma, P. K. Mishra, S. Talegaonkar, & B. Vaidya, Metal nanoparticles: A theranostic nanotool against cancer. Drug Discovery Today, 20 (2015) 1143–1151. https://doi.org/10.1016/j.drudis.2015.05.009. K. T. Nguyen, Targeted nanoparticles for cancer therapy: Promises and challenges. Journal of Nanomedicine & Nanotechnology, 02 (2011). https://doi.org/10.4172/2157-7439.1000103e. T. Sun, Y. S. Zhang, B. Pang, D. C. Hyun, M. Yang, & Y. Xia, Engineered nanoparticles for drug delivery in cancer therapy. Angewandte Chemie - International Edition, 53 (2014) 12320– 12364. https://doi.org/10.1002/anie.201403036. K. K. Harish, V. Nagasamy, B. Himangshu, & K. Anuttam, Metallic nanoparticle: A review. Biomed J Sci &Tech Res, 4 (2018).

Metal Nanoparticles for Cancer Therapy

55.

56.

57.

58.

59.

60.

61.

62.

63.

83

https://doi.org/10.26717/BJSTR.2018.04.001011. M. A. Safwat, G. M. Soliman, D. Sayed, & M. A. Attia, Gold nanoparticles enhance 5-fluorouracil anticancer efficacy against colorectal cancer cells. International Journal of Pharmaceutics, 513 (2016) 648–658. https://doi.org/10.1016/j.ijpharm.2016.09.076. N. Ngernyuang, W. Seubwai, S. Daduang, P. Boonsiri, T. Limpaiboon, & J. Daduang, Targeted delivery of 5-fluorouracil to cholangiocarcinoma cells using folic acid as a targeting agent. Materials Science and Engineering C, 60 (2016) 411–415. https://doi.org/10.1016/j.msec.2015.11.062. M. A. Safwat, G. M. Soliman, D. Sayed, & M. A. Attia, Fluorouracilloaded gold nanoparticles for the treatment of skin cancer: Development, in vitro characterization, and in vivo evaluation in a mouse skin cancer xenograft model. Molecular Pharmaceutics, 15 (2018) 2194–2205. https://doi.org/10.1021/acs.molpharmaceut.8b00047. P. Manivasagan, S. Bharathiraja, N. Q. Bui, B. Jang, Y. O. Oh, I. G. Lim, & J. Oh, Doxorubicin-loaded fucoidan capped gold nanoparticles for drug delivery and photoacoustic imaging. International Journal of Biological Macromolecules, 91 (2016) 578–588. https://doi.org/10.1016/j.ijbiomac.2016.06.007. L. C. Tsai, H. Y. Hsieh, K. Y. Lu, S. Y. Wang, & F. L. Mi, EGCG/gelatin-doxorubicin gold nanoparticles enhance therapeutic efficacy of doxorubicin for prostate cancer treatment. Nanomedicine, 11 (2016) 9–30. https://doi.org/10.2217/nnm.15.183. V. Ramalingam, K. Varunkumar, V. Ravikumar, & R. Rajaram, Target delivery of doxorubicin tethered with PVP stabilized gold nanoparticles for effective treatment of lung cancer. Scientific Reports, 8 (2018). https://doi.org/10.1038/s41598-018-22172-5. S. Borker & V. Pokharkar, Engineering of pectin-capped gold nanoparticles for delivery of doxorubicin to hepatocarcinoma cells: An insight into mechanism of cellular uptake. Artificial Cells, Nanomedicine, and Biotechnology, 46 (2018) 826–835. https://doi.org/10.1080/21691401.2018.1470525. A. Hekmat, A. A. Saboury, & A. Divsalar, The effects of silver nanoparticles and doxorubicin combination on DNA structure and its antiproliferative effect against T47D and MCF7 cell lines. Journal of Biomedical Nanotechnology, 8 (2012) 968–982. https://doi.org/10.1166/jbn.2012.1451. Y. Li, M. Guo, Z. Lin, M. Zhao, M. Xiao, C. Wang, T. Xu, T. Chen, & B. Zhu, Polyethylenimine-functionalized silver nanoparticle-

84

64.

65.

66.

67.

68.

69.

70.

71.

72.

Chapter 3

based co-delivery of paclitaxel to induce HepG2 cell apoptosis. International Journal of Nanomedicine, 11 (2016) 6693–6702. https://doi.org/10.2147/IJN.S122666. Y. Y. Gao, H. Chen, Y. Y. Zhou, L. T. Wang, Y. Hou, X. H. Xia, & Y. Ding, Intraorgan targeting of gold conjugates for precise liver cancer treatment. ACS Applied Materials and Interfaces, 9 (2017) 31458–31468. https://doi.org/10.1021/acsami.7b08969. Y. Sun, Q. Wang, J. Chen, L. Liu, L. Ding, M. Shen, J. Li, B. Han, & Y. Duan, Temperature-sensitive gold nanoparticle-coated pluronic-PLL nanoparticles for drug delivery and chemophotothermal therapy. Theranostics, 7 (2017) 4424–4444. https://doi.org/10.7150/thno.18832. L. T. Curtis, C. G. England, M. Wu, J. Lowengrub, & H. B. Frieboes, An interdisciplinary computational/experimental approach to evaluate drug-loaded gold nanoparticle tumour cytotoxicity. Nanomedicine, 11 (2016) 197–216. https://doi.org/10.2217/nnm.15.195. P. Manivasagan, S. Bharathiraja, N. Q. Bui, I. G. Lim, & J. Oh, Paclitaxel-loaded chitosan oligosaccharide-stabilized gold nanoparticles as novel agents for drug delivery and photoacoustic imaging of cancer cells. International Journal of Pharmaceutics, 511 (2016) 367–379. https://doi.org/10.1016/j.ijpharm.2016.07.025. F. Li, X. Zhou, H. Zhou, J. Jia, L. Li, S. Zhai, & B. Yan, Reducing both Pgp overexpression and drug Efflux with anticancer goldpaclitaxel nanoconjugates. PLOS ONE, 11 (2016) e0160042. https://doi.org/10.1371/journal.pone.0160042. Y. Jeong, S. T. Kim, Y. Jiang, B. Duncan, C. S. Kim, K. Saha, Y.-C. Yeh, B. Yan, R. Tang, S. Hou, C. Kim, M.-H. Park, & V. M. Rotello, Nanoparticle–dendrimer hybrid nanocapsules for therapeutic delivery. Nanomedicine, 11 (2016) 1571–1578. https://doi.org/10.2217/nnm-2016-0034. N. Zhang, H. Chen, A. Y. Liu, J. J. Shen, V. Shah, C. Zhang, J. Hong, & Y. Ding, Gold conjugate-based liposomes with hybrid cluster bomb structure for liver cancer therapy. Biomaterials, 74 (2016) 280–291. https://doi.org/10.1016/j.biomaterials.2015.10.004. C. Ong, J. Z. Z. Lim, C.-T. Ng, J. J. Li, L.-Y. L. Yung, & B.-H. Bay, Silver nanoparticles in cancer: Therapeutic efficacy and toxicity. Current Medicinal Chemistry, 20 (2013) 772–781. https://doi.org/10.2174/092986713805076685. L. Curtis, C. England, M. Wu, J. Lowengrub, & H. Frieboes, An interdisciplinary computational/ experimental approach to evaluate

Metal Nanoparticles for Cancer Therapy

73.

74.

75.

76.

77.

78.

79.

80.

85

drug-loaded gold nanoparticle tumour cytotoxicity. (2016). https://doi.org/10.2217/nnm.15.195. S. Dhar, F. X. Gu, R. Langer, O. C. Farokhzad, & S. J. Lippard, Targeted delivery of cisplatin to prostate cancer cells by aptamer functionalized Pt(IV) prodrug-PLGA-PEG nanoparticles. Proceedings of the National Academy of Sciences, 105 (2008) 17356–17361. https://doi.org/10.1073/pnas.0809154105. M. Kutwin, E. Sawosz, S. Jaworski, M. Hinzmann, M. Wierzbicki, A. Hotowy, M. Grodzik, A. Winnicka, & A. Chwalibog, Investigation of platinum nanoparticle properties against U87 glioblastoma multiforme. Archives of Medical Science, 6 (2017) 1322–1334. https://doi.org/10.5114/aoms.2016.58925. M. M. Kemp, A. Kumar, S. Mousa, E. Dyskin, M. Yalcin, P. Ajayan, R. J. Linhardt, & S. A. Mousa, Gold and silver nanoparticles conjugated with heparin derivative possess anti-angiogenesis properties. IOP PUBLISHING NANOTECHNOLOGY Nanotechnology, 20 (2009) 455104–455111. https://doi.org/10.1088/0957-4484/20/45/455104. D. Bartczak, O. L. Muskens, T. Sanchez-Elsner, A. G. Kanaras, & T. M. Millar, Manipulation of in vitro angiogenesis using peptidecoated gold nanoparticles. ACS Nano, 7 (2013) 5628–5636. https://doi.org/10.1021/nn402111z. J. Baharara, F. Namvar, T. Ramezani, N. Hosseini, & R. Mohamad, Green synthesis of silver nanoparticles using achillea biebersteinii flower extract and its antiangiogenic properties in the rat aortic ring model. Molecules, 19 (2014) 4624–4634. https://doi.org/10.3390/molecules19044624. J. Baharara, F. Namvar, M. Mousavi, T. Ramezani, & R. Mohamad, Anti-angiogenesis effect of biogenic silver nanoparticles synthesised using saliva officinalis on chick chorioalantoic membrane (CAM). Molecules, 19 (2014) 13498–13508. https://doi.org/10.3390/molecules190913498. D. B. Chithrani, S. Jelveh, F. Jalali, M. van Prooijen, C. Allen, R. G. Bristow, R. P. Hill, & D. A. Jaffray, Gold nanoparticles as radiation sensitizers in cancer therapy. Radiation Research, 173 (2010) 719– 28. https://doi.org/10.1667/RR1984.1. J. F. Hainfeld, D. N. Slatkin, & H. M. Smilowitz, The use of gold nanoparticles to enhance radiotherapy in mice. Physics in Medicine and Biology, 49 (2004) N309-15. https://doi.org/10.1088/00319155/49/18/n03.

86

81.

82.

83.

84.

85.

86.

87. 88.

89.

Chapter 3

R. Xu, J. Ma, X. Sun, Z. Chen, X. Jiang, Z. Guo, L. Huang, Y. Li, M. Wang, C. Wang, J. Liu, X. Fan, J. Gu, X. Chen, Y. Zhang, & N. Gu, Ag nanoparticles sensitize IR-induced killing of cancer cells. Cell Research, 19 (2009) 1031–1034. https://doi.org/10.1038/cr.2009.89. R. Lu, D. Yang, D. Cui, Z. Wang, & L. Guo, Egg white-mediated green synthesis of silver nanoparticles with excellent biocompatibility and enhanced radiation effects on cancer cells. International Journal of Nanomedicine, 7 (2012) 2101–2107. https://doi.org/10.2147/IJN.S29762. T. Kong, J. Zeng, X. Wang, X. Yang, J. Yang, S. McQuarrie, A. McEwan, W. Roa, J. Chen, & J. Z. Xing, Enhancement of radiation cytotoxicity in breast-cancer cells by localized attachment of gold nanoparticles. Small, 4 (2008) 1537–1543. https://doi.org/10.1002/smll.200700794. J. F. Hainfeld, H. M. Smilowitz, M. J. O, F. Avraham Dilmanian, & D. N. Slatkin, Gold nanoparticle imaging and radiotherapy of brain tumours in mice. J Hainfeld Nanomedicine, 8 (2013) 1601–1609. https://doi.org/10.2217/NNM.12.165. D. B. Chithrani, S. Jelveh, F. Jalali, M. van Prooijen, C. Allen, R. G. Bristow, R. P. Hill, & D. A. Jaffray, Gold Nanoparticles as radiation sensitizers in cancer therapy. Radiation Research, 173 (2010) 719. https://doi.org/10.1667/RR1984.1. Muhammad Afiq KA, R. AbRashida, R. MatLazim, N.Dollah, K. Abdul Razak, W.N.Rahmana, Evaluation of radiosensitization effects by platinum nanodendrites for 6 MV photon beam radiotherapy. Radiation Physics and Chemistry, 150 (2018) 40–45. https://doi.org/10.1016/j.radphyschem.2018.04.018. S. K. Lee, M. S. Han, S. Asokan, & C.-H. Tung, Effective gene silencing by multilayered siRNA-coated gold nanoparticles. Small, 7 (2011) 364–370. https://doi.org/10.1002/smll.201001314. S. Guo, Y. Huang, Q. Jiang, Y. Sun, L. Deng, Z. Liang, Q. Du, J. Xing, Y. Zhao, P. C. Wang, A. Dong, & X.-J. Liang, Enhanced gene delivery and siRNA silencing by gold nanoparticles coated with charge-reversal polyelectrolyte ACS Nano, 4 (2010) 5505–5511. https://doi.org/10.1021/nn101638u. J. Shen, H.-C. Kim, C. Mu, E. Gentile, J. Mai, J. Wolfram, L. Ji, M. Ferrari, Z. Mao, & H. Shen, Multifunctional gold nanorods for siRNA gene silencing and photothermal therapy. Advanced Healthcare Materials, 3 (2014) 1629–1637. https://doi.org/10.1002/adhm.201400103.

Metal Nanoparticles for Cancer Therapy

90.

91.

92.

93.

94.

95. 96.

97. 98. 99.

87

T. B. Huff, L. Tong, Y. Zhao, M. N. Hansen, J.-X. Cheng, & A. Wei, Hyperthermic effects of gold nanorods on tumour cells. Nanomedicine, 2 (2007) 125–132. https://doi.org/10.2217/17435889.2.1.125. L. Liu, F. Ni, J. Zhang, X. Jiang, X. Lu, Z. Guo, & R. Xu, Silver nanocrystals sensitize magnetic-nanoparticle-mediated thermoinduced killing of cancer cells. Acta Biochimica et Biophysica Sinica, 43 (2011) 316–323. https://doi.org/10.1093/abbs/gmr015. E. A. Thompson, E. Graham, C. M. Macneill, M. Young, G. Donati, E. M. Wailes, B. T. Jones, & N. H. Levi-Polyachenko, Differential response of MCF7, MDA-MB-231, and MCF 10A cells to hyperthermia, silver nanoparticles and silver nanoparticle-induced photothermal therapy. International Journal of Hyperthermia, 30 (2014) 312–323. https://doi.org/10.3109/02656736.2014.936051. Jannathul Firdhouse M & Lalitha P, Apoptotic efficacy of biogenic silver nanoparticles on human breast cancer MCF-7 cell lines. Progress in Biomaterials, 4 (2015) 113–121. https://doi.org/10.1007/s40204-015-0042-2. H.-C. Huang, K. Rege, & J. J. Heys, Spatiotemporal temperature distribution and cancer cell death in response to extracellular hyperthermia induced by gold nanorods. ACS Nano, 4 (2010) 2892– 2900. https://doi.org/10.1021/nn901884d. A. Sharma, A. K. Goyal, & G. Rath, Recent advances in metal nanoparticles in cancer therapy. Journal of Drug Targeting, 26 (2018) 617–632. https://doi.org/10.1080/1061186X.2017.1400553. B. .OĊERZVNL J. Depciuch, M. 3DUOLĔVND-Wojtan, & J. Baran, Applications of noble metal-based nanoparticles in medicine. International Journal of Molecular Sciences, 19 (2018). https://doi.org/10.3390/ijms19124031. S. Roussakow, The history of hyperthermia rise and decline. Conference Papers in Medicine, 428027 (2013) 1–40. https://doi.org/10.1155/2013/428027. N. Tran & T. J. Webster, Magnetic nanoparticles: Biomedical applications and challenges. Journal of Materials Chemistry, 20 (2010) 8760–8767. https://doi.org/10.1039/c0jm00994f. A. El-Ansary & S. Al-Daihan, On the toxicity of therapeutically used nanoparticles: An overview. Journal of Toxicology, 754810 (2009). https://doi.org/10.1155/2009/754810.

CHAPTER 4 NANOTECHNOLOGY IN GASTROENTEROLOGY RAJAN RAJABALAYA*1 AND SHEBA RANI DAVID2 1

PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2 School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA *Corresponding author: Dr Rajan Rajabalaya Senior Assistant Professor PAPRSB Institute of Health Sciences Universiti Brunei Darussalam Jalan Tungku Link BE1410 Bandar Seri Begawan Brunei Darussalam Email: [email protected] Phone: + 6732460922

Abstract The gastrointestinal tract (GIT) is regarded as an ideal site for drug delivery due to its high patient adherence. Further, the oral dosage form can be taken easily and is often less expensive than other forms. Despite these benefits, the GIT presents several barriers that protect the body from other foreign substances, including the drug. Because of these barriers, many drugs with low stability, solubility and bioavailability in GIT often require high doses, resulting in unwanted side effects. Nanotechnology applications for drug delivery have garnered huge popularity. The nanoparticles can overcome the limitations of conventional formulations by protecting the drug and controlling its release into the target site. Moreover, the nanoparticles for drug delivery are more advantageous than conventional formulations because they can enhance the drug solubility and stability, increase epithelium permeability, improve drug bioavailability, and allow tissue targeting with fewer side effects. Besides their use in drug delivery, the

Nanotechnology in Gastroenterology

89

unique physiochemical properties of nanoparticles also facilitate the detection, prevention and treatment of diseases. This chapter aims to integrate the potential applications of nanotechnology in gastroenterology, particularly in drug delivery, imaging, tissue engineering and theranostics.

Abbreviations 5ASA 5-FU AGNP AR Bor Bor/Cs/ChsFA CD CRC CTS-Pasp CysCS/PMLA DSS EC FDA GIT GNP IBD ICP-MS IL KPV MSN Nab NIR NLR NP NS NSR PCDC PCL PEG-PEI PLGA PMLA

5-Amino salicylic acid 5-Fluororacil GNP-bearing amoxicillin Aspect ratio Bortezomib Bortezomib-chitosan-chondroitin sulphate-conjugated folic acid nanoparticles Crohn’s diseases Colorectal cancer Chitosan-polyaspartic acid Cysteine-chitosan/PMLA nanoparticles Dextran sodium sulphate Ethyl cellulose Food and Drug Administration Gastrointestinal tract Gliadin nanoparticles Inflammatory bowel disease Inductively coupled plasma mass spectrometry Interleukin Tripeptide Lys-Pro-Val Mesoporous silica nanoparticles Nanoparticle albumin-bound Near-infrared fluorescent InAs (ZnS) Nano long rod Nanoparticles Nanosphere Nano short rod Pressure-controlled colon delivery capsule Poly(caprolactone), Polyethylene glycol-polyethyleneimine Poly-(lactic-co-glycolic acid) Poly (ȕ-L- Malic acid)

90

QDs ROS SiNP TNB TNF-Į UC

Chapter 4

Quantum dots Reactive oxygen species Silica nanoparticles Trinitrobenzene sulfonic acid, Tumour necrosis factor-alfa Ulcerative colitis

1. Introduction Gastroenterology refers to the study of normal function and diseases of the gastrointestinal tract (GIT), ranging from the oesophagus to the rectum, including the gall bladder, liver and pancreas [1]. The GIT is often vulnerable to infection. Such diseases may include inflammatory bowel disease, gastric cancer, colorectal cancer and infection by Helicobacter pylori (H pylori). Many options, such as oral, topical, parenteral and inhalation, are available for drug delivery. Oral drug administration is the most widely accepted route for drug delivery because of its high patient adherence, cost-effectiveness and easy administration. However, the oral drug delivery system presents various challenges mainly due to an inconsistent absorption rate of the drug and the GIT and, thus, poor drug bioavailability, commonly associated with dissolution, permeability and solubility factors [2]. Moreover, the oral drug delivery’s efficiency is also limited because of GIT barriers. Fortunately, the emergence of nanotechnology has somehow offered solutions to these limitations. Nanotechnology is described as ‘the design, characterisation, production and application of structures, devices and systems by controlling shape and size at the nanoscale (non-inflamed tissue>healthy tissue). The enhanced permeability and high numbers of immune cells were believed to explain increased particles’ uptake inside the colitis tissue. Moreover, the Me5ASA–SiNP could selectively adhere to the inflamed tissue and forming a drug reservoir at that site, decreasing the drug distribution to the healthy tissue. Additionally, after reaching the inflamed tissues, the nanoparticles release the drug, increasing the therapeutic efficacy.

Nanotechnology in Gastroenterology

101

A similar in vivo experiment aimed to verify the ability of nanoparticles for drug delivery and evaluate the therapeutic efficacy of anti-inflammatory tripeptide Lys-Pro-Val (KPV)-loaded nanoparticles (NP-KPV) on dextran sodium sulphate (DSS)-induced colitis mouse model. The NP-KPV was encapsulated into a polysaccharide gel comprising alginate and chitosan to target the inflamed colon. The colitis-induced mice with NP-KPV were protected against inflammatory and histologic parameters than mice with DSS. This result illustrates that the nanoparticles can target the antiinflammatory agents to the specific inflammation sites and allow the dose reduction compared to free KPV but with the same therapeutic efficacy [23]. Further, a study of 5-Amino salicylic acid-loaded poly(‫ܭ‬-caprolactone) nanoparticles on colitis mice showed the same results [32].

5.3 Helicobacter pylori infection Helicobacter Pylori (H. Pylori) was first discovered by Australian scientists Barry Marshall and Robin Warren in 1982 [33]. H. Pylori is classified as a gram-negative bacterium with a spiral on the surface of the gastric epithelium. It affects roughly 50% of humans worldwide and is commonly associated with GIT infections such as gastritis, gastric cancer, and peptic and duodenal ulcers. H. pylori is considered virulent because it has four to six flagella, enhancing their mobility to penetrate the mucous gel barrier and infect the epithelial cells. The invasion of H. Pylori to the epithelium cells is necessary for an active infection. The survival of H. pylori in the acidic pH of the gastric environment is facilitated by their ability to produce phospholipase, urease, cytotoxin and adhesin [34]. For example, the urease can neutralise the acidic environment by hydrolysing the urea in gastric juice into carbon dioxide and ammonia, favouring the survival of H. pylori. These mechanisms created by H. pylori could cause mucosal injury in the gastroduodenal artery through direct mucosal damage, alteration in host inflammatory responses, hypergastrinemia and elevated acid secretion [29]. During infection, the H. Pylori secrete urease to neutralise the acidic region of the stomach. When using their flagella, H. Pylori moves towards the gastric epithelial cells, followed by the adherence of the H. Pylori to the carbohydrates of the mucosa and epithelial cells by the adhesin-like protein, causing colonisation and persistent infection [35]. Finally, the H. Pylori secretes phospholipase, urease and cytotoxin, damaging the host tissue. In response to the infection, the epithelium layer releases chemokines to initiate innate immunity and activation of neutrophils, leading to clinical diseases, such as gastritis and ulcers. Although the immune responses are induced, H. Pylori can resist the local immune responses by decreasing the recognition

102

Chapter 4

Figure 4-1: Helicobacter pylori creates neutralised buffer zone by producing urease in the acidic zone, whereas bacterial mucinous erodes the mucus layer. Nanoparticles. Created with BioRender.com

Nanotechnology in Gastroenterology

103

of the immune sensor and altering the antigen uptake and other genetic mechanisms [34]. The treatment plan used as first-line therapy in treating H. Pylori is a triple therapy. This plan includes a proton pump inhibitor and two antibiotics, generally clarithromycin, amoxicillin and metronidazole twice a day for 7– 14 days [34–36]. However, this triple therapy is limited by several factors in clinical practices. Recently, sequential, non-bismuth quadruple, hybrid and concomitant therapy have shown promising results against H. Pylori (6,7). Although these therapies have shown a high eradication rate against H. pylori, they are still unsatisfactory [36]. The factors contributing to the therapeutic inefficacy are poor patient compliance, antibiotic instability at gastric pH, antibiotic resistance and the short gastric residence time [34]. Several efforts have been directed to design a drug delivery system that can overcome problems faced by conventional treatments. Therefore, by considering the pathophysiology of H. Pylori and difficulties presented in clinical practices, the oral targeted drug delivery that can increase gastric retention at the reaction site is an ideal strategy. Such efforts are made to deliver the antimicrobials directly to the infected sites, adhere antibiotics to the mucosal layer and release antibiotics continuously into the gastric environment [37]. The nanoparticles have a large specific surface, allowing for high interactions with biological surfaces thus making them ideal for mucoadhesive purposes. Umamaheshwari et al. [37] evaluated the effectiveness of amoxicillin-loaded mucoadhesive gliadin nanoparticles (AGNP) in eradicating H. Pylori. The AGNP was entrapped with rhodamine isothiocyanate, and their gastric mucoadhesive property was examined. Their study reported that after several hours of treatment, a high percentage of AGNP was still present in the stomach, proving the gastric residence time of AGNP was prolonged because of mucoadhesion. Further, the findings determined that the AGNP was more efficient and required a lower dose to eradicate H. Pylori over the amoxicillin-free solution completely. The mucoadhesive properties of gliadin nanoparticles were attributed to several interactions, such as hydrogen bonding, van der Waals force and mechanical penetration [38]. Moreover, the hydrophobicity and solubility characteristics of gliadin enabled the formulation of nanoparticles to protect the loaded drug and control its release [37]. Another in vivo study encapsulated clarithromycin into ethyl cellulose nanoparticles for H. Pylori clearance [39]. The encapsulated clarithromycin-loaded ethyl cellulose nanoparticles demonstrated a significant eradication of H. Pylori in the stomach than free clarithromycin-treated mice. The efficient eradication by encapsulated

104

Chapter 4

clarithromycin was attributed to its adherence along the gastric mucosa, helping localise the drug and allowing for the accumulation of drug concentrations. Additionally, a study was performed using cysteine-conjugated chitosan/PMLA multifunctional nanoparticles, encapsulating amoxicillin for eradicating H. pylori [40]. The amoxicillin-Cys-CS/PMLA nanoparticles are sensitive to pH changes and have a mucoadhesive profile to release amoxicillin into the targeted site. The results demonstrated that the growth of H. pylori was inhibited to a greater extent using amoxicillin-CysCS/PMLA nanoparticles compared to the unmodified amoxicillin. Therefore, these studies have shown that the nanoparticles have proven to be a promising approach for eradicating H. pylori mainly through their mucoadhesive properties. Thus, when the drug is encapsulated into a mucoadhesive nanoparticle, the mucoadhesion properties will prolong the drug’s release across the gastric mucosa, allowing the drug’s accumulation at the infection sites for effective therapy.

5.4 Gastric cancer Gastric cancer is defined as a malignancy of the stomach lining. It is classified as one of the cancer-causing deaths globally [41]. Advanced gastric carcinomas are divided as follows: polypoid growth (type I), fungating growth (type II), ulcerating growth (type III), and diffusely infiltrating growth (type IV) or termed as ‘linitis plastica’ in the signet ring cell carcinoma, according to Borrmann’s classification. T The gastric cancer stages are 1a, 1b, 2, 3 and 4. Various factors can lead to the occurrence of gastric cancer. Figure 4-2 presents the staging and types of gastric cancer. The factors could be attributed to infection, environmental or host-related factors, such as age, gender and diet [41–42]. Despite the high prevalence of gastric cancer in the Asian region, its treatment and diagnosis are considerably limited. The only curative therapy known for gastric cancer is surgery. Perioperative, adjuvant chemotherapy and radiation therapy can only enhance outcomes [43]. For example, chemotherapy combined with cisplatin and fluoropyrimidine in patients with Stage IV gastric cancer is a commonly used treatment. Further, trastuzumab is commonly used in patients with human epidermal growth factor receptor 2-positive gastric cancer [43]. However, these therapies often affect the normal tissues [41]. Thus, an intervention that can address this challenge is desired for the efficient treatment of gastric cancer.

Nanotechnology in Gastroenterology

105

Figure 4-2: Gastric cancer staging consisting of 1a until 4 and Bormann classification of gastric cancer with four types. Created with BioRender.com

106

Chapter 4

The unique physicochemical properties of nanomedicine have made it suitable for gastric cancer treatment. Compared to conventional treatment, the nanoparticles could decrease side effects and increase the treatment efficacy via a site-targeting mechanism [41]. In gastric cancer, the nanoparticles directly transport the anticancer drugs (e.g., Fluorouracil, taxane, antimetabolite, etc.) to the targeted site without being distributed to the non-cancerous sites [44]. The greater specificity of targeting cancer cells by nanosized materials is facilitated by two fundamental processes: active and passive targeting. Active targeting refers to any ligand favouring attachment towards malignant relative to nonmalignant cells. In contrast, passive targeting involves enhanced permeability and retention effect to elevate the nanoparticles concentration in the tumour [44]. Some studies demonstrated the use of nanoparticles in treating gastric cancer. Zhang et al. [45] experimented with chitosan-polyaspartic acid (CTS-Pasp) nanoparticles. The CTS-Pasp nanoparticles were utilised to deliver 5-fluorouracil (5-FU) to the cancer cells. The results revealed that CTS-Pasp-5FU has a greater tumour inhibition rate than free 5-FU. Moreover, Wu et al. [46] used PEG-modified polyethyleneimine (PEGmodified PEI). They evaluated its capability of delivering siRNA to suppress the activity of CD44v6 in gastric cancer cells in vitro in a gene therapy study. Conversely, Zhang et al. [47] investigated the microtubuleinhibitory effect of nanoparticle albumin-bound paclitaxel (Nab-paclitaxel) compared to free oxaliplatin and epirubicin using human gastric cancer cell lines: AGS, NCI-N87 and SNU16. Based on the findings, gastric cancer cell proliferation inhibition was more significant in Nab-paclitaxel than in oxaliplatin and epirubicin.

5.5 Colorectal cancer Colorectal cancer (CRC) involves the development of cancer in the colon and/or the rectum. It has one of the highest cancer prevalence and is responsible for thousands of deaths worldwide [48]. The CRC often progresses slowly and usually begins from non-cancerous polyp progression in the colon or rectum. The existing treatments for CRC are surgery, radiotherapy, chemotherapy (e.g., 5-FU, oxaliplatin and capecitabine) and targeted therapy (e.g., bevacizumab). However, the current therapy possesses some limitations, especially its toxic effects on the non-cancerous cells and the development of drug resistance [48]. The use of nanoparticles for drug delivery has been investigated to deliver anticancer drugs to CRC cells while decreasing the drug distribution into non-cancerous cells.

Nanotechnology in Gastroenterology

107

Shashank Tummala and M. N. Satish Kumar used 5-FU chitosan nanoparticles to localise the drug in the colon and reduce the toxicity to noncancerous cells [49]. They concluded that the 5-FU chitosan nanoparticles could target the drug to colorectal cancer without affecting the healthy cells. Further, the chitosan nanoparticles release the drug via a sustained released mechanism. Safwat et al. [50] used gold nanoparticles to increase the efficacy of 5-FU against colorectal cancer. Their findings showed that the 5-FU/ glutathione-gold nanoparticles have higher efficacy in treating CRC than free 5-FU. Another study used bortezomib-chitosan-chondroitin sulphate-conjugated folic acid nanoparticles (Bor/Cs/Chs-FA) to selectively deliver the nanoparticles into the colorectal cells [51]. The expression of the folate receptors was higher in colorectal cancer cells lines (HT-29 and HCT116). Thus, the antitumour study using HCT116 showed that the Bor/Cs/Chs-FA nanoparticles exhibit better antitumour effects and fewer toxic effects than those without folate-targeting ligands.

Carrier

Silica nanoparticles (SiNP)

Tripeptide KVP (Lys-ProVal) nanoparticles

PCL nanoparticles

Gliadin nanoparticles (GNP)

Ethyl cellulose (EC) nanoparticles

Drug

5-Amino salicylic acid (5ASA)

-

5-Amino salicylic acid (5ASA)

Amoxicillin

Clarithromycin

H. pylori infection

H. pylori infection

IBD

IBD

IBD

Disease

In vivo, mice

In vivo, Gerbil

In vivo, Mice

In vivo, mice

Invitro/ vivo In vivo, mice

Methods

Chapter 4

The H. Pylori eradication on infected mice using clarithromycin-loaded EC nanoparticles was assessed. The free drug was used as a control.

The gastric mucoadhesive and the clearance of H. pylori using Rhodamine isothiocyanateentrapped GNP were evaluated.

NPs were prepared by double emulsion/solvent evaporation and loaded with KVP. The NPKPV were encapsulated into a polysaccharide gel polymer consisting of alginate and chitosan. The mice with TNB-induced were orally given either 5ASA, 5ASA-NP or saline (control).

The SiNP targeting ability and therapeutic efficiency were assessed in colitis mice.

Table 4-4: Drug delivery in nano-gastroenterology

108

Orally administered 5ASAPCL-NP slowed drug release and allowed to lower 5ASA dose. The H. Pylori eradication by AGNP was more effective than amoxicillin because of the longer duration of GI residence due to mucoadhesive properties. High H. pylori eradication using the clarithromycinloaded EC nanoparticles compared to the free drugs.

The SiNP has potential in the targeted drug delivery as it was found to be siteselective, and it allows the reduction in drug dose. The mice with dextran sodium sulphate (DSS) and NP-KPV were shown to be more protected from inflammation than mice with DSS alone.

Outcomes

Preclinical study

Preclinical study

Preclinical study

Preclinical study

Preclinical study

Status

2014 [39]

2004 [37]

2007 [32]

2010 [23]

Year [Ref] 2008 [31]

PolymersiRNA nanoparticle

-

Nanoparticle albumin-bound (Nab)

Chitosanpolyaspartic acid (CTSPasp) nanoparticles

5fluorouracil

Paclitaxel

Cysteinechitosan/PMLA nanoparticles (CysCS/PMLA)

Amoxicillin

Gastric cancer

Gastric cancer

Gastric cancer

H. pylori infection

In vitro

In Vitro

In vivo, mice

In vitro, Human stomach carcinoma cell line (AGS)

Human gastric cancer cell lines AGS, NCI-N87 and SNU16 were used to test the microtubuleinhibitory effect of Nabpaclitaxel.

The SCG-7901 gastric carcinoma cell line mass was injected into male BABL/c nude mice. The mice were divided into CTS-Pasp-5FU, 5-FU, CTS-Pasp and normal saline groups. SiRNA targeting CD44v6 was incorporated in NPs. SGC-7901 cell culture was treated with polymer-siRNA NPs.

The Cys-CS/PMLA nanoparticles were used to encapsulate amoxicillin. The impact of the nanoparticles on AGS was examined by MTT assay.

PEG-PEI showed great promise for altering gene expression in gastric cancer treatment because it has high gene transfection efficiency and low cytotoxicity. The Nab-paclitaxel showed a high antiproliferative potency with a lowest effective dosage

The Cys-CS/PMLA nanoparticles are sensitive to pH and possess mucoadhesive properties. In comparison to unmodified amoxicillin, the amoxicillin-loaded Cys-CS/PMLA nanoparticles effectively inhibit the H. pylori growth The CTS-Pasp-5FU NPs have a higher inhibition rate of tumour growth than FU.

Nanotechnology in Gastroenterology

Phase II clinical trial

Preclinical study

Preclinical study

Preclinical study

2013 [47]

2010 [46]

2008 [45]

2018 [40]

109

Gold nanoparticles

Chitosan polymeric nanoparticles

Folate-targeting self-assembled nanoparticles using polymer chitosan and chondroitin sulphate

Colorectal cancer

Colorectal cancer

Colorectal cancer

In vitro

In vitro

In vivo, mice

The efficacy of 5-FU/GSH-gold nanoparticles was assessed by flow cytometry in colorectal cancerous tissues.

High folate receptors expression in colorectal cell lines HT-29 and HCT116 thus, easy interactions with folate-targeting nanoparticles. The HCT116 was injected into mice, and the Bor antitumour efficacy with and without folate targeting was evaluated. In vitro drug release study was done by the USP dissolution apparatus

Chapter 4

The 5-FU chitosan nanoparticles are protected from the gastric environment and localise the drug in the colon area with a sustained-release drug profile. Prevent cytotoxicity to noncancerous cells. The 5-FU/GSH- gold nanoparticles have a higher anticancer effect than free 5-FU. The gold nanoparticles enhance the efficacy of 5-FU

Bor/Cs/Chs-FA inhibits the tumour progression with lower toxicity to noncancerous cells and tissues than control, free drug and formulation without folate targeting.

Preclinical

Preclinical

Preclinical

2016 [50]

2015 [49]

2018 [51]

PCL-poly(caprolactone), TNB-trinitrobenzene sulfonic acid, PMLA- Poly (Malic acid), PEG-PEI- polyethylene glycol-polyethyleneimine, RefReference

5-Fluororacil

5- fluorouracil

Bortezomib (Bor)

110

Nanotechnology in Gastroenterology

111

6. Gastrointestinal imaging using nanoparticles The use of nanoparticles is limited to drug delivery and extends as a diagnostic tool. Many experiments aimed to determine the possible use of nanoparticles in molecular imaging to diagnose and treat cancer [52]. This focus is mainly due to their enhanced cellular uptake and their ability to target specific sites through surface modifications [7]. Moreover, their surface chemistry, magnetic properties, tunable absorption and emission properties show the opportunity of using nanoparticles as probes for disease detection [52]. Several nanoparticles, such as liposomes, micelles, metallic nanoparticles, quantum dots and dendrimers, are currently investigated for molecular investigation. Alvarez et al. [53] examined the use of an anti-mesothelin antibodyconjugated quantum rod in diagnosing and treating mesothelin-overexpressing oesophageal adenocarcinomas. The immunohistochemical analysis found that mesothelin’s expression is limited to oesophageal adenocarcinoma and related metastasis. This finding verifies the potential application of mesothelin as a marker for the diagnosis and therapy of oesophageal adenocarcinoma. Another study used near-infrared fluorescent (NIR) InAs(ZnS) quantum dots to evaluate the biodistribution and clearance of the nanoparticles by varying the length of short-chain polyethylene glycol (PEGs) [54]. The tissue- or organ-specific biodistribution and clearance heavily depended on the PEG chain length. Additionally, Kumar et al. [55] performed an in vivo biodistribution and clearance experiment using multimodal organically modified silica (ORMOSIL) nanoparticles. The ORMOSIL nanoparticles were conjugated with NIR and radiolabelled with Iodine (124I). The biodistribution of nanoparticles was determined using optical fluorescence imaging and measuring the radioactivity from the collected organs. The finding revealed a high accumulation of ORMOSIL nanoparticles in the liver, spleen and stomach relative to the kidney, heart and lung. In contrast, the clearance study indicated hepatobiliary excretion. The quantum dots (QDs) are defined as semiconductor crystals of nanometre dimension with a distinctive conductive characteristic determined by their size. T The QD has unique photophysical properties such as high fluorescent quantum field, photochemical robustness and resistance to a photobleaching effect, making them excellent for biomedical applications [56]. Moreover, a previous study successfully used cadmium telluride QDs as a proton flux sensor for detecting the influenza virus [57]. Another study used semiconductor QDs in the fluorometric assay to detect Escherichia coli (E. coli). The QDs maintained high fluorescence intensities longer than typical dye bleaches, allowing the rapid and precise identification of E. coli [58].

Chapter 4

Applications

Oesophageal cancer

Imaging

Imaging

Nanoparticle

Anti-mesothelin antibodyconjugated quantum rod

Near-infrared fluorescent InAs (ZnS) (NIR) quantum dots

Multimodal organically modified silica (ORMOSIL)

In vivo

In vivo

In vitro / in vivo In vitro

A successive IV injection of PEGylated NIR QD was administered to rats, and their distribution and clearance were evaluated using NIR fluorescence imaging. The ORMOSIL nanoparticles were conjugated with NIR and radio-labelled with 124I. The biodistribution of nanoparticles was determined using optical fluorescence imaging and measuring the radioactivity from the collected organs.

The presence of mesothelin was investigated by examining the biopsy specimens using immunohistochemical.

Methods

Table 4-5: Gastrointestinal imaging using nanoparticles

112

The finding revealed a high accumulation of ORMOSIL nanoparticles in the liver, spleen and stomach compared to the kidney, heart and lung. The clearance study, on the other hand, indicated hepatobiliary excretion.

The chain length affects the nanoparticles’ distribution and clearance.

The mesothelin was found to be restricted in oesophageal adenocarcinoma and related metastasis, and it has potential use in the diagnosis and treatment of mesothelinoverexpressing oesophageal adenocarcinomas.

Results

[55]

[54]

[53]

Reference

Nanotechnology in Gastroenterology

113

7. Nanotoxicity Although nanoparticles have gained much interest, their possible toxicity and long-term adverse effects are still under-explored. Nanotoxicology refers to a study relating to the toxicity of the nanoparticles, and it is crucial to understand their potential toxicity before using them for various applications. Several routes of the administration of nanoparticles to humans are also available, including inhalation, injection, oral ingestion and transdermal routes [59]. Therefore, lungs, skin, the central nervous system and GIT are potentially affected organs arising from nanoparticle toxicity [60]. The physicochemical characteristics of nanoparticles, such as size, specific surface area, AR and surface properties, significantly correlate with nanotoxicity [60]. The particles’ size and surface area are known to be the main factors leading to toxicity. The small-sized nanoparticles exhibit higher toxicity because they have a high surface-area-to-volume ratio, enabling greater chemical and biological interactions [59]. Moreover, as mentioned in Section 5.1, the cell uptake and distribution are considerably influenced by the size of the nanoparticles, with the smallest having higher cell uptake and distribution, increasing the likelihood of nanoparticles accumulation and toxicity (the smaller the particles, the higher the uptake and distribution). Conversely, the surface properties of the nanoparticles may facilitate the adsorption of ions and biomolecules, which can affect the cellular responses elicited and induce toxicity. However, the underlying mechanism of nanotoxicity is related to the generation of reactive oxygen species (ROS), which can cause oxidative stress in the tissue [61]. The nanoparticles significantly associated with toxicity are metal nanoparticles, such as high doses of copper, zinc, silver and gold [62, 66]. For example, silver nanoparticles are frequently used as antibacterial agents [67]. However, they have been reported to cause toxicities. The permeation of silver nanoparticles to the cell membrane often triggers the release of its ions, Ag+, potentially causing cytotoxicity [67–69]. Some silver dosedependent animal toxicity-related side effects are weight loss, liver damage and altered blood levels [70–71]. The toxicity mechanisms related to the use of silver nanoparticles include DNA toxicity, cytokine induction and oxidative stress [72]. ROS generation during cell uptake can result in oxidative stress and genotoxic effects. Additionally, the high amount of ROS can cause cell death by apoptosis or necrosis [67, 72]. Further, the release of potentially toxic ions depends on the pH because any pH variations along the GIT can affect the dissolution and the mechanism of ingested nanoparticles (69). At acidic pH, the dissolution and denudation of nanoparticles are fast. Thus, the pH shift may influence the local toxicity

Chapter 4

114

and the penetration of nanoparticles into GIT barriers (70). However, most studies on nanoparticles toxicity have focused on their biodistribution and systemic effects. In contrast, the potential effects on the gastrointestinal tract are still under-studied. Table 4-6 summarises nanotoxicity studies by various researchers, such as Khanna P, Ong C, Bay B, Baeg [59]. Table 4-6: Overview of the different nanoparticles used in nanomedicine and their toxicity [59]. Nanoparticles

Applications

Polymeric nanoparticles Polysaccharide Drug delivery [73] chitosan nanoparticles Poly- (lactic-co- drug delivery for cancer glycolic acid) therapy [74] (PLGA) Inorganic nanoparticles Ceramic Cancer drug delivery nanoparticles [75] Metallic nanoparticles Supermagnetic Magnetic resonance iron oxide imaging contrast nanoparticles enhancement, immunoassays and cancer drug carrier systems [77–78] Gold shell Imaging and therapeutics nanoparticles [80]

Toxicity

Report – Nil Report – Nil

Oxidative stress/cytotoxic activity in the lungs, liver, heart and brain [76] Oxidative stress and disturbance in iron homeostasis [79]

Hepatic and splenic toxicity [81]

8. Approved nanomedicines in gastroenterology Some nanoparticles for gastroenterology have been approved by the Food and Drug Administration (FDA). Table 4-7 shows this list of approved nanoparticles.

Nanotechnology in Gastroenterology

115

Table 4-7: The list of approved nanomedicine in gastroenterology [82] Name Cimzia®/ certolizumab pegol (UCB)

Onivyde® (Merrimack)

Abraxane®/ABI007 (Celgene)

Material Nanoparticle Indication description advantage(s) (s) Polymer nanoparticles PEGylated Improved Crohn's antibody circulation time disease fragment and greater (Certolistability in vivo zumab) Liposome formulations Liposomal Enhanced delivery Pancreatic Irinotecan to tumour site; Cancer reduce systemic toxicity arising from side effects Protein nanoparticles Albuminincrease solubility; Pancreatic bound increase delivery cancer paclitaxel to tumour nano-particles Inorganic and metallic nanoparticles SPION coated Superparamagnetic Imaging with silicone character agent

Year approved 2008

2015

2013

GastroMARK™; 2001 Lumirem® (AMAG pharmaceuticals) SPION- Superparamagnetic iron oxide nanoparticles 6RXUFH Bobo D, Robinson KJ, Islam J, Thurecht KJ, Corrie SR. Nanoparticle-Based 0HGLFLQHV A Review of FDA-Approved Materials and Clinical Trials to Date. Pharm Res.   –

9. Conclusion The rising interest in nanotechnology has led to a new field called nanomedicine. Nanomedicine offers various potential uses in gastroenterology, such as drug delivery, imaging, tissue engineering and theranostics. This literature review offers insight into the possible interactions of nanoparticles along the gastrointestinal tract, gastrointestinal disease processes, and nanotechnology in disease prevention, treatment and diagnosis. However, future research is needed to interpret this information into toxicity studies and extend the research into human studies. In the future, the successful developments of nanotechnology will give rise to significant advantages, especially in every aspect of human life. Medicine and tissue engineering are among the leading areas that will benefit from nanotechnology advancements.

116

Chapter 4

References 1. 2.

3.

4.

5. 6.

7.

8.

9. 10.

R. Riasat & N. Guangjun, Effects of nanoparticles on gastrointestinal disorders and therapy. Journal of Clinical Toxicology, 6 (2016). https://doi.org/10.4172/2161-0495.1000313. N. Ü. Okur, P. I. Siafaka, & E. H. Gökçe, Challenges in oral drug delivery and applications of lipid nanoparticles as potent oral drug carriers for managing cardiovascular risk factors. Current Pharmaceutical Biotechnology, 22 (2021) 892–905. https://doi.org/10.2174/1389201021666200804155535. A. Albanese, P. S. Tang, & W. C. W. Chan, The effect of nanoparticle size, shape, and surface chemistry on biological systems. Annual Review of Biomedical Engineering, 14 (2012) 1–16. https://doi.org/10.1146/annurev-bioeng-071811-150124. H. Laroui, D. S. Wilson, G. Dalmasso, K. Salaita, N. Murthy, S. V Sitaraman, & D. Merlin, Nanomedicine in GI. American Journal of Physiology. Gastrointestinal and Liver Physiology, 300 (2011) G371-83. https://doi.org/10.1152/ajpgi.00466.2010. W. H. De Jong & P. J. a Borm, Drug delivery and nanoparticles: applications and hazards. International Journal of Nanomedicine, 3 (2008) 133–149. https://doi.org/10.2147/IJN.S596. S. K. Lai, Y.-Y. Wang, & J. Hanes, Mucus-penetrating nanoparticles for drug and gene delivery to mucosal tissues. Advanced Drug Delivery Reviews, 61 (2009) 158–171. https://doi.org/10.1016/j.addr.2008.11.002. G. Brakmane, M. Winslet, & A. M. Seifalian, Systematic review: The applications of nanotechnology in gastroenterology. Alimentary Pharmacology and Therapeutics, 36 (2012) 213–221. https://doi.org/10.1111/j.1365-2036.2012.05179.x. E. M. Pridgen, F. Alexis, & O. C. Farokhzad, Polymeric nanoparticle drug delivery technologies for oral delivery applications. Expert Opinion on Drug Delivery, 12 (2015) 1459–1473. https://doi.org/10.1517/17425247.2015.1018175. Nanotechnology and nanomaterials in the treatment of lifethreatening diseases (Elsevier, 2014). https://doi.org/10.1016/C2013-0-13374-0. C. H. Lin, C. H. Chen, Z. C. Lin, & J. Y. Fang, Recent advances in oral delivery of drugs and bioactive natural products using solid lipid nanoparticles as the carriers. Journal of Food and Drug Analysis, 25 (2017) 219–234. https://doi.org/10.1016/j.jfda.2017.02.001.

Nanotechnology in Gastroenterology

11.

12.

13. 14. 15.

16.

17.

18.

19.

20.

117

L. M. Ensign, R. Cone, & J. Hanes, Oral drug delivery with polymeric nanoparticles: The gastrointestinal mucuc barrier. NIH Public Access, 64 (2013) 557–570. https://doi.org/10.1016/j.addr.2011.12.009.Oral. M. Liu, J. Zhang, W. Shan, & Y. Huang, Developments of mucus penetrating nanoparticles. Asian Journal of Pharmaceutical Sciences, 10 (2014) 275–282. https://doi.org/10.1016/j.ajps.2014.12.007. and S. S. K. Behera A.L, Patil S.V., Nanosizing of drugs: A promising approach for drug delivery. Der Pharmacia Sinica, 1 (2010) 20–28. S. A. A. Rizvi & A. M. Saleh, Applications of nanoparticle systems in drug delivery technology. Saudi Pharmaceutical Journal, 26 (2017) 64–70. https://doi.org/10.1016/j.jsps.2017.10.012. W. H. De Jong, W. I. Hagens, P. Krystek, M. C. Burger, A. J. A. M. Sips, & R. E. Geertsma, Particle size-dependent organ distribution of gold nanoparticles after intravenous administration. Biomaterials, 29 (2008) 1912–1919. https://doi.org/10.1016/j.biomaterials.2007.12.037. A. Banerjee, J. Qi, R. Gogoi, J. Wong, & S. Mitragotri, Role of nanoparticle size, shape and surface chemistry in oral drug delivery. Journal of Controlled Release, 238 (2016) 176–185. https://doi.org/10.1016/j.jconrel.2016.07.051. S. E. A. Gratton, P. A. Ropp, P. D. Pohlhaus, J. C. Luft, V. J. Madden, M. E. Napier, & J. M. DeSimone, The effect of particle design on cellular internalization pathways. Proceedings of the National Academy of Sciences, 105 (2008) 11613–11618. https://doi.org/10.1073/pnas.0801763105. Y. Zhao, Y. Wang, F. Ran, Y. Cui, C. Liu, Q. Zhao, Y. Gao, D. Wang, & S. Wang, A comparison between sphere and rod nanoparticles regarding their in vivo biological behaviour and pharmacokinetics. Scientific Reports, 7 (2017) 1–11. https://doi.org/10.1038/s41598-017-03834-2. L. Fuentes, S. Lebenkoff, K. White, C. Gerdts, K. Hopkins, J. E. Potter, D. Grossman, P. E. Project, & R. Sciences, The impact of aspect ratio on the biodistribution and tumour homing of rigid softmatter nanorods. 93 (2016) 292–297. https://doi.org/10.1016/j.contraception.2015.12.017.Women. N. J. S. M Manikandan, K Kannan, R Manavalan, Design of nanoparticles for colon target drug delivery – A review. Research Journal of Pharmaceutical, Biological and Chemical, 2 (2011) 128–

118

21. 22.

23.

24.

25.

26. 27.

28.

29.

30.

Chapter 4

139. N. Kumar & R. Kumar, Nanotechnology and nanomaterials in the treatment of life-threatening diseases (2014). https://doi.org/10.1016/B978-0-323-26433-4.00003-8. M. Manikandan, K. Kannan, R. Manavalan, & N. J. Sundresh, Dsign of nanoparticles for colon target drug delivery – A review. Research Journal of Pharmaceutical, Biological and Chemical Sciences REVIEW, 2 (2011) 128–139. H. Laroui, G. Dalmasso, H. T. T. Nguyen, Y. Yan, S. V. Sitaraman, & D. Merlin, Drug-loaded nanoparticles targeted to the colon with polysaccharide hydrogel reduce colitis in a mouse model. Gastroenterology, 138 (2010) 843-853.e2. https://doi.org/10.1053/j.gastro.2009.11.003. a Dowling, R. Clift, N. Grobert, D. Hutton, R. Oliver, O. O’neill, J. Pethica, N. Pidgeon, J. Porritt, J. Ryan, & Et Al., Nanoscience and nanotechnologies: Opportunities and uncertainties. London The Royal Society The Royal Academy of Engineering Report, 46 (2004) 618–618. https://doi.org/10.1007/s00234-004-1255-6. I. J. T. Sci, S. Singh, V. K. Pandey, R. Prakash Tewari, & V. Agarwal, Nanoparticle based drug delivery system: Advantages and applications. Indian Journal of Science and Technology, 4 (2011) 25–29. https://doi.org/10.17485/ijst/2011/v4i3/29960. M. Gagliardi, Novel biodegradable nanocarriers for enhanced drug delivery. Therapeutic Delivery, 7 (2016) 809–826. https://doi.org/10.4155/tde-2016-0051. A. Viscido, A. Capannolo, G. Latella, R. Caprilli, & G. Frieri, Nanotechnology in the treatment of inflammatory bowel diseases. Journal of Crohn’s and Colitis, 8 (2014) 903–918. https://doi.org/10.1016/j.crohns.2014.02.024. B. R. R. de Mattos, M. P. G. Garcia, J. B. Nogueira, L. N. Paiatto, C. G. Albuquerque, C. L. Souza, L. G. R. Fernandes, W. M. da S. C. Tamashiro, & P. U. Simioni, Inflammatory bowel disease: An overview of immune mechanisms and biological treatments. Mediators of Inflammation, 2015 (2015) 1–11. https://doi.org/10.1155/2015/493012. Marie A. Chisholm-Burns, Terry L. Schwinghammer, Barbara G. Wells, Patrick M. Malone, Jill M. Kolesar, Joseph T. DiPiro, Pharmacotherapy principles and practice, 4th Edition (2013) McGraw Hill, New York USA, S. Zhang, R. Langer, & G. Traverso, Nanoparticulate drug delivery systems targeting inflammation for treatment of inflammatory bowel

Nanotechnology in Gastroenterology

31.

32.

33. 34.

35. 36.

37.

38.

39.

40.

119

disease. Nano Today, 16 (2017) 82–96. https://doi.org/10.1016/j.nantod.2017.08.006. B. Moulari, D. Pertuit, Y. Pellequer, & A. Lamprecht, The targeting of surface modified silica nanoparticles to inflamed tissue in experimental colitis. Biomaterials, 29 (2008) 4554–4560. https://doi.org/10.1016/j.biomaterials.2008.08.009. D. Pertuit, B. Moulari, T. Betz, A. Nadaradjane, D. Neumann, L. Ismaïli, B. Refouvelet, Y. Pellequer, & A. Lamprecht, 5-amino salicylic acid bound nanoparticles for the therapy of inflammatory bowel disease. Journal of Controlled Release, 123 (2007) 211–218. https://doi.org/10.1016/j.jconrel.2007.08.008. G. Watts, Nobel prize is awarded to doctors who discovered H pylori. BMJ (Clinical research ed.), 331 (2005) 795. https://doi.org/10.1136/bmj.331.7520.795. S. Zhao, J.-B. Zhang, B. Wang, Y. Lv, Guo-JunLv, & X.-J. Ma, Gastroretentive drug delivery systems for the treatment of Helicobacter pylori. World Journal of Gastroenterology, 20 (2014) 9321–9329. https://doi.org/10.3748/wjg.v20.i28.9321. C. Y. Kao, B. S. Sheu, & J. J. Wu, Helicobacter pylori infection: An overview of bacterial virulence factors and pathogenesis. Biomedical Journal, 39 (2016) 14–23. https://doi.org/10.1016/j.bj.2015.06.002. D. Lopes, C. Nunes, M. C. L. Martins, B. Sarmento, & S. Reis, Eradication of helicobacter pylori: Past, present and future. Journal of Controlled Release, 189 (2014) 169–186. https://doi.org/10.1016/j.jconrel.2014.06.020. R. B. Umamaheshwari, S. Ramteke, & N. K. Jain, Anti-Helicobacter pylori effect of mucoadhesive nanoparticles bearing amoxicillin in experimental gerbils model. AAPS PharmSciTech, 5 (2004) 60–68. https://doi.org/10.1208/pt050232. A. O. Elzoghby, M. M. Elgohary, & N. M. Kamel, Implications of protein- and peptide-based nanoparticles as potential vehicles for anticancer drugs. (2015) 169–221. https://doi.org/10.1016/bs.apcsb.2014.12.002. P. Pan-In, W. Banlunara, N. Chaichanawongsaroj, & S. Wanichwecharungruang, Ethyl cellulose nanoparticles: Clarithomycin encapsulation and eradication of H. pylori. Carbohydrate Polymers, 109 (2014) 22–27. https://doi.org/10.1016/j.carbpol.2014.03.025. M. Arif, Q. J. Dong, M. A. Raja, S. Zeenat, Z. Chi, & C. G. Liu, Development of novel pH-sensitive thiolated chitosan/PMLA nanoparticles for amoxicillin delivery to treat Helicobacter pylori.

120

41. 42.

43.

44. 45.

46.

47.

48.

49.

Chapter 4

Materials Science and Engineering C, 83 (2018) 17–24. https://doi.org/10.1016/j.msec.2017.08.038. S. A. Kuddus, Nanoparticles to deal with gastric cancer. Journal of Gastrointestinal Cancer and Stromal Tumours, 02 (2017) 1–5. https://doi.org/10.4172/2572-4126.1000112. B. Hu, N. El Hajj, S. Sittler, N. Lammert, R. Barnes, & A. MeloniEhrig, Gastric cancer: Classification, histology and application of molecular pathology. Journal of Gastrointestinal Oncology, 3 (2012) 251–261. https://doi.org/10.3978/j.issn.2078-6891.2012.021. M. Orditura, G. Galizia, V. Sforza, V. Gambardella, A. Fabozzi, M. M. Laterza, F. Andreozzi, J. Ventriglia, B. Savastano, A. Mabilia, E. Lieto, F. Ciardiello, & F. De Vita, Treatment of gastric cancer. World Journal of Gastroenterology, 20 (2014) 1635–1649. https://doi.org/10.3748/wjg.v20.i7.1635. W. H. Gmeiner & S. Ghosh, Nanotechnology for cancer treatment. Nanotechnology Reviews, 3 (2014). https://doi.org/10.1515/ntrev2013-0013. D. Y. Zhang, X. Z. Shen, J. Y. Wang, L. Dong, Y. L. Zheng, & L. L. Wu, Preparation of chitosan-polyaspartic acid-5-fluorouracil nanoparticles and its anti-carcinoma effect on tumour growth in nude mice. World Journal of Gastroenterology, 14 (2008) 3554–3562. https://doi.org/10.3748/wjg.14.3554. Y. Wu, W. Wang, Y. Chen, K. Huang, X. Shuai, Q. Chen, X. Li, & G. Lian, The investigation of polymer-siRNA nanoparticle for gene therapy of gastric cancer in vitro. International Journal of Nanomedicine, 5 (2010) 129–136. https://doi.org/10.2147/IJN.S8503. C. Zhang, N. Awasthi, M. A. Schwarz, S. Hinz, & R. E. Schwarz, Superior antitumour activity of nanoparticle albumin-bound paclitaxel in experimental gastric cancer. PLoS ONE, 8 (2013) 1–10. https://doi.org/10.1371/journal.pone.0058037. S. P. Chandran, S. B. Natarajan, S. Chandraseharan, & M. S. B. Mohd Shahimi, Nano drug delivery strategy of 5-fluorouracil for the treatment of colorectal cancer. Journal of Cancer Research and Practice, 4 (2017) 45–48. https://doi.org/10.1016/j.jcrpr.2017.02.002. A. P. Shashank Tummala, M.N. Satish Kumar, Formulation and characterization of 5-Fluorouracil enteric coated nanoparticles for sustained and localized release in treating colorectal cancer. Saudi Pharmaceutical Journal, 23 (2015) 308–314. https://doi.org/10.1016/j.jsps.2014.11.010.

Nanotechnology in Gastroenterology

50.

51.

52.

53.

54.

55.

56. 57.

58.

121

M. A. Safwat, G. M. Soliman, D. Sayed, & M. A. Attia, Gold nanoparticles enhance 5-fluorouracil anticancer efficacy against colorectal cancer cells. International Journal of Pharmaceutics, 513 (2016) 648–658. https://doi.org/10.1016/j.ijpharm.2016.09.076. Z. C. Soe, B. K. Poudel, H. T. Nguyen, R. K. Thapa, W. Ou, M. Gautam, K. Poudel, S. G. Jin, J.-H. Jeong, S. K. Ku, H.-G. Choi, C. S. Yong, & J. O. Kim, Folate-targeted nanostructured chitosan/chondroitin sulfate complex carriers for enhanced delivery of bortezomib to colorectal cancer cells. Asian Journal of Pharmaceutical Sciences, 14 (2019) 40–51. https://doi.org/10.1016/j.ajps.2018.09.004. S. K. Nune, P. Gunda, P. K. Thallapally, Y. Lin, M. Laird, & C. J. Berkland, Nanoparticles for biomedical imaging. Expert Opinion in Drug Delivery, 6 (2009) 1175–1194. https://doi.org/10.1517/17425240903229031.Nanoparticles. H. Alvarez, P. L. Rojas, K. T. Yong, H. Ding, G. Xu, P. N. Prasad, J. Wang, M. Canto, J. R. Eshleman, E. A. Montgomery, & A. Maitra, Mesothelin is a specific biomarker of invasive cancer in the Barrettassociated adenocarcinoma progression model: translational implications for diagnosis and therapy. 1DQRPHGLFLQH Nanotechnology, Biology, and Medicine, 4 (2008) 295–301. https://doi.org/10.1016/j.nano.2008.06.006. H. S. Choi, B. I. Ipe, P. Misra, J. H. Lee, M. G. Bawendi, & J. V Frangioni, Tissue- and organ-selective biodistribution of NIR fluorescent quantum dots. Nano Letters, 9 (2009) 2354–2359. https://doi.org/10.1021/nl900872r. R. Kumar, I. Roy, T. Ohulchanskky, & L. Vathy, In vivo biodistribution and clearance studies using multimodal organically modified silica nanoparticles. ACS nano, 4 (2010) 699–708. https://doi.org/10.1021/nn901146y.In. V. G. Reshma & P. V. Mohanan, Quantum dots: Applications and safety consequences. Journal of Luminescence, 205 (2018) 287–298. https://doi.org/10.1016/j.jlumin.2018.09.015. Z. Yun, D. Zhengtao, Y. Jiachang, T. Fangqiong, & W. Qun, Using cadmium telluride quantum dots as a proton flux sensor and applying to detect H9 avian influenza virus. Analytical Biochemistry, 364 (2007) 122–127. https://doi.org/10.1016/j.ab.2007.02.031. M. A. Hahn, J. S. Tabb, & T. D. Krauss, Detection of single bacterial pathogens with semiconductor quantum dots. Analytical Chemistry, 77 (2005) 4861–4869. https://doi.org/10.1021/ac050641i.

122

59. 60. 61.

62.

63.

64.

65. 66.

67.

68.

Chapter 4

P. Khanna, C. Ong, B. Bay, & G. Baeg, Nanotoxicity: An interplay of oxidative stress, Inflammation and Cell Death. Nanomaterials, 5 (2015) 1163–1180. https://doi.org/10.3390/nano5031163. B. Viswanath & S. Kim, Influence of nanotoxicity on human health and environment: The alternative Ssrategies. Rev. Environ. Contam. Toxicol. (2016) 61–104. https://doi.org/10.1007/398_2016_12. P. P. Fu, Q. Xia, H. M. Hwang, P. C. Ray, & H. Yu, Mechanisms of nanotoxicity: Generation of reactive oxygen species. Journal of Food and Drug Analysis, 22 (2014) 64–75. https://doi.org/10.1016/j.jfda.2014.01.005. X. D. Zhang, H. Y. Wu, D. Wu, Y. Y. Wang, J. H. Chang, Z. Bin Zhai, A. M. Meng, P. X. Liu, L. A. Zhang, & F. Y. Fan, Toxicologic effects of gold nanoparticles in vivo by different administration routes. International Journal of Nanomedicine, 5 (2010) 771–781. https://doi.org/10.2147/IJN.S8428. T.-K. Hong, N. Tripathy, H.-J. Son, K.-T. Ha, H.-S. Jeong, & Y.-B. Hahn, A comprehensive in vitro and in vivo study of ZnO nanoparticles toxicity. Journal of Materials Chemistry B, 1 (2013) 2985. https://doi.org/10.1039/c3tb20251h. Z. Chen, H. Meng, G. Xing, C. Chen, Y. Zhao, G. Jia, T. Wang, H. Yuan, C. Ye, F. Zhao, Z. Chai, C. Zhu, X. Fang, B. Ma, & L. Wan, Acute toxicological effects of copper nanoparticles in vivo. Toxicology Letters, 163 (2006) 109–120. https://doi.org/10.1016/j.toxlet.2005.10.003. Y. P. Jia, B. Y. Ma, X. W. Wei, & Z. Y. Qian, The in vitro and in vivo toxicity of gold nanoparticles. Chinese Chemical Letters, 28 (2017) 691–702. https://doi.org/10.1016/j.cclet.2017.01.021. C. Lopez-Chaves, J. Soto-Alvaredo, M. Montes-Bayon, J. Bettmer, J. Llopis, & C. Sanchez-Gonzalez, Gold nanoparticles: Distribution, bioaccumulation and toxicity. In vitro and in vivo studies. 1DQRPHGLFLQH Nanotechnology, Biology, and Medicine, 14 (2018) 1–12. https://doi.org/10.1016/j.nano.2017.08.011. M. C. Stensberg, Q. Wei, E. S. McLamore, D. M. Porterfield, A. Wei, & M. S. Sepúlveda, Toxicological studies on silver nanoparticles: challenges and opportunities in assessment, monitoring and imaging. Nanomedicine, 6 (2011) 879–898. https://doi.org/10.2217/nnm.11.78. A. Barbasz, M. 2üZLHMD & M. Roman, Toxicity of silver nanoparticles towards tumoural human cell lines U-937 and HL-60. Colloids and Surfaces % Biointerfaces, 156 (2017) 397–404. https://doi.org/10.1016/j.colsurfb.2017.05.027.

Nanotechnology in Gastroenterology

69.

70.

71.

72.

73.

74.

75.

76. 77.

123

M. Van Der Zande, R. J. Vandebriel, E. Van Doren, E. Kramer, Z. Herrera Rivera, C. S. Serrano-Rojero, E. R. Gremmer, J. Mast, R. J. B. Peters, P. C. H. Hollman, P. J. M. Hendriksen, H. J. P. Marvin, A. A. C. M. Peijnenburg, & H. Bouwmeester, Distribution, elimination, and toxicity of silver nanoparticles and silver ions in rats after 28day oral exposure. ACS Nano, 6 (2012) 7427–7442. https://doi.org/10.1021/nn302649p. Y. S. Kim, M. Y. Song, J. D. Park, K. S. Song, H. R. Ryu, Y. H. Chung, H. K. Chang, J. H. Lee, K. H. Oh, B. J. Kelman, I. K. Hwang, & I. J. Yu, Subchronic oral toxicity of silver nanoparticles. Particle and Fibre Toxicology, 7 (2010) 20. https://doi.org/10.1186/17438977-7-20. N. Hadrup & H. R. Lam, Oral toxicity of silver ions, silver nanoparticles and colloidal silver - A review. Regulatory Toxicology and Pharmacology, 68 (2014) 1–7. https://doi.org/10.1016/j.yrtph.2013.11.002. Sylvie Gaillet & J. M. Rouanet, Silver nanoparticles: Their potential toxic effects after oral exposure and underlying mechanisms - A review. Food and Chemical Toxicology, 77 (2015) 58–63. https://doi.org/https://doi.org/10.1016/j.fct.2014.12.019. S. A. Agnihotri, N. N. Mallikarjuna, & T. M. Aminabhavi, Recent advances on chitosan-based micro- and nanoparticles in drug delivery. Journal of Controlled Release, 100 (2004) 5–28. https://doi.org/10.1016/j.jconrel.2004.08.010. C. Fornaguera, A. Dols-Perez, G. Calderó, M. J. García-Celma, J. Camarasa, & C. Solans, PLGA nanoparticles prepared by nanoemulsion templating using low-energy methods as efficient nanocarriers for drug delivery across the blood–brain barrier. Journal of Controlled Release, 211 (2015) 134–143. https://doi.org/10.1016/j.jconrel.2015.06.002. A. K. Cherian, A. C. Rana, & S. K. Jain, Self-assembled carbohydrate-stabilized ceramic nanoparticles for the parenteral delivery of insulin. Drug Development and Industrial Pharmacy, 26 (2000) 459–463. https://doi.org/10.1081/DDC-100101255. D. Cui, F. Tian, C. S. Ozkan, M. Wang, & H. Gao, Effect of single wall carbon nanotubes on human HEK293 cells. Toxicology Letters, 155 (2005) 73–85. https://doi.org/10.1016/j.toxlet.2004.08.015. A. K. Gupta & M. Gupta, Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials, 26 (2005) 3995–4021. https://doi.org/10.1016/j.biomaterials.2004.10.012.

124

78.

79.

80.

81.

82.

Chapter 4

M. Pöttler, A. Staicu, J. Zaloga, H. Unterweger, B. Weigel, E. Schreiber, S. Hofmann, I. Wiest, U. Jeschke, C. Alexiou, & C. Janko, Genotoxicity of superparamagnetic iron oxide nanoparticles in granulosa cells. International Journal of Molecular Sciences, 16 (2015) 26280–26290. https://doi.org/10.3390/ijms161125960. S. Laurent, A. A. Saei, S. Behzadi, A. Panahifar, & M. Mahmoudi, Superparamagnetic iron oxide nanoparticles for delivery of therapeutic agents: Opportunities and challenges. Expert Opinion on Drug Delivery, 11 (2014) 1449–1470. https://doi.org/10.1517/17425247.2014.924501. L. R. Hirsch, A. M. Gobin, A. R. Lowery, F. Tam, R. A. Drezek, N. J. Halas, & J. L. West, Metal nanoshells. Annals of Biomedical Engineering, 34 (2006) 15–22. https://doi.org/10.1007/s10439-0059001-8. A. M. Alkilany & C. J. Murphy, Toxicity and cellular uptake of gold nanoparticles: What we have learned so far? Journal of Nanoparticle Research, 12 (2010) 2313–2333. https://doi.org/10.1007/s11051010-9911-8. D. Bobo, K. J. Robinson, J. Islam, K. J. Thurecht, & S. R. Corrie, Nanoparticle-based medicines: A review of FDA-approved materials and clinical trials to date. Pharmaceutical Research, 33 (2016) 2373–2387. https://doi.org/10.1007/s11095-016-1958-5.

CHAPTER 5 NANOMEDICINES IN ORAL CANCER THERAPIES SHEBA RANI DAVID*1 AND RAJAN RAJABALAYA2 1

School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2

*Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract Oral cancer, also known as head or neck cancer, is a prevalent cancer type with remarkable morbidity. It accounts for two-thirds of cancer-related mortalities worldwide. Oral cancer is defined as cancer of the oral cavities and oropharynx. A total of 90% of the diagnosed oral cancers comprise oral squamous cell carcinoma (OSCC). Although several treatments are currently available for managing oral cancers, these treatments appear to have limitations. Despite that outcome, investigations on nanoplatforms have been emerging because they have introduced a promising strategy in oral cancer therapies. Their capability to deliver substances might be attributed to improving efficiency in delivering anticancer agents and reducing cytotoxic effects. This standard review chapter aims to summarise the current development of different nanoparticle-based medicines, such as polymers, gold nanoparticles, magnetic nanoparticles, dendrimers, nanoshells

126

Chapter 5

and plant-based approaches, for the treatment of oral cancer. Spectacular in vitro and in vivo results with minimised toxicity was observed in using nanoparticles to deliver cancer therapeutics. Enhancing the laser radiation with polymer and gold nanoparticles approaches was also noted in this chapter. Ultimately, the prospects of further investigation into nano-based medicine used in oral cancer will be discussed to bolster the use of nanobased medicine as one of the oral cancer treatment alternatives in clinical settings.

Abbreviations ((PDPN Ab)-AuNPDOX) 5-FU-Cur-NE 5-FUNE AuNPs Cat Cat-NPs CDDP Cur-NE Cur-NP CSL DOX DTX EGF EPR FGF hNOK HNPD HP NCCN NLS-AuNPs OSCC PAMAM PD PEG PEI PhCsNPs

PDPN antibody and DOX 5-fluorouracil-and-curcuminnanoemulsions 5-fluorouracil-nanoemulsions Gold nanoparticles Catechol Catechol-modified chitosan (also known as hyaluronic acid nanoparticles) Cisplatin Curcumin-nanoemulsions Curcumin-loaded nanoparticles Doxorubicin Docetaxel Epidermal growth factor Enhanced permeability and retention Fibroblast growth factors Human normal oral keratinocytes HN-1 mediated PEGylated doxorubicin Hematoporphyrin National Comprehensive Cancer Network Nuclear targeting gold nanospheres Oral Squamous Cell Carcinoma 5 polyamidoamine PEGylated doxorubicin Polyethylene glycol Polyethylenimine A plant-based nanomedicine named phloretin that is loaded in chitosan

Nanomedicines in Oral Cancer Therapies

PLA PLGA PLGA/NR7 RPTD/HP VEGF

127

Polylactic acid Poly-lactic-co-glycolic acid PLGA-PEG Cisplatin conjugated with NR7 peptide Encapsulated PEGylated doxorubicin with hematoporphyrin Vascular endothelial growth factor

1. Introduction The sixth most common cancer form, oral cancer, is responsible for about two-thirds of cancer-related deaths globally, exacerbating health and socioeconomic problems [1–2]. Genetic and epigenetic factors significantly affect emerging new health issues [3]. These factors have been strongly associated with the advancement of oral cancer. Oral cancer has developed immensely due to associated habits or epigenetic factors, such as smoking and alcohol assumption. In this perspective, smoking includes both tobacco and smokeless tobacco because smokeless tobacco contains nicotine, nitrosamines and carcinogens, just like tobacco smoking. In fact, smokeless tobacco is one of the significant risk factors contributing to the high prevalence of oral cancer [1]. Moreover, genetic factors, such as familial and genetic predispositions and ethnicity, play an essential role in the occurrence of oral cancer in an individual [3]. The pathology of oral cancer is not a straightforward process; it involves many alterations and mutations of genes and growth factors, such as the epidermal growth factor (EGF) and vascular endothelial growth factor (VEGF) [4]. Commonly, oral cancer involves cancer in the tongue, lips, the mucosae of the buccal and alveolar regions and the roof and floor of the mouth. Oral squamous cell carcinoma affects at least nine out of ten people with oral cancers, with lymphomas, melanomas, or sarcomas accounting for the remainder [2]. Additionally, the early detection of oral cancers and malignant disorders, which can undergo malignant transformation, is crucial for predicting the progress of oral cancer. Presently, oral cancer is diagnosed via biopsy to determine the histopathology of the cancer cells. It is considered the gold standard in diagnosing oral cancer [5].

1.1 Current oral cancer treatment and drawbacks Currently, oral cancer treatments comprise traditional treatments, such as chemotherapy, radiation therapy or surgery alone or in combinations [6].

128

Chapter 5

When the tumour is dissectible, surgery is usually considered the most conventional treatment accompanied by either postoperative radiation therapy or chemotherapy. The accompanied treatment is decided by several measures, such as tumour size and patient’s conditions [7]. However, these treatments seemed to have drawbacks, particularly chemotherapy. This is due to the uncontrollable effect of chemotherapy that manifests high toxicity because it kills non-specific fast-growing cells [6]. It is undeniable that chemotherapy’s side effects significantly impact the patients’ quality of life, potentially impairing their ability to perform their everyday tasks. The poor management of toxicity and adverse events of chemotherapy, such as nausea and vomiting, could also hinder the treatment’s effectiveness because of unnecessary dose reduction or postponement of treatment [8]. With that in mind, health care professionals have shown increased interest in developing individualised treatment regimens for oral cancer. The National Comprehensive Cancer Network (NCCN) was established in 1990 to standardise treatment guidelines for optimising the health outcomes of patients with oral cancer. The 1997 NCCN guidelines stated that a single treatment of either surgery or radiation therapy was recommended for an early stage of oral cancer. This regimen increases survival by up to three years. Increasing survival rates have also been shown in the NCCN guidelines that recommended surgery and chemoradiotherapy as an adjuvant in treating late-stage oral cancer [9]. Chemoradiotherapy, or chemoradiation, is the concurrent use of chemotherapy and radiotherapy. The chemotherapy agent enhances the sensitisation of tumours to radiotherapy, improving the spread of local tumours [10]. This treatment is considered superior to conventional chemotherapy, especially in inoperable patients. Although several toxicities are related to this chemoradiation, there is still scope for improvement. Because of the chemoradiation’s direct effect on the tumour without affecting many healthy cells, the implementation of nanomedicine might enhance the targeting effect of the chemoradiation agent, outweighing the toxicity problems [10–11].

1.2 Nanoparticles as a new approach in oral cancer treatment Nanomedicine is a combination of medical approaches using nanotechnology that refers to repairing, constructing and controlling the biology of humans using the tenets of nanotechnology [12]. It is a revolutionary branch of medicine that applies nanoparticles in drug delivery, especially chemotherapeutics. Nanomedicine was first introduced in 1950 with the emergence of polyvinyl pyrrolidone mescaline as a conjugate of the medicine and the polymer, followed by the discovery of liposomes in the

Nanomedicines in Oral Cancer Therapies

129

1960s. Most innovations of nanoparticles for the past years have been focusing on their application in therapeutics and are currently still under investigation [11]. The potential use of nanomedicine can be considered from two different perspectives: whether it can be utilised as a cure for a specific disease or can cause harm to patients [12]. In 1995, Doxil became the first nanotherapeutics anticancer drug discovered and approved by the Food and Drug Administration in the United States [13]. Since then, vigorous developments in nanomedicine have been made in various formulations. Notably, these efforts have improved over the past years [14]. Several advantages are available from the use of nanotechnology in medicine. These include the ability to compartmentalise several elements into one drug that can be customised to make the full use of therapeutic functions safely in their personalised ways and their ability to differentiate cancer from normal tissue. Moreover, nanomedicine tends to have electromagnetic properties that could aid therapeutic applications [13]. Although prominent nanotechnology applications are present in the medical field, no nanoparticles-based treatment has been approved for oral cancer [4].

1.3 Nanoparticles characterisation The unique characteristics of nanoparticles have proven beneficial in targeting the delivery method. Nanoparticles have a small size of fewer than 100 nanometres and are biodegradable and biocompatible, supporting their applications in chemotherapy and immunotherapy [10–11]. The nanoparticles’ different sizes and shapes affect their formulations and pharmacokinetics depending on how the shape complements the biologic sites to perform their respective functions [11]. Additionally, sizes and shapes are considered the most crucial parameters in determining the nanoparticles’ biodistribution and tumour penetration ability [15]. These parameters can be established through various methods, such as x-ray diffraction and spectroscopy [16]. Because the distribution of nanoparticles depends on their size and shape, it is essential to customise the nanoparticles to accumulate the anticancer agent in the targeted tissues [11,17]. From evaluating different sizes of gold nanoparticles (AuNPs) ranging from 15 nm to 200 nm, the smallest size was most highly accumulated in the organs. Still, the nanoparticles should cross the tumour vessels, with sizes ranging between 40 and 200 nm, but not pass through the normal vessel walls, with pores’ sizes ranging from 6 to 12 nm [15]. This action ensures that the drug only and specifically penetrates the tumour. Another purpose of customising the size of the nanoparticles is to

130

Chapter 5

ensure that the nanoparticles are quickly removed from the other nontargeted tissues to avoid high cytotoxicity because the size of nanoparticles plays an important role in clearing the nanoparticles from the body [11, 15, 17]. Usually, nanoparticles smaller than 10 nm are rapidly removed via the kidneys. In contrast, nanoparticles larger than 10 nm are removed in the liver with or without the aid of the mononuclear phagocyte system, depending on the shape of the nanoparticles [11]. Elongated nanoparticles are easily embodied when placed 90 degrees to the macrophage. Although smaller nanoparticles might benefit drug delivery and removal, smaller nanoparticles yield higher toxicity [15]. In contrast, sometimes, charges and coatings of nanoparticles also affect the biodistribution of nanoparticles in the tissues. It was discovered that positively charged nanoparticles have a lower diffusion coefficient and penetration ability onto the skin, whereas negatively charged nanoparticles show an opposite result. For instance, positively charged nanoparticles usually accumulate more in the lungs than in other tissues, resulting in effective lung cancer treatment [15]. Hence, it can be deduced that charges should be personalised depending on the expected accumulation area of the nanoparticles to optimise the cancer treatment. The coatings, which are considered crucial for protecting the nanodrug from the immune system, also play an important role in the mechanism of action. An example of coating used is polyethylene glycol (PEG), which can help extend the halflife of the nanoparticles. However, at the same time, the implementation of coating required some discussion to make it more favourable, especially concerning its solubility, density and hydrophobicity, because it will affect the biodistribution and elimination of the nanoparticles.

1.4 Classifications of nanodrug delivery Nanodrug delivery has been around for quite some time in the pharmaceutical industry. The development of nanodrug has progressed from its first generation back in the 1950s to the third generation of nanodrugs currently still under development. Four known nanodrug delivery systems are available [4]: passive, active, immune, and magnetic targeting. They can be used accordingly depending on the groups targeted [18]. The study of the pathophysiology of most cancers showed that the nanodrug accumulation in the tumour is the initial method recognised to be beneficial in targeting the tumour alone without affecting normal tissues. Hence, the first generation of nanodrugs’ invention was based on the passive targeting of drug delivery without ligands [17]. This is due to the enhanced permeability and retention (EPR) arising in the leaky intratumoral blood vessels that enhance the

Nanomedicines in Oral Cancer Therapies

131

passive targeting effect into the tumour via diffusion. This effect was proven to manage the primary tumour [4, 17] effectively. This method is beneficial since it improves the nanodrugs’ ability to sustain the distribution and retention time, allowing the drug to reach the tumour [18]. However, more studies found that the EPR effect is insufficient in targeting the tumour and reducing the side effects of the nanodrug [17]. Because of the first generation of nanodrugs drawbacks, further research was conducted to improve the targeting mode and the efficacy of the nanodrug. Consequently, a second-generation nanodrug was developed to target the tumour, known as smart nanocarriers, actively. They are responsive to external stimuli [4]. A ligand of high specificity is usually attached to the active targeting nanodrugs. This connection allows them to selectively and specifically bind to the target cell, reducing the binding of nanodrugs with healthy cells. The difference between active targeting and passive targeting nanodrugs is that active targeting nanodrugs can overcome barriers, such as anatomic and physiologic barriers, that hinder the drug from reaching the cancer cells. Nevertheless, the efficiency of the active targeting nanodrug depends on the nanodrug’s ability to prevent it from releasing and degrading prematurely in the body. The efficiency also depends on its density to avoid early detection by the mononuclear phagocytes [18]. The discovery of the passive and active targeting nanodrug delivery has opened the door to more advanced nanodrug delivery methods for specifically targeting the cancer cells according to their characteristics. Hence, it will lead to innovative immune and magnetic nanodrug delivery processes. The method of immune targeting nanomedicine is adopted from both the active and passive targeting drug delivery processes. It is done with the infusion of the tumour antigen to stimulate the production of antibodies to fight the tumour cells via active targeting or using external factors to apply the antitumour effect. In contrast, magnetic targeting is concerned with the use of magnetic nanoparticles, such as ferric oxide, that amplify the chemotherapy effect by their ability to locate the tumours. These nanodrug delivery systems intend to optimise the ligand functions and reduce the nanodrugs’ rapid opsonisation before serving their purposes [18].

132

Chapter 5

2. Different targeting sites for oral cancer in nanomedicine delivery The foundation of nanodrug delivery systems has been used to establish oral squamous cell carcinoma (OSCC) treatment and applied in four different targeted areas: the tumour vessels, intracellular fluid and extracellular matrix, and stromal cells dendritic cells. Each site plays a significant role in the growth of oral cancer cells. Therefore, the effect of nanodrug delivery needs to be targeted in those areas. The blood vessels and stromal cells are the important suppliers of nutrition and growth factors such as fibroblast growth factors (FGF) and vascular endothelial growth factors for the oral cancer cell to grow and metastasise. Therefore, it is a brilliant idea to target and supply that particular area with anticancer therapeutics to exert its antitumour effect directly on the tumour and disrupt the supplementation of growth factors to the tumour. In contrast, dense intracellular fluid and extracellular matrix allow the slow penetration of drugs, giving a good opportunity for nanodrug delivery to perform its function effectively in the targeted areas. The purpose of targeting oral tumour-associated dendritic cells is to trigger the immune system to release the antitumour response or produce the cytotoxic T cells to kill the oral tumour cells [18].

2.1 Nanotechnology avenues for oral cancer treatment In nanomedicine, nanoparticles are used in different methods of oral cancer treatment in different avenues. The nanodrugs are either encapsulated, dissolved or entrapped while behaving as a passive host to perform their functions in the delivery system, especially in oral cancer treatment [19]. Using nanoparticles in the treatment of oral cancer was introduced to overcome the drawbacks of conventional oral cancer strategies [6]. However, their use tends to be accompanied by different carrier variations that enhance the functions of the nanodrugs, especially when targeting cancer cells. The different nanotechnology approaches, such as polymers, AuNPs, magnetic nanoparticles, dendrimers, nanoshells, and natural products, such as chitosan and curcumin, were studied in vitro using different oral cancer cell lines in vivo (Table 5-1 and Table 5-2). Nanomedicine applications have gained prominence over the past few years. Some have already been approved by the FDA for use, whereas some are still undergoing clinical trials. The lists of nanomedicine applications are outlined in Table 5-3. Moreover, some nanomedicines are also patented. Data were extracted from clinicaltrials.gov. FDA approved and products under various stages of development have been listed. (Table 5-4).

SCC-25

CAL27

HSC-4

Polymers

Polymers

Polymers

HN6

OSC19 OSC20 HSC-3

Nanotechnology Avenues

Oral Cancer Cell lines

Doxorubicin

Cisplatin

Cisplatin

Active Ingredients

The cytometric profiles observed that the HNPD internalised the cancer cells from the CAL-27 and SCC-25 cell lines more than PD. From the MTT assays results, the IC50 concluded that the HNPD displays higher cytotoxicity than PD.

The PLGA-PEG Cisplatin shows more internalisation (more than 20%) than the free drug (less than 10%) but at the same time enhances its cytotoxicity effects as the dose increases, in which its cytotoxicity is superior over the free cisplatin. The free cisplatin has a more potent inhibitory effect over the NC-6004 cisplatin. Although, concerning toxicity, NC-6004 displays less toxicity towards the kidney than free cisplatin. This can be demonstrated by measuring the serum creatinine to examine the severity of kidney damage.

PLGA-PEG Cisplatin is formed into spherical micelles coated with the non-polar PLGA, and polar PEG as the outer layer ranges around 100 to 135 nm maximum in size. The cisplatin is polymerised using a hydrophilic polymeric micellar nanoparticle, NC6004, resulting in polymeric micelles with a diameter of approximately 30 nm. PEGylated doxorubicin is mediated with HN-1, developing HN-1 mediated PEGylated doxorubicin (HNPD) micelles with a diameter of 140 nm approximately.

Key Findings

Nanoparticles Characteristics

Table 5-1. List of in vitro studies for nanotechnology avenues for oral cancer treatment

Nanomedicines in Oral Cancer Therapies

[20]

[21]

[20]

References

133

Docetaxel

Doxorubicin

Polymer

Gold nanoparticles

SCC-9

HSC-3

Doxorubicin

Polymers

CAL-27

134

[23]

[24]

The PEGylated PLGA DTX observed that the encapsulated DTX releases the active drugs slowly and in a sustained manner and has a higher toxicity effect than the free DTX. Both forms of AuNPs, release-resistant DOX nanoparticles and DOX-releasing PHresponsive nanoparticles, showed increasing localisation in the nucleus of the HSC-3 cell lines. Still, no significant change is seen in healthy HaCat cells, but they have two times higher cellular uptake than the free DOX. Although, there are differences in the pattern of cell death. The flow cytometry revealed that the release-resistant DOX nanoparticles caused necrosis that also contributes to cytotoxicity, whereas the DOX-releasing PH-responsive nanoparticles showed that the cell death was caused by apoptosis.

The encapsulated PEGylated DOX with the photosensitizer hematoporphyrin (HP) is a spherical nanoparticle with a size of approximately 180 nm.

Encapsulated docetaxel (DTX) using PLGA was synthesised. It was characterised that the DTX-loaded PLGA have a spherical shape with an estimated size of 120 nm.

Nano-constructed gold sphere with an average size of 30 nm, in which the active ingredient yields 60% of the gold nanoparticles surface.

[22]

It exhibited more potent inhibitory effects and induced apoptosis with a combination of laser radiation. Without laser radiation, the PEGylated DOX was still able to induce apoptosis but only up to a certain range, indicating that DOX was only partially released from the HP without radiation. It was also shown that the PEGylated DOX internalised the cancer cells more efficiently. However, the PEGylated DOX exhibit more cytotoxicity effects compared to HP alone.

Chapter 5

Doxorubicin

5-Fluororacil

-

Gold nanoparticles

Gold nanoparticles

Magnetic nanoparticles

CAL-27

HSC-3

OECM1

The typical 10 nm diameter of an iron core with gold shell nanoparticles exhibits anticancer characteristics by autophagy, mediated by activating the mitochondria. It stops the growth of the cancer cells without affecting the healthy cells.

Two types of nuclear targeting gold nanospheres (NLSAuNPs) were used as a carrier for a chemotherapeutic agent, 5-fluorouracil with an average size of 15 and 30 nm.

Gold nanoparticles were fabricated with polymer PEG, and its function is enhanced by conjugating it with PDPN antibody and DOX ((PDPN Ab)-AuNP-DOX). The size of the spherical (PDPN Ab)AuNP-DOX was estimated to be 150 nm. When compared, the NLS-AuNPs were observed to be the most effective in enhancing the efficacy of 5-FU compared to the RGDconjugated AuNPs. It was identified that the cell viability reduces as the concentration of 5fluorouracil increases, and the lowest cell availability was noted to be with the treatment using the 30 nm NLS-AuNPs compared to the 15 nm NLS-AuNPs and the RGD-conjugated AuNPs. A low dose of 10 μg/ml successfully killed 80% of the OECM1 cancer cells, but the human normal oral keratinocytes (hNOK) cells remained unaffected until the dose reached 100 μg/ml. Aged gold shell–coated iron core nanoparticles were also tested in both cells, resulting in no notable reduction in the cell viability of the cells, aligning with the hypothesis that the gold shell–coated iron core nanoparticles are biocompatible.

Via flow cytometry, it was seen that (PDPN Ab)-AuNP-DOX (IC50 = 0.35μM) possessed up to 3 times more potent cellular uptake and antitumour effects compared to free DOX (IC50 = 0.41μM). However, treatment with this approach exhibited much superior antitumour performance with the combination of laser radiation (IC50 = 0.26μM).

Nanomedicines in Oral Cancer Therapies

[27]

[26]

[25]

135

Dendrimers

Nanoshells

UMSCC-1 UMSCC17B UMSCC22B

KB Cells

136

Anti-HER2 antibody

Methotrexate

[28]

[29]

Using mRNA levels as biomarkers, the use of dendrimers is threefold more effective in treating different tumour types by exhibiting their antitumour effects with reference to FBP-Į expressions compared to free methotrexate.

Both conjugated and unconjugated nanoshells have proved their ability to induce apoptosis in both cell lines. However, the unconjugated nanoshells showed lower cell mortality rates. The cells mortality rate for the unconjugated nanoshells was noted to be accountable for 1.5% of apoptosis in the KB cells and only up to 0.9% in HeLaS3 cell lines, whereas upon using the anti-HER2 conjugated nanoshells, a higher rate (69.4%) of apoptosis was spotted in the KB cell lines. Still, HeLaS3 cell lines revealed a lower rate of apoptosis estimated to be around 4%.

The dendritic polymers were attached with generation 5 polyamidoamine (PAMAM) dendrimer onto folic acid and methotrexate, with approximately 26,530 g/mol molar mass.

Gold-silica nanoshells were conjugated with anti-HER2 nanobody, in which the gold and silica used in this process were 3nm and 100 nm in size, respectively. Thus, this results in gold-silica nanoshells with a diameter of 100 nm and a thickness of 10 nm.

Chapter 5

Doxorubicin

Phloretin, Chitosan

Curcumin, Chitosan

Natural products (Chitosan)

Natural products (Phloretinloaded Chitosan)

Natural products (Curcuminloaded Chitosan)

HN-22

KB cells

SCC-9

The curcumin-loaded nanoparticles (Cur-NP CSL) were designed in a spherical shape with a size ranging from 104 to 125 nm with a positive charged outer surface.

A plant-based nanomedicine named phloretin is loaded in chitosan (PhCsNPs), spherical with a size distribution ranging between 80 to 100 nm.

Catechol-modified chitosan or also known as hyaluronic acid nanoparticles (Cat-NPs) were utilised as a carrier of oral cancer therapeutics, doxorubicin, shaped into a sphere with an approximate size of 160 nm.

The release of DOX was rapid in the first 3 hours via the Cat-NPs, but the amount of DOX released was equal to that of free DOX. However, the inhibitory effect of the Cat-NPs was superior to the free DOX, which was demonstrated by the IC50 values of 2.95 μg/ml and 1.51 μg/ml, respectively. It was also revealed that the Cat-NPs are more cytotoxic than free DOX. The Cat-NPs showed more extensive cell death (22%) than the free DOX (16%). The PhCsNPs were able to inhibit cell growth at a lower concentration. Its anticancer effect was measured using IC50 values around 20.34 μg/ml, which were then compared to doxorubicin with IC50 values of 2.08 μg/ml. Concerning the cytotoxicity effect of PhCsNPs, it was equalised to doxorubicin's IC50 values of PhCsNPs described that the cell death was caused by necrosis. The cytotoxicity effect was monitored every 24 hours, and it was observed that the free curcumin contributed to high toxicity. After 72 hours of incubation, the percentage of viable cells' depletion was 90% with the treatment of free curcumin, whereas the Cur-NP CSL reduced only 45% of viable cells. This result supported the IC50 value of the Cur-NP CSL, which was higher than the free curcumin.

Nanomedicines in Oral Cancer Therapies

[32]

[31]

[30]

137

SCC152

SCC090

OSCC25

OSCC4

138

Natural products (Curcumin)

Natural products (Curcumin)

5-FU, Curcumin

Curcumin

Different preparations of 5fluorouracil and curcumin in spherical nanoemulsions, whether designed to be alone or in combination, were uniformly distributed to be between 150 to 200 nm in size. Three types of 5-fluorouracil and/or curcuminnanoemulsions were tested: 5fluorouracil-nanoemulsions (5FUNE), curcuminnanoemulsions (Cur-NE), and 5-fluorouracil-and-curcuminnanoemulsions (5-FU-Cur-NE)

Curcumin was designed to be a carrier for the natural product, Curcumin, in a microemulsion form with a size range of 40 to 50 nm.

Chapter 5 Most of the curcumin-incorporated microemulsion concentrations showed an increase in cytotoxicity by more than 25%, especially in the OSCC-4 cell lines. This curcumin-incorporated microemulsion drug delivery was also tested to be accompanied using an ultrasound to see whether it affects the cytotoxicity of the curcumin-incorporated microemulsion. It was determined that the ultrasound enhances the cytotoxicity more, causing it to be more lethal to cells. All three forms showed high cytotoxicity in both cell lines and high reduction of viability cells, especially in the SCC-090 cell line, up to 14% after 96 hours of incubation. However, it was suggested that 5-FUNE have the superior anticancer effect, and Cur-NE contributes the least cytotoxicity among all three available forms. However, the IC50 values of 5-FU in SCC-090 and SCC-152 were higher in 5-FUNE (51.79 μg/ml and 62.04 μg/ml) in comparison to the 5-FU-Cur-NE (31.05 μg/ml and 37.46 μg/ml) and in Cur-NE (79.16 μg/ml and 84.18 μg/ml) which suggests that the 5-FU-Cur-NEto be most effective in treating both cell lines. [34]

[33]

Hep-2

Natural products (Curcumin) 5-FU, Curcumin

Curcumin and 5-fluorouracil were separately loaded on PLGA as a nanoplatform to be used as a delivery method in treating oral cancer. The average size of free PLGA was around 50 to 52 nm, while curcumin-loaded PLGA was approximately 62 nm and the average size of 5-fluorouracilloaded PLGA was approximately 120 nm.

The curcumin-loaded PLGA and 5-fluorouracilloaded PLGA used were at different concentrations of 26.4 μg/ml and 36.5 μg/ml, respectively. Caspase 3 expression was used as a biomarker for apoptosis as it acts as one of the major control of cell death. It was noticed that the higher PLGA ratio compared to the drug induces faster release rate profiles and vice versa. There was no statistical difference between the cytotoxicity of the curcumin-loaded PLGA and 5-fluorouracil-loaded PLGA. Although, in 24-hour duration, curcumin-loaded PLGA showed the higher caspase-3 expression. Meanwhile, in the 48-hour duration, it was dominated by 5-fluorouracil-loaded PLGA.

Nanomedicines in Oral Cancer Therapies

[35]

139

Chapter 5

Polymers

Polymers

Polymers

SCC-25 tumourbearing mice

CAL-27 cellsbearing mice

Nanotechnology Avenues

Orthotropic tongue cancer mouse model

In vivo

Doxorubicin

Doxorubicin

Cisplatin

Active Ingredients

The encapsulated PEGylated doxorubicin with hematoporphyrin (RPTD/HP) is a spherical nanoparticle with a size of approximately 180 nm.

PEGylated doxorubicin (PD) is mediated with HN-1, developing HN-1 mediated PEGylated doxorubicin (HNPD) micelles with a diameter of 140 nm approximately.

Nanoparticles Characteristics The cisplatin is polymerised using a polymeric micellar nanoparticle, NC-6004, that is hydrophilic. The diameter of the NC-6004 cisplatin is approximately 30 nm. PD and HNPD significantly showed higher inhibitory effects among the three types of doxorubicin nanoparticles. However, HNPD delivered more antitumour effects and effectively halted the tumour growth and contributed to the lowest toxicity, which was deduced from monitoring the weight of the mice. With the laser radiotherapy, slowing down of tumour growth was noted, resulting in the smallest tumour compared to the treatment using free DOX. No significant changes in body weight were observed with the RPTD/HP. Therefore, it is safe for humans.

There was a significant reduction of lymphatic tumours metastasis in the administration of the NC-6004, and a high concentration of NC-6004 micelles was found accumulated in the lymph nodes. However, free cisplatin was not administered to compare the results.

Key Findings

Table 5-2. List of in vitro studies for nanotechnology avenues for oral cancer treatment

140

[22]

[20]

[21]

References

Gold nanoparticles

Dendrimers

CAL-27 cellsbearing mice

Mice with UM-SCC Xenograft Methotrexate

Doxorubicin

The reduction of tumour size was superior with the administration of (PDPN Ab)-AuNP-DOX with an inhibitory rate of 70% compared to free DOX at 51%. In combination with 15 minutes of laser radiation, treatment with (PDPN Ab)AuNP-DOX also showed the highest increase in the tumour temperature. This approach has shown a reduction of systemic cytotoxicity in animals compared to free methotrexate. Still, two out of six mice were experiencing tumour growth instead of retardation.

Gold nanoparticles were fabricated with polymer PEG, and its function is enhanced by conjugating it with PDPN antibody and DOX ((PDPN Ab)-AuNP-DOX). The dendritic polymers were attached with generation 5 polyamidoamine (PAMAM) dendrimer onto folic acid and methotrexate, with approximately 26,530 g/mol molar mass.

Nanomedicines in Oral Cancer Therapies

[28]

[25]

141

Chapter 5

142

2.2 Synthesis of different nanotechnology avenues 2.2.1 Polymers A recent innovative application that used self-assembled polymeric nanoparticles has received positive deliberation for its capability as a carrier in the nanodrug delivery used in cancer treatment. This polymeric nanodrug provides several advantages, including load capacity, morphology, specificity and the ability to reduce side effects of the treatment [36]. Some of the common anticancer agents used with polymers are cisplatin, doxorubicin, docetaxel and 5-fluorouracil, with polyethylene glycol (PEG), polylactic acid (PLA), and poly-lactic-co-glycolic acid (PLGA) are used to formulate nanodrugs [37]. Several studies on the use of these agents have found differences in the therapeutic and toxicity effects between the free drug and polymerised nanodrug of the same drug. Different polymers require different preparation methods. An example is using PLGA and PEG as a carrier for cisplatin (CDDP). The PLGA-PEG cisplatin is conjugated with NR7 peptide (PLGA/NR7) to target the epidermal growth factor (EGF) and restrict the development of cancer. When the free cisplatin is dissolved in DMSO, the cisplatin is formed into spherical micelles coated with the non-polar PLGA and polar PEG in the outer layer [36]. A different study tested HN-1–mediated PEGylated doxorubicin (HNPD) and free PEGylated doxorubicin (PD) separately against the OSCC tumour cell lines CAL-27 and SCC-25. The conjugation of the HN-1 peptide prepared the PD to assist in penetrating the nanodrug into the tumour cells. HNPD micelles were synthesised via the fusion of doxorubicin with PEG by forming the succinyl linkage and coating the HN-1 peptide as the outer layer using the outer layer nanoprecipitation method. This was done to ensure that the nano properties of the drug are not violated [20]. Because using docetaxel is known to be efficient in fighting against cancer cells, the implementation of polymers on docetaxel is now considered for oral cancer treatment, where DTX using PLGA was compared to free docetaxel on an SCC-9 OSCC cell line used in vitro assays. The PEGylated PLGA DTX was synthesised by adding DTX to the polymers, then adding polyvinyl alcohol while whirling the viscous solution that results in PEGylated PLGA DTX. The PEGylated PLGA DTX has a negative charged outer layer, preventing it from being opsonised and entering the reticuloendothelial system [23].

Nanomedicines in Oral Cancer Therapies

143

2.2.2 Gold nanoparticles Gold nanoparticles (AuNPs) have been used to improve cancer diagnosis and the therapeutics approach for more than two decades ago. The conjugation of the AuNPs with different ligands will enhance their ability to target the cancer cells, such as the function of IgG-conjugated AuNPs in targeting the EGFR in OSCC. The AuNPs can also function as a carrier for chemotherapeutics, such as doxorubicin (DOX) and 5-fluorouracil, which may aid in the cellular uptake, solubility and liability against drug resistance. However, using AuNPs alone is enough to induce cell death in OSCC cells [24]. AuNPs also halt cell growth and significantly regulate the cell cycle by sensitising the tumour cells against radiation [26]. Thus, it can be concluded that using AuNPs is practical in the combination treatment of cancer, especially with radiation therapy. Generally, the synthesis of AuNPs is performed via citrate reduction. Then, they are stabilised using the polymer such as PEG-SH and peptides, followed by the loading of DOX or 5-fluorouracil particles through a coupling reaction [24–26]. A different trial that aimed to use the gold nanoplatform against oral cancer fabricated the AuNPs with the polymer PEG and enhanced its function by conjugating it with PDPN antibody and DOX, also known as the (PDPN Ab)-AuNP-DOX system. This system sought to target the tumour and release its anticancer possessions effectively. The rationale for the conjugation of PDPN antibody in this approach is to enhance the accumulation of the therapeutics via target cell recognition, given that the PDPN antibody is usually overexpressed in squamous cell cancer types [25]. 2.2.3 Magnetic nanoparticles In pursuing anticancer therapeutics with little toxicity towards non-cancerous, healthy cells and organs, the iron core with gold shell nanoparticles was developed to harness the iron core’s magnetic forces protected from being oxidised easily by the outer gold shell. The coating of iron is crucial because it was previously reported that using iron alone will provoke cytotoxicity because of the redox reaction of the iron. In brief, the gold shell–coated iron core nanoparticles were synthesised by sequential synthesising the reverse micelles in a cetyltrimethylammonium bromide/N-butanol/water system which was then purified by washing them in ethanol. The gold shell–coated iron core nanoparticles exhibit anticancer characteristics by autophagy, mediated from activating the mitochondria

144

Chapter 5

and stopping the cancer cells’ growth without affecting the healthy cells [27]. 2.2.4 Dendrimers The progress of the nanodrug delivery system has promoted a potential carrier for targeted therapy in the oral cancer cell via dendrimers or noncationic dendritic polymers. This carrier is sometimes attached to targeting ligands, such as folate or EGF, or anticancer agents, such as paclitaxel and methotrexate. For instance, the conjugation of dendrimers can be demonstrated by the attachment of a fifth-generation 5-polyamidoamine (PAMAM) dendrimer onto the folic acid and methotrexate. It was synthesised by coating the surface of PAMAM dendrimers with acetic anhydride to avoid non-specific targeting of the dendrimer, followed by its reaction with a folic acid ester. Finally, the active ingredient methotrexate was added via esterification [28]. 2.2.5 Nanoshells Using nanoshells as a therapeutic option for OSCC has been established in a previous study. However, it must be combined with photothermal therapy. Nanoshells were believed to have higher accumulation on the tumours due to their nano size and exhibit more retention effects. An example of this approach in treating oral cancer is gold-silica nanoshells conjugated with anti-HER2. This combination was synthesised using a multi-step process that started with the fabrication of nanoshells, followed by the preparation of AuNPs and then the conjugation of the anti-HER2 onto the gold-silica nanoshells. Anti-HER2 nanobody was used specifically because a previous study emphasised that the high rate of overexpression of the antibody, HER2, was found in a patient with OSCC [29]. 2.2.6 Natural products: Chitosan and curcumin Recent studies have focused on the effectiveness of the treatment delivery system alone and considered using natural resources to reduce the toxicity of these treatments. These studies have shown that plant-based chemicals can provide anticancer effects, which might be useful in nanomedicine avenues for oral cancer [31]. Simultaneously, local drug administration has always been a priority in ensuring safer and convenient use in oral cancer therapy, especially the mucoadhesive drug delivery method delivering the therapeutics to the targeted location [30]. Considering both factors, this drug delivery system uses natural products such as chitosan and curcumin to convey their nanotherapeutic effects against oral cancer by adhering to and

Nanomedicines in Oral Cancer Therapies

145

interacting with the mucous membrane. The mucoadhesive drug delivery method can be activated with polymer, such as the natural products, which can be functionalised with thiols or catechol (Cat) via a nucleophilic addition forming the covalent bond while activating the adhesive effects [30]. More than 3,000 plants have been recorded to possess anticancer characteristics. Some are now being studied and experimented with for clinical use [38]. Chitosan is a biopolymer derived from the structural part of crustaceans via de-acetylation, proven to increase cell permeability and cellular uptake [39]. In contrast, the aromatic herb curcumin is a yellow polyphenol extract found in turmeric. It is scientifically known as Curcuma Longa that exhibit anti-carcinogenic properties [32–34]. It was reported that curcumin could induce apoptosis in OSCC, including the caspase pathway, by suppressing the activation of IKK-mediated 1)Nȕ and the gene expression regulated by the 1)Nȕ, such as COX-2 and MMP-9 [32,35]. These natural products were preferred because of their abundance and versatility to be shaped into different forms and their special pharmacological activities. The function of chitosan in oral cancer therapies, alone or in combination with other therapeutic agents, was compared in studies with standard anticancer drugs such as DOX. An example of chitosan against oral cancer is the Cat-modified chitosan nanoparticles, synthesised by the ionic gelation of the catechol-functionalised succinyl chitosan with the catechol-carrying hyaluronic acid [30]. Another study used the same ionic gelation process when synthesising chitosan but functionalised with a natural product called phloretin, resulting in phloretin-loaded chitosan nanoparticles. Phloretin was used because of its pharmacological properties, such as exhibiting the anticancer effect, which enhanced the effect of chitosan in combatting oral cancer [31]. The utilisation of curcumin was also explored in different methods and combinations to evaluate its effectiveness and toxicity. One study explored curcumin microemulsion's stability, effectiveness, and cytotoxicity in the nanotherapeutic approach of treating OSCC with and without ultrasoundaided delivery. The curcumin microemulsions were prepared by adding carrier oil (soybean oil), soybean lecithin, Tween 80 and deionised water to the curcumin powder and then mixed to form curcumin microemulsions [33]. It is also reasonable to compare the study using different nanoemulsion formulations by loading it with curcumin and 5-fluorouracil, alone or in combination, to evaluate its effectiveness and cytotoxicity in OSCC [34].

146

Chapter 5

The curcumin-loaded nanoemulsion, 5-fluorouracil-loaded nanoemulsion and 5-fluorouracil–curcumin nanoemulsion were prepared by applying a homogenisation process using ultrasonic procedures [34]. This combination of curcumin and 5-fluorouracil was again used in another study but by loading them on the PLGA nanomedicine delivery system and preparing them with a double emulsion method, which was tested on human laryngeal SCC, Hep-2 [35].

147

Manufacturer Safaa Elbaz, Cairo University UNC Lineberger Comprehensive Cancer Center, AstraZeneca, Celgene Hadassah Medical Organization Alza Regulon Cytimmune Sciences

Active ingredient

Luteolin or flavonoid natural extract

Carboplatin/NabPaclitaxel/ Durvalumab

Ammonium Polyethylenimine

Stealth Liposomal Cisplatin

Liposomal Cisplatin

71)Į bound to colloidal gold nanoparticles

Product

Luteolin

Carboplatin/ NabPaclitaxel/ Durvalumab

Polyethylenimine (PEI) nanoparticles incorporated in soft liner silicon

SPI-77

Lipoplatin

Aurimmune (CYT6091)

Head and Neck Cancer

Head and Neck Squamous Cell Carcinoma Head and Neck Squamous Cell Carcinoma Head and Neck Cancer Head and Neck Cancer

Oral Cancer

Indication

Clinical Phase 2

Clinical Phase 3 Clinical Phase 3

Clinical phase 1

Clinical phase 2

FDAapproved date/clinical trial status Clinical phase 1

-

-

-

NCT01007240

NCT03174275

NCT03288298

Clinicaltrials.gov identifier

[45]

[44]

[43]

[42]

[41]

[40]

Reference

Table 5-3. List of FDA-approved nanomedicine for oral cancer and nanomedicines currently undergoing clinical trials.

Nanomedicines in Oral Cancer Therapies

Chapter 5

Oxaliplatin Nanoparticles and Method for Preparing Same

Polymeric Nanoparticles

Polymeric Nanoparticles

Lipid Nanoparticles

Polymeric Nanoparticles

WO2016020697 A1

US9393201 B2

WO2015023797 A1

US20150140109 A1

1

2

3

4

Docetaxel-Based Prolonged-Release Cancer Treatment Drug

Lipophilic Nanoparticles for Drug Delivery

Pharmaceutical Compositions of Polymeric Nanoparticles

Patent Title

Classification

Patent No.

Sr. No. Shrikhande, Shruti Bajaj, Amrita N Malhotra, Geena Raut, Preeti Sung Jae Lee Young Hoon Kim Sang Heon Lee Kab Sig Kim Thaxton, Shad, C. Gordon, Leo, I. Mutharasan, Raja, Kannan Grun, Casey, N. Foit, Linda Evgenij Severin Irina Zykova Victor Gulenko Maksim Iurchenko

Inventor(S)

Oy Filana Ltd Unichempharm Ltd

Northwestern University

JW Pharmaceutical Corporation, Bio-Synectics, Inc.

Cipla Limited, King, Lawrence

Organization

Table 5-4. List of patents approved nano-based medicines for oral cancer treatments

148

May 21, 2013

Feb. 19, 2015

Jul. 12, 2012

Feb. 11, 2016

Publication/ Application Date & Year

[49]

[48]

[47]

[46]

Reference

Docetaxel-Based Prolonged-Release Cancer Treatment Drug Smart Polymeric Nanoparticles Which Overcome Multidrug Resistance to Cancer Chemotherapeutics and Treatment-Related Systemic Toxicity

Metallic Nanoparticles

Polymeric Nanoparticles

Polymeric Nanoparticles

US8911786 B2

US8784895 B2

WO2013171382 A1

EP2648760 A4

7

8

9

Polymeric Nanoparticles

6

Polymeric Nanoparticles

US9023395 B2

5

Formulation of Active Agent Loaded Activated PLGA Nanoparticles for Targeted Cancer NanoTherapeutics Nanoparticle Comprising Rapamycin and Albumin as Anticancer Agent Multifunctional Metal Nanoparticles Having a Polydopamine-Based Surface and Methods of Making and Using the Same

Maitra Anirban Pramanik Dipankar

Desai Neil P. Soon-Shiong Patrick Trieu Vuong Messersmith Phillip B. Black, Iv Kvar C. L. Yi Ji Rivera Jose G. Severin, Evgenij Zykova, Irina Gulenko, Victor Iurchenko, Maksim

Arthur R. C. Braden Jamboor K. Vishwanatha

Nanomedicines in Oral Cancer Therapies

Univ Johns Hopkins

Oy Filana Ltd Unichempharm Ltd

Northwestern University

Abraxis Bioscience, LLC

University Of North Texas Health Science Center At Fort Worth

Oct. 16, 2013

Nov. 21, 2013

Sep. 20, 2012

Jul. 22, 2010

Oct. 16, 2008

[54]

[53]

[52]

[51]

[50]

149

-

NAH, JaeWoon JUNG, Teok Rae CHAE, Su Young JANG, Mi Kyeong Choi, Chang Yong

Anticancer Agent Loaded Hydrophobic Bile Acid Conjugated Hydrophilic Chitosan Oligosaccharide Nanoparticles and Preparation Method Thereof Ceramic-Based Nanoparticles for Entrapping Therapeutic Agents for Photodynamic Therapy and Method of Using Same

Polymeric Nanoparticles (Naturalbased)

Polymeric Nanoparticles

WO2007086651 A1

CA2513759 C

13

14

The Research Foundation of State University Of New York Health Research Inc.

-

Polymeric Nanoparticles

WO2007108618 A1

12

HONG, Soon Hai CHUNG, Jin Hyuk

Water-Soluble Organometallic Nanoparticles and Method for Preparing the Same

11

JW Pharmaceutical Corporation Bio-Synectics, Inc.

Sung Jae Lee Young Hoon Kim Sang Heon Lee Kab Sig Kim

WO2011034394 A2

Oxaliplatin Nanoparticles and Method for Preparing Same

Polymeric Nanoparticles

US8242165 B2

10

Creighton University

Dash Alekha K. Trickler William J.

Chapter 5 Mucoadhesive Nanoparticles for Cancer Treatment

Polymeric Nanoparticles (Naturalbased)

150

Aug. 12, 2004

Aug. 02, 2007

Sep. 27, 2007

Mar. 24, 2011

Nov. 5, 2009

[59]

[58]

[57]

[56]

[55]

Nanomedicines in Oral Cancer Therapies

151

3. Application of nanodrug delivery system in oral cancer Currently, the availability of treatment choices for advanced oral cancer is minimal, and the effects are still considered to be at sub-therapeutic levels. Although various conventional therapies such as surgery or chemotherapy are available, there will be a consequential impact on patients’ overall health and quality of life. With improved drug delivery systems, new therapeutic approaches with low toxicity are now developed to defeat the challenges of administering anticancer agents. These new therapeutic approaches for oral cancer treatments involve administering anticancer agents using drugloaded nanoparticles and other nanotechnology avenues [60]. A general schematic diagram (Figure 5-1) of the OSCC, refers to the cancer present between the vermilion border of the lips as well as the junction of the hard and soft palates (the posterior one-third of the tongue) and presents the general treatment options available for the OSCC. Smoking and drinking are significant risk factors for OSCC. Nanotechnology has been used for chemotherapy, radiotherapy and immunotherapy. The use of nanotechnology variants was demonstrated in many studies and is potentially safe for oral cancer treatment. Alongside their biocompatibility properties, nanoparticles are also known for their large surface area. Therefore, they can enhance their bioactivity and control release properties [60]. Many nanotechnology alternatives, such as polymers, AuNPs, magnetic nanoparticles, dendrimers, and nanoshells, and natural products, such as chitosan and curcumin, are present for managing oral cancer cells. Therefore, it would be fair to compare all the in vitro and in vivo tests of the new nanotechnology avenues to evaluate the efficacy and cytotoxicity of chemotherapeutics using the nanotechnology methods. Different oral cancer cell lines are used as model systems in the studies. Some of the common ones used are HSC-3, CAL-27, SCC-9 and KB cell lines. Most of the nanomedicines used in treating oral cancers in this review are spherical in shape, with sizes ranging from 15 nm to 180 nm.

152

Chapter 5

Figure 5-1: Oral squamous cell carcinoma (OSCC) has different stages of cancer which are treated using surgery, chemotherapy, radiotherapy and immunotherapy Created with BioRender.com

Nanomedicines in Oral Cancer Therapies

153

Different nanotechnology approaches can be compared using the same cancer cell lines to conclude which method provides a better option. Three different studies with two different nanotechnology avenues were analysed using HSC-three cancer cell lines as the response model. Polymers and AuNPs have the same size of approximately 30 nm. Each nanotechnology method was used as carriers for different active ingredients, namely CDDP, DOX and 5-fluorouracil, where the polymer, NC-6004, was used to load the CDDP. In contrast, DOX and 5-fluorouracil were loaded in gold nanospheres. The rationale for having a small size is to enhance their ability to permeate and exhibit the EPR effect via the leaky tumour capillary, thus increasing the accumulation of drugs and retention time on the tumour sites to execute its antitumour effects [21]. This hypothesis was confirmed because the doxorubicin-loaded AuNPs showed an escalation of nucleus localisation and cellular uptake of the HSC-3 cells. This led to enhancing the efficacy of the AuNP-coated DOX and 5-fluorouracil. In contrast, CDDP-loaded NC-6004 showed an accumulation of the micelles. However, there was a slight contradiction of results regarding its inhibitory effects compared to AuNPs. However, the CDDP-loaded NC-6004 improved in reducing cytotoxicity, further acknowledged by the in vivo test on the orthotropic tongue cancer mouse model. The less potent inhibitory effects of CDDPloaded NC-6004 might be attributed to the properties of the polymer NC6004 because this was the first study using NC-6004 in oral cancer. NC6004 was shown to have antitumour effects against gastric cancer; therefore, it might not be as effective on oral cancer cells [21]. However, it can be concluded that NC-6004 is safe to administer and test in humans. Three different approaches that carried DOX as the active ingredients were studied using CAL-27 cancer cell lines. Two of these studies involved using PEG polymers, and one of them uses the AuNP approach attached to PEG polymer. All three methods have the same spherical shape with a 140–180 nm size range. The sizes are within the accepted range (40–200 nm) that would be able to pass through the tumour vessels and will be able to accumulate in the tumour vessels [15]. These three different methods have been demonstrated to internalise the cancer cells more efficiently and effectively in inhibiting the growth of cancer cells, especially when the treatment is combined with laser radiation. This result was confirmed again by the in vivo test of CAL-27 cells-bearing mice, where 15 minutes of laser treatment had shown an increase in tumour temperature that would result in cell death. The enhancement of inhibitory effects in the combined treatment might be attributed to the properties of the conjugates attached to the nanoparticles. For example, hematoporphyrin acts as the photosensitiser, easing the laser radiation to target the tumour and assisting the apoptosis

154

Chapter 5

process [22]. However, the polymerised nanoparticles still induced cytotoxic effects superior to administering the DOX alone. It is reasonable to believe that the polymer, PEG, is insufficient to prevent the active component from becoming cytotoxic. This action may be improved by changing the polymer-to-drug ratio. Other studies used SCC-9 oral cancer cell lines to evaluate the efficacy and cytotoxicity of nanotechnology avenues using PLGA polymer and chitosan, where both have the same average size of 120 nm and are spherical. The PLGA polymer was loaded with a chemotherapeutic agent, docetaxel. Meanwhile, chitosan was designed to be loaded with curcumin. Both demonstrated contradicting cytotoxicity results wherein the synthetic PLGA polymer resulted in a higher cytotoxicity effect. Meanwhile, the other study demonstrated that the curcumin-loaded chitosan has a lower cytotoxicity effect than free curcumin. The high cytotoxicity effect of the PLGA polymer might be attributed to PLGA’s inability to block the cytotoxic effects of docetaxel. In contrast, the lower toxicity level of curcumin-loaded chitosan might be explained by the properties of the outer chitosan layer, where it uniquely possesses adjustable properties to regulate the drug release and affect the loaded anticancer agents [61]. Apart from SCC-9, oral cancer KB cell lines were also used as a model system to analyse two different approaches to using nanotechnology in oral cancer treatment: nanoshells and chitosan (the natural biopolymer). Nanoshell is generally a metallic shell–covered nanoparticle. In this study, gold-silica is used as the shell covering anti-HER2 nanobody resulting in gold-silica nanoshells of 100-nm diameter. In contrast, chitosan was used as a carrier of a plant-based nanomedicine called phloretin found in the apple or bark of pear and cherry trees. Chitosan has a size ranging from 80 to 100 nm [31]. The gold-silica nanoshell and the phloretin-loaded chitosan can induce apoptosis in the KB cells. Although the phloretin-loaded chitosan still provoked the cytotoxic effects equal to that of free doxorubicin, it was deduced that it kills the cancer cells via necrosis, which might be harmful if administered to patients. Therefore, the components used in phloretin-loaded therapeutics must be modified to reduce lethal cytotoxicity effects. Meanwhile, the use of nanoshells should be explored more in vivo to ensure its safety to be administered in humans, as it looks promising to treat oral cancer.

Nanomedicines in Oral Cancer Therapies

155

4. Challenges of targeted drug delivery using nanoparticles Tumours of different sizes and structures produce a heterogeneous cell expression for targeted therapy. Thus, targeted therapy is not an easy process because it only destroys some parts of the tumour cells that express the same expression as it does. Some parts of the tumour might survive with the target and continue to increase. Further, sometimes while targeting tumours, other tumour cells, such as tumour stem cells, may be overlooked. This action can cause the relapse of the tumour, making it challenging to remove the whole tumour [61]. The complexity in delivering the nanotherapeutics to the targeted sites might also contribute to less effective therapies. Normally, the internalisation process occurs once the nanomedicine is bound to the target receptor. The active ingredient should be released inside the cell to function. However, problems might arise during the internalisation process when the drug cannot avoid the lysosome, preventing the drug from reaching the target sites [61]. Another concern related to drug delivery using nanoparticles is the safety and toxicology effects. Although the small size benefits drug delivery as a nanocarrier, the structure and physicochemical properties might be altered because of several interactions in the body, producing unwanted side effects [39]. Therefore, it is crucial to test the nano-based medicine in vitro to evaluate its toxicity. Substantial research is required because nanoscale therapeutics and toxicology are considered underexplored areas in literature research.

5. Conclusion Above all, oral cancer treatment aims to induce cancer cell death without affecting the healthy cells, which is possible via targeted and distinctive site delivery of the therapeutic drugs to attain its optimum potential. According to this objective, the carrier-mediated delivery system will become an excellent option for overcoming the current issues with drug delivery systems. Extensive in vitro and in vivo studies will be used as model systems in investigating the pharmacokinetic and pharmacodynamics criteria as the guides to interpret the proper drug dose and release mechanism to treat oral cancer. In the meantime, nano-based medicines are becoming more prominent in the pharmaceutical industry as a promising strategy for combatting cancer. Their efficacy towards oral cancer is under research for it to be applied in clinical settings. The accomplishments demonstrated in

156

Chapter 5

nano-based medicine research have shown immense potential in boosting patients' health outcomes and quality of life with oral cancer. The continuation of research from different multidisciplinary teams is crucial in recognising the importance of finding a practical approach to overcome the challenges of improving the drug delivery in oral cancer treatment and any cancer in general. Improving clinicians’ knowledge of the effectiveness of nano-based medicine in oral cancer would be helpful in developing the awareness of the nano-based medicine approach via high-quality and evidence-based research. Its potential effect should be highlighted for diagnostic tests, lymph node evaluation as a guide before surgery, and managing primary and recurrent oral cancer. Indeed, challenges must be addressed before nano-based medicine can be applied in clinical practice as an alternative to overcome cancer. Thus, it is comprehensible that future prospective studies must interpret the treatment strategies.

References 1. 2.

3.

4.

5.

6.

J. Ali, B. Sabiha, H. U. Jan, S. A. Haider, A. A. Khan, & S. S. Ali, Genetic etiology of oral cancer. Oral Oncology, 70 (2017) 23–28. https://doi.org/10.1016/j.oraloncology.2017.05.004. D. Maleki, M. Ghojazadeh, S.-S. Mahmoudi, S.-M. Mahmoudi, F. Pournaghi-Azar, A. Torab, R. Piri, S. Azami-Aghdash, & M. Naghavi-Behzad, Epidemiology of oral cancer in Iran: A systematic review. Asian Pacific Journal of Cancer Prevention, 16 (2015) 5427–5432. https://doi.org/10.7314/APJCP.2015.16.13.5427. M. Kumar, R. Nanavati, T. Modi, & C. Dobariya, Oral cancer: Etiology and risk factors: A review. Journal of Cancer Research and Therapeutics, 12 (2016) 458. https://doi.org/10.4103/0973-1482.186696. S. Marcazzan, E. M. Varoni, E. Blanco, G. Lodi, & M. Ferrari, Nanomedicine, an emerging therapeutic strategy for oral cancer therapy. Oral Oncology, 76 (2018) 1–7. https://doi.org/10.1016/j.oraloncology.2017.11.014. X.-J. Chen, X.-Q. Zhang, Q. Liu, J. Zhang, & G. Zhou, Nanotechnology: A promising method for oral cancer detection and diagnosis. Journal of Nanobiotechnology, 16 (2018) 52. https://doi.org/10.1186/s12951-018-0378-6. A. I. Irimie, L. Sonea, A. Jurj, N. Mehterov, A. A. Zimta, L. Budisan, C. Braicu, & I. Berindan-Neagoe, Future trends and emerging issues

Nanomedicines in Oral Cancer Therapies

7.

8.

9.

10. 11.

12. 13. 14.

15. 16. 17.

157

for nanodelivery systems in oral and oropharyngeal cancer. International Journal of Nanomedicine, 12 (2017) 4593–4606. https://doi.org/10.2147/IJN.S133219. M. Tangthongkum, V. Kirtsreesakul, P. Supanimitjaroenporn, & P. Leelasawatsuk, Treatment outcome of advance staged oral cavity cancer: Concurrent chemoradiotherapy compared with primary surgery. European Archives of Oto-Rhino-Laryngology, 274 (2017) 2567–2572. https://doi.org/10.1007/s00405-017-4540-9. D. Lorusso, E. Bria, A. Costantini, M. Di Maio, G. Rosti, & A. Mancuso, Patients’ perception of chemotherapy side effects: Expectations, doctor-patient communication and impact on quality of life - An Italian survey. European Journal of Cancer Care, 26 (2017) e12618. https://doi.org/10.1111/ecc.12618. S. Cheraghlou, A. Schettino, C. K. Zogg, & B. L. Judson, Changing prognosis of oral cancer: An analysis of survival and treatment between 1973 and 2014. The Laryngoscope, 128 (2018) 2762–2769. https://doi.org/10.1002/lary.27315. S. M. Miller & A. Z. Wang, Nanomedicine in chemoradiation. Therapeutic Delivery, 4 (2013) 239–250. https://doi.org/10.4155/tde.12.147. P. G., N. Kalarikkal, & S. Thomas, Challenges in nonparenteral nanomedicine therapy. Theory Appl. Nonparenteral Nanomedicines (Elsevier, 2021), pp. 27–54. https://doi.org/10.1016/B978-0-12820466-5.00002-8. B. Virupakshappa, Applications of nanomedicine in oral cancer. Oral Health and Dental Management, 11 (2012) 62–8. J. Wolfram & M. Ferrari, Clinical cancer nanomedicine. Nano Today, 25 (2019) 85–98. https://doi.org/10.1016/j.nantod.2019.02.005. H. He, L. Liu, E. E. Morin, M. Liu, & A. Schwendeman, Survey of clinical translation of cancer nanomedicines—Lessons learned from successes and failures. Accounts of Chemical Research, 52 (2019) 2445–2461. https://doi.org/10.1021/acs.accounts.9b00228. R. Zein, W. Sharrouf, & K. Selting, Physical properties of nanoparticles that result in improved cancer targeting. Journal of Oncology, 2020 (2020) 1–16. https://doi.org/10.1155/2020/5194780. I. Khan, K. Saeed, & I. Khan, Nanoparticles: Properties, applications and toxicities. Arabian Journal of Chemistry, 12 (2019) 908–931. https://doi.org/10.1016/j.arabjc.2017.05.011. A. Wicki, D. Witzigmann, V. Balasubramanian, & J. Huwyler, Nanomedicine in cancer therapy: Challenges, opportunities, and clinical applications. Journal of Controlled Release, 200 (2015)

158

18.

19.

20.

21.

22.

23.

24.

25.

26.

Chapter 5

138–157. https://doi.org/10.1016/j.jconrel.2014.12.030. M.-C. Q. Y. Zhu, L.-M. Wen, R. Li, W. Dong, S.-Y. Jia, Recent advances of nanodrug delivery system in oral squamous cell carcinoma treatment. European Review for Medical and Pharmacological Sciences, 23 (2019) 9445–9453. G. Calixto, B. Fonseca-Santos, M. Chorilli, & J. Bernegossi, Nanotechnology-based drug delivery systems for treatment of oral cancer: A review. International Journal of Nanomedicine, (2014) 3719. https://doi.org/10.2147/IJN.S61670. Y. Wang, G. Wan, Z. Li, S. Shi, B. Chen, C. Li, L. Zhang, & Y. Wang, PEGylated doxorubicin nanoparticles mediated by HN-1 peptide for targeted treatment of oral squamous cell carcinoma. International Journal of Pharmaceutics, 525 (2017) 21–31. https://doi.org/10.1016/j.ijpharm.2017.04.027. K. Endo, T. Ueno, S. Kondo, N. Wakisaka, S. Murono, M. Ito, K. Kataoka, Y. Kato, & T. Yoshizaki, Tumour-targeted chemotherapy with the nanopolymer-based drug NC-6004 for oral squamous cell carcinoma. Cancer Science, 104 (2013) 369–374. https://doi.org/10.1111/cas.12079. S. Shi, L. Zhang, M. Zhu, G. Wan, C. Li, J. Zhang, Y. Wang, & Y. Wang, Reactive oxygen species-responsive nanoparticles based on PEGlated prodrug for targeted treatment of oral tongue squamous cell carcinoma by combining photodynamic therapy and chemotherapy. ACS Applied Materials & Interfaces, 10 (2018) 29260–29272. https://doi.org/10.1021/acsami.8b08269. P. Gupta, M. Singh, R. Kumar, J. Belz, R. Shanker, P. D. Dwivedi, S. Sridhar, & S. P. Singh, Synthesis and in vitro studies of PLGADTX nanoconjugate as potential drug delivery vehicle for oral cancer. International Journal of Nanomedicine, 13 (2018) 67–69. https://doi.org/10.2147/IJN.S124995. M. M. Essawy, S. M. (Oဨ6KHLNK H. S. Raslan, H. S. Ramadan, B. Kang, I. M. Talaat, & M. M. Afifi, Function of gold nanoparticles in oral cancer beyond drug delivery: Implications in cell apoptosis. Oral Diseases, 27 (2021) 251–265. https://doi.org/10.1111/odi.13551. Z. Liu, J. Shi, B. Zhu, & Q. Xu, Development of a multifunctional gold nanoplatform for combined chemo-photothermal therapy against oral cancer. Nanomedicine, 15 (2020) 661–676. https://doi.org/10.2217/nnm-2019-0415. M. A. Mackey & M. A. El-Sayed, Chemosensitization of cancer cells via gold nanoparticle-induced cell cycle regulation. Photochemistry

Nanomedicines in Oral Cancer Therapies

27.

28.

29.

30.

31.

32.

33.

34.

159

and Photobiology, 90 (2014) 306–312. https://doi.org/10.1111/php.12226. Y.-N. Wu, D.-B. Shieh, L.-X. Yang, H.-S. Sheu, R. Thordarson, D.H. Chen, & F. Braet, Characterization of iron core–gold shell nanoparticles for anticancer treatments: Chemical and structural transformations during storage and use. Materials, 11 (2018) 2572. https://doi.org/10.3390/ma11122572. B. B. Ward, T. Dunham, I. J. Majoros, & J. R. Baker, Targeted dendrimer chemotherapy in an animal model for head and neck squamous cell carcinoma. Journal of Oral and Maxillofacial Surgery, 69 (2011) 2452–2459. https://doi.org/10.1016/j.joms.2010.12.041. F. Sheikholeslami, R. Fekrazad, Rasaee, S. Ardestani, Hakimiha, Farokhi, & Kalhori, Treatment of oral squamous cell carcinoma using anti-HER2 immunonanoshells. International Journal of Nanomedicine, (2011) 2749. https://doi.org/10.2147/IJN.S24548. C. Pornpitchanarong, T. Rojanarata, P. Opanasopit, T. Ngawhirunpat, & P. Patrojanasophon, Catechol-modified chitosan/hyaluronic acid nanoparticles as a new avenue for local delivery of doxorubicin to oral cancer cells. Colloids and Surfaces % Biointerfaces, 196 (2020) 111279. https://doi.org/10.1016/j.colsurfb.2020.111279. A. V. A. Mariadoss, R. Vinayagam, V. Senthilkumar, M. Paulpandi, K. Murugan, B. Xu, G. K.M., V. S. Kotakadi, & E. David, Phloretin loaded chitosan nanoparticles augments the pH-dependent mitochondrial-mediated intrinsic apoptosis in human oral cancer cells. International Journal of Biological Macromolecules, 130 (2019) 997–1008. https://doi.org/10.1016/j.ijbiomac.2019.03.031. L. Mazzarino, G. Loch-Neckel, L. D. S. Bubniak, S. Mazzucco, M. C. Santos-Silva, R. Borsali, & E. Lemos-Senna, Curcumin-loaded chitosan-coated nanoparticles as a new approach for the local treatment of oral cavity cancer. Journal of Nanoscience and Nanotechnology, 15 (2015) 781–791. https://doi.org/10.1166/jnn.2015.9189. M.-H. Lee, Lin, Thomas, Chen, Shen, Yang, W.-J. Yang, & Thomas, In vitro suppression of oral squamous cell carcinoma growth by ultrasound-mediated delivery of curcumin microemulsions. International Journal of Nanomedicine, 7 (2012) 941-951. https://doi.org/10.2147/IJN.S28510. S. Srivastava, S. Mohammad, A. B. Pant, P. R. Mishra, G. Pandey, S. Gupta, & S. Farooqui, Co-delivery of 5-fluorouracil and curcumin nanohybrid formulations for improved chemotherapy against oral squamous cell carcinoma. Journal of Maxillofacial and Oral Surgery,

160

35.

36.

37.

38.

39. 40. 41. 42. 43.

44.

Chapter 5

17 (2018) 597–610. https://doi.org/10.1007/s12663-018-1126-z. S. M. Masloub, M. H. Elmalahy, D. Sabry, W. S. Mohamed, & S. H. Ahmed, Comparative evaluation of PLGA nanoparticle delivery system for 5-fluorouracil and curcumin on squamous cell carcinoma. Archives of Oral Biology, 64 (2016) 1–10. https://doi.org/10.1016/j.archoralbio.2015.12.003. Z.-Q. Wang, K. Liu, Z.-J. Huo, X.-C. Li, M. Wang, P. Liu, B. Pang, & S.-J. Wang, A cell-targeted chemotherapeutic nanomedicine strategy for oral squamous cell carcinoma therapy. Journal of Nanobiotechnology, 13 (2015) 63. https://doi.org/10.1186/s12951015-0116-2. S. A. Gharat, M. Momin, & C. Bhavsar, Oral squamous cell carcinoma: current treatment strategies and nanotechnology-based approaches for prevention and therapy. Critical ReviewsTM in Therapeutic Drug Carrier Systems, 33 (2016) 363–400. https://doi.org/10.1615/CritRevTherDrugCarrierSyst.2016016272. D. Nandi, A. Singal, & A. Nag, Drug resistance in cancer and role of nanomedicine-based natural products. Bioact. Nat. Prod. Manag. Cancer from Bench to Bedside (Singapore: Springer Singapore, 2019), 177–218. https://doi.org/10.1007/978-981-13-7607-8_9. D. Chakraborty, Nanoparticles as 'smart' pharmaceutical delivery. Frontiers in Bioscience, 18 (2013) 1030. https://doi.org/10.2741/4161. S. Elbaz, Therapeutic effect of luteolin natural extract versus its nanoparticles on tongue squamous cell carcinoma cell line. (2017). https://clinicaltrials.gov/ct2/show/NCT03288298. J. Weiss, Carboplatin, nab-paclitaxel, durvalumab before surgery and adjuvant therapy in head and neck squamous cell carcinoma. (2017). https://www.clinicaltrials.gov/ct2/show/NCT03174275. A. B. Sharon, Antibacterial properties of silicon incorporated with quaternary ammonium polyethylenimine nanoparticles. (2009). https://clinicaltrials.gov/ct2/show/NCT01007240. S. C. White, P. Lorigan, G. P. Margison, J. M. Margison, F. Martin, N. Thatcher, H. Anderson, & M. Ranson, Phase II study of SPI-77 (sterically stabilised liposomal cisplatin) in advanced non-small-cell lung cancer. British Journal of Cancer, 95 (2006) 822–828. https://doi.org/10.1038/sj.bjc.6603345. T. Boulikas, Clinical overview on Lipoplatin: A successful liposomal formulation of cisplatin. Expert Opinion on Investigational Drugs, 18 (2009) 1197–1218. https://doi.org/10.1517/13543780903114168.

Nanomedicines in Oral Cancer Therapies

45.

46. 47. 48. 49. 50. 51. 52.

53. 54. 55. 56. 57. 58.

161

S. K. Libutti, G. F. Paciotti, A. A. Byrnes, H. R. Alexander, W. E. Gannon, M. Walker, G. D. Seidel, N. Yuldasheva, & L. Tamarkin, Phase I and pharmacokinetic studies of CYT-6091, a novel PEGylated colloidal gold-rhTNF nanomedicine. Clinical Cancer Research, 16 (2010) 6139–6149. https://doi.org/10.1158/10780432.CCR-10-0978. S. Shrikhandea. N. Bajajg. M. Raut, Pharmaceutical compositions of polymeric nanoparticles, WO 2016/020697 Al, 2016. Y. Sung Jae Lee, S. H. Kim, Hoon, & K. S. K. Lee, Oxaliplatin nanoparticles and method for preparing same, US9393201B2, 2012. Shad C. Thaxton, Leo I. Gordon, Raja Kannan Mutharasan, Casey N. Grun, Linda Foit, Lipophilic nanoparticles for drug delivery, WO2015023797A1, 2015. Evgenij Severin, Irina Zykova, Victor Gulenko, Maksim Iurchenko, Docetaxel-based prolonged-release cancer treatment drug, US20150140109A1, 2015. Arthur R. C. Braden, Jamboor K. Vishwanatha, Formulation of Active Agent Loaded Activated PLGA Nanoparticles for Targeted Cancer Nano-Therapeutics, US9023395 B2, 2015. Neil P. Desai Patrick, Soon-Shiong, Vuong Trieu, Nanoparticle comprising rapamycin and albumin as anticancer agent, US8911786B2, 2014. Phillip B. Messersmith, Kvar C. L. Black, IV, Ji Yi, Jose G. Rivera, Multifunctional metal nanoparticles having a polydopamine-based surface and methods of making and using the same, US8784895B2, 2014. Evgeni Severin, Irina Zykova, Victor Gulenko, Docetaxel-based prolonged-release cancer treatment drug, WO2013171382A1, 2013. Anirban Maitra, Dipankar Pramanik, Smart polymeric nanoparticles which overcome multidrug resistance to cancer chemotherapeutics and treatment-related systemic toxicity, EP2648760A4, 2015. Alekha K. Dash, William J. Trickler, Mucoadhesive nanoparticles for cancer treatment, US8242165B2, 2012. Seungjae Lee, Kim Young-hoon, Lee Sang-heon, Gap-sik Kim, Oxaliplatin nanoparticles and method for preparing same, WO2011034394A2, 2011. Soon Hai Hong, Jin Hyuk Chung, Water-soluble organometallic nanoparticles and method for preparing the same, WO2007108618A1, 2007. Jae-Woon Nah, Teok Rae Jung, Su Young Chae, Mi Kyeong Jang, Chang Yong Choi, Anticancer agent loaded hydrophobic bile acid

162

59.

60.

61.

Chapter 5

conjugated hydrophilic chitosan oligosaccharide nanoparticles and preparation method thereof, WO2007086651A1, 2007. Paras N. Prasad, Indrajit Roy, Earl J. Bergey, Tymish Y. Ohulchansky, Haridas Pudavar, Ravindra K. Pandey, Janet Morgan, Allan Oseroff, Ceramic based nanoparticles for entrapping therapeutic agents for photodynamic therapy and method of using same, CA2513759C, 2004. Ketabat, Pundir, Mohabatpour, Lobanova, Koutsopoulos, Hadjiiski, Chen, Papagerakis, & Papagerakis, Controlled drug delivery systems for oral cancer treatment—Current status and future perspectives. Pharmaceutics, 11 (2019) 302. https://doi.org/10.3390/pharmaceutics11070302. V. T. Chivere, P. P. D. Kondiah, Y. E. Choonara, & V. Pillay, Nanotechnology-based biopolymeric oral delivery platforms for advanced cancer treatment. Cancers, 12 (2020) 522. https://doi.org/10.3390/cancers12020522.

CHAPTER 6 EMERGING NANOTECHNOLOGIES FOR CANCER IMMUNOTHERAPY SHEBA RANI DAVID*1 AND RAJAN RAJABALAYA2 1

School of Pharmacy, University of Wyoming, Laramie, Wyoming, USA PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2

*Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract Cancer is the second most common disease that affects the health of a large number of humans worldwide. It is also responsible for millions of deaths annually. Cancer can affect many organs and tissues because it modifies the functions of both cellular and molecular mechanisms. Surgery, radiation and chemotherapy are used as standard therapies for cancer. However, despite the availability of these standard therapies, some cancers still have limited therapeutic options. The malfunctioning of the immune system is attributed as one of the causes of cancer. Thus, enhancing and repairing the immune system can effectively distinguish and eliminate cancer. Notably, the immune system should be the primary target of immunotherapy. Cancer immunotherapy has been recognised as an effective therapy in recent years because it has demonstrated fruitful results in treating different cancer types. Immune checkpoint blockade and cancer vaccines are examples of

164

Chapter 6

successful immunotherapy. These treatment regimens have started a new age in cancer therapy. Various approaches to emerging nanotechnology have also been introduced in this chapter to enhance cancer immunotherapy. This chapter will focus its attention on several immunotherapy agents used in cancer and discuss the application of nanotechnology in drug targeting and some structures of nanoparticles that could potentially enhance the immunotherapy for cancer.

Abbreviations aAPCs AcFu ADCC ADCP APCs CAR T-cell CDC CEA cHL CTL CTLA-4 DAMPs DC DC-SIGN, Fc FDA Fe3 O4 -ZnO GITR GM-CSF HSCT i:C OVA MDSCs MPLA NK NPs NSCLC ODN PAMAM

Artificial antigen-presenting cells Acetylated fucoidan Antibody-dependent cell-mediated cytotoxicity Antibody-dependent cellular phagocytosis Antigen-presenting cells Chimeric antigen receptor T-cell Complement-dependent cytotoxicity Carcinoembryonic antigen Classical Hodgkin lymphoma Cytotoxic T lymphocyte Cytotoxic T lymphocyte–associated antigen 4 Damage-associated molecular patterns Dendritic cells Dendritic cell-specific intercellular adhesion molecule3-grabbing non-integrin interstitium Fragment crystallisable Food and Drug Administration Iron oxide-zinc oxide Glucocorticoid-Induced tumour necrosis factor receptor Granulocyte-macrophage colony-stimulating factor Haematopoietic stem cell transplantation Polyinosinic:polycytidylic acid Ovalbumin Myeloid-derived suppressive cells Monophosphoryl lipid A Natural killer cells Nanoparticles Non-small cell lung cancer Oligodeoxynucleotide Polyamidoamine

Emerging Nanotechnologies for Cancer Immunotherapy

PAP PBCA PEG– SWCNTs P-LPS PMASH PPI PTX TAA TAMs TCR TLR TME TNF- Į Tregs

165

Prostatic acid phosphatase Polybutyl cyanoacrylate Polyethylene glycol-modified single-walled carbon nanotubes Pleomorphic lipopolysaccharide Disulphide cross-linked polymethacrylic acid Poly propylene imine Paclitaxel Tumour-associated antigen Tumour-associated macrophages T-cell receptor Toll-like receptor Tumour microenvironment Tumour necrosis factor Į Regulatory T cells

1. Introduction With 8.2 million fatalities per year worldwide, cancer is the second most common lethal disease that deteriorates the health of patients suffering from the disease [1]. It is characterised as a condition in which the cells grow out of control, often attacking neighbouring tissues affecting any part of the body and spreading through the lymph systems and the blood [2]. Several treatments, such as surgery, radiation, chemotherapy and immunotherapy, are available for cancer [1]. Chemotherapy is a well-known treatment that works by delivering anticancer drugs systemically to patients to reduce the uncontrolled proliferation of cancerous cells. However, the desired outcome for treating cancer is often not met due to anticancer agents’ non-specific targeting, which produces many side effects and causes poor drug efficacy [3]. Many advancements have been made to treat cancer. However, despite existing treatments, most patients die, indicating the urgent need for new and more effective cancer therapy [4]. Emerging nanotechnology-based methods lead to target-based drug development regimens that optimise the chance of curing patients with cancer [5]. These treatments have sparked much interest in the immunotherapy field because they open up new avenues in cancer therapy [6]. Cancer immunotherapy, which functions by stimulating the host’s immune system to identify and kill tumour cells, has shown excellent potential in inhibiting the growth of tumours and recurrence that could last for a long time compared to radiotherapy and chemotherapy [7]. Although several

166

Chapter 6

other methods have the potential to overcome the host’s immunity in cancer progression, cancer immunotherapy can recover the suppressed immune system of the patient, potentially leading to the disease’s eradication [4]. A variety of cancer immunotherapy approaches, such as cancer vaccines, immune checkpoint blockers, T-cell transfer therapy, immune system modulators and monoclonal antibodies, have proven their effectiveness in many patients [4]. These immunotherapeutic agents are formulated with nanoparticles and delivered via appropriate delivery routes for healing the patient (Figure 6-1).

2. Cancer immunotherapy Cancer immunotherapy aims to boost the patient’s immune response to a tumour. Immunotherapy effectiveness can be affected by several factors, such as the immunosuppressive tumour microenvironment [8]. There are two common yet remarkable immune system mechanisms that eradicate tumour progression in healthy people: cancer immunoediting and cancer immunity cycle.

2.1 Cancer immunoediting Cancer immunoediting is how immune cells defend against cancer development by shaping the immunogenicity of developing tumours [9]. It can work either by controlling/preventing or by promoting the growth of tumours [10]. The process is performed through three phases: elimination, equilibrium and escape. The innate and adaptive immune systems detect and eradicate tumour cells in the elimination phase. This process is termed cancer immunosurveillance [11]. Tumour cells that survived the elimination phase will move into the equilibrium phase, where the immune system controls the development of total tumour cells [12]. In this phase, adaptive immunity plays a role in restricting the growth of untraceable tumour cell variants and edits tumour cell immunogenicity [13]. The remaining tumour cells carrying different mutations cause them to be resistant to immune detection, thus producing a new population of tumour clones with diminished immunogenicity and leading to the progression of the cells into the escape phase [14]. When the immune system pressures the genetically unstable cells, tumour variants are generated that are not recognised by the immune system. These cells become insensitive to effector mechanisms and acquire an immunosuppressive tumour microenvironment [15].

Emerging Nanotechnologies for Cancer Immunotherapy

Figure 6-1: Overview of cancer immunotherapy Created with BioRender.com

167

168

Chapter 6

2.2 Cancer immunity cycle The cancer immunity cycle, comprising seven steps developed by Daniel Chen and Ira Mellman, is an anticancer immune response that effectively eradicates cancer cells by initiating stepwise events. The first step involves the dendritic cells releasing neoantigens generated by oncogenesis. For this step to elicit an anticancer T-cell response, there must be signals that specify immunity if peripheral tolerance to the tumour antigens is induced. The signals might contain pro-inflammatory cytokines and factors discharged by dying tumour cells. In the second step, the dendritic cells present the captured antigens on MHC I and MHC II molecules to T cells. The antigen presentation causes the priming and activation of effector T cells responses against the cancer-specific antigens (step 3) that are viewed as foreign or against incomplete central tolerance. The activation of effector T cells causes the trafficking of the cells to the tumours (step 4) and infiltrates them into the tumour bed (step 5). The effector T cells then recognise and bind the cancer cells by interacting the T-cell receptor (TCR) with its antigen linked to MHC I (step 6), eventually killing the target cancer cell (step 7). Additional tumour-associated antigens are released when cancer cells are killed to enhance the breadth and depth of the response in consequent revolutions of the cycle (step 1) [16]. Thus, the cycle continues and protects the healthy person from developing cancer. Table 6-1 Cancer immunity cycle with examples Source: Chen DS, Mellman I. Oncology Meets Immunology: The Cancer Immunity Cycle. Immunity

Steps 1. Release of cancer antigens 2. Cancer antigen presentation

(+) Stimulators Immunogenic or necrotic cell death

(–) Inhibitors Tolerogenic or apoptotic cell death

• Pro-inflammatory cytokines (e.g., TNF-Į IL1, IFN-Į • Immune cell factors: CD40L/CD40 • Endogenous adjuvants released from dying tumours: CDN (STING ligand), ATP, HMGB1 • Gut microbiome products: TLR ligands

IL-10, IL-4, IL-13

Emerging Nanotechnologies for Cancer Immunotherapy

3. Priming and activation 4. Trafficking of T cells into tumours 5. Infiltration of T cells into tumours 6. Recognition of cancer cells by T cells

169

CD28:B7.1, CD137 (4-1BB)/CD137L, OX40:OX40L, CD27:CD70, HVEM, GITR, IL-2, IL-12 CX3CL1, CXCL9, CXCL10, CCL5

CTLA4:B7.1, PDL1:PD-1, PDL1:B7.1, prostaglandins -

LFA1:ICAM1, selectins

VEGF, endothelin B receptor

IFN-Ȗ T cell granule content

PD-L1:PD-1, PDL1:B7.1, TIM3:phospholipids, BTLA, VISTA, LAG-3, IDO, Arginase, MICA:MICB, B7H4, 7*)ȕ

2.3 Immunosuppressive tumour microenvironment The tumour microenvironment (TME) provides the required habitat for the tumour cells to grow and adapt to the surroundings, thus preventing them from being detected and removed by host immunosurveillance [17]. The tumour cells can also facilitate an advanced invasion ability in the TME [18]. Evidence has confirmed that the cancer cells must collaborate with the surrounding normal cells and reprogram them [19]. They interact by creating an organ-like structure between the tumour cells and normal cells, attempting for a local invasion, fast proliferation and metastases [20]. Several normal cells are available in the TME, comprising infiltrating inflammatory cells, carcinoma-associated fibroblasts, bone marrow-derived haematopoietic and endothelial progenitor cells [17]. These cells are recruited by the cancer cells via the secretion of stimulatory growth factors, cytokines and chemokines, which eventually cause these recruited cells to elicit growth-promoting signals, intermediate metabolites and alter the tissue structure to manufacture and strengthen the immunosuppressive TME [20]. Tumour-associated macrophages (TAMs), regulatory T cells (Tregs) and myeloid-derived suppressive cells (MDSCs) are some of the immunesuppressive cells that are attracted to the TME [21]. A clinical survey suggested that the leading risk factor for tumour formation is chronic inflammation because the excessive stimulation of inflammatory mediators

170

Chapter 6

can contribute to cancer formation and the spreading of the cancer cells to other parts of the body [18]. One of the strategies for cancer immunotherapy is targeting the TME by delivering various immunomodulatory substances [22] to reverse the effects of immunosuppressive conditions [21].

2.4 Cancer immunotherapeutic agents Many immunotherapies aimed to recognise and kill cancer cells by controlling the immune system have been introduced in the fight against cancer [23]. The therapy works by protecting the host and its features in specificity and memory that can target the specific carcinogen/pathogen by quickly responding to them [24]. There are two approaches to cancer immunotherapy. The first approach is where agents are used for targeting the tumour directly. In contrast, the second approach is the use of agents that can stimulate immune cells [25]. Different types of cancer immunotherapy agents are currently available in the arsenal of oncologists (Figure 6-2).

Figure 6-2: Types of cancer immunotherapy. Created with BioRender.com

Emerging Nanotechnologies for Cancer Immunotherapy

171

3. Immunotherapy agents used in tumour targeting 3.1 Monoclonal antibodies, radioimmunotherapy, antibodydrug conjugates and immunotoxins Monoclonal antibodies are one of the immunotherapy agents developed to directly target the tumour [26]. Rituximab, the first monoclonal antibody, was accepted in 1997 by US Food and Drug Administration (FDA). Since then, numerous monoclonal antibodies have been approved for many cancer treatments, such as B-cell malignancies, colon cancer and breast cancer [25]. Two types of monoclonal antibodies, namely naked and conjugated monoclonal antibodies, have been produced to coordinate the mechanisms of disease progression in cancer formation. Naked monoclonal antibodies have no material such as drugs or radioactive particles attached to them. In contrast, those comprising a chemotherapeutic drug, toxin or radioactive isotope are conjugated monoclonal antibodies [27]. Monoclonal antibodies can promote tumour eradication by various mechanisms, such as executing cells directly by apoptosis induction, distribution of cytotoxic agents, receptor blockade or the agonist activity, radiation, immune-mediated cell eradication, or acting on the tumour vasculature and stroma [28]. Tumour cell death by naked monoclonal antibodies is caused by fragment crystallisable (Fc)-dependent and independent mechanisms [25]. Fcdependent mechanisms are antibody-dependent cell-mediated cytotoxicity (ADCC), which is mediated by natural killer cells and macrophages, complement-dependent cytotoxicity (CDC) and antibody-dependent cellular phagocytosis (ADCP) mediated by macrophages. However, the Fcindependent mechanism directly involves the induction of apoptosis after the antibody binds to its receptor or by obstructing the interactions between the receptor and ligand [25]. However, for the conjugated monoclonal antibodies, the mechanism is different from that of naked monoclonal antibodies, where antibody-tumour antigen interaction can assist in the delivery of toxins to tumour cells, eventually causing apoptosis or promoting apoptosis-targeted radioimmunotherapy [27]. Additionally, monoclonal antibodies were conjugated to radioisotopes, such as Iodine131 or Yttrium-90, to yield radioimmunotherapy agents or conjugated with cytotoxic molecules auristatin E or emtansine to produce antibody-drug conjugates further to improve the efficacy of monoclonal antibodies [25]. Because tumour-associated monoclonal antibodies coupled to cytotoxic radionuclides may bind selectively to tumour antigens and unleash targeted cytotoxic radiation, radioimmunotherapy using radionuclide-labelled monoclonal antibodies is used in cancer therapy [29]. The use of immunotoxins

172

Chapter 6

to target the tumour is also one of the approaches used in developing cancer immunotherapy, where an immune molecule such as a cytokine is conjugated to a toxin. The receptor on the tumour cell’s surface will bind to its substrate, namely cytokine, and distribute the toxin into the cell via receptor-mediated endocytosis [25]. Table 6-2 shows some monoclonal antibodies approved by the FDA. Table 6-2 Food and Drug Administration (FDA) approved anticancer immunotherapy [20]. Drug Name

Immunotherapy

Cancer Type

Rituximab

Monoclonal Antibody: CD20

Trastuzumab

Monoclonal Antibody: Erb B2 (HER-2) Monoclonal antibody: CD274

CD20-positive B-cell NonHodgkin’s lymphoma Metastatic breast cancer

Durvalumab

Alemtuzumab

Monoclonal Antibody: CD52

Ibritumomab

Monoclonal Antibody: CD20

Tositumomab

Monoclonal Antibody: CD20

Cetuximab

Monoclonal Antibody: EGFR

locally advanced or metastatic urothelial carcinoma B-cell Chronic lymphocytic leukaemia B-cell nonHodgkin’s lymphoma CD20-positive B-cell NonHodgkin’s lymphoma Metastatic colorectal and head and neck carcinoma

FDA approval time 26/11/1997

25/09/1998 01/05/2017

07/05/2001 19/02/2002 27/06/2003

12/02/2004

Emerging Nanotechnologies for Cancer Immunotherapy

Bevacizumab

Monoclonal Antibody: VEGF-A

Panitumumab

Monoclonal Antibody: EGFR

Sipuleucel-T

Cancer vaccine

Talimogene laherparepvec Ipilimumab

Oncolytic virus therapy: HSV-1 Immune checkpoint inhibitors: antiCTLA-4 Immune checkpoint inhibitors: antiPD-1

Pembrolizumab

Nivolumab

Atezolizumab

Immune checkpoint inhibitors: antiPD-1

Immune checkpoint inhibitors: antiPD-1

173

Metastatic colorectal and non-small cell lung carcinoma Metastatic colorectal carcinoma Metastatic castrationresistant prostate cancer Advanced melanoma Advanced melanoma

26/02/2004

Advanced refractory melanoma and non-small cell lung cancer Unresectable or metastatic melanoma Squamous nonsmall cell lung cancer classical Hodgkin lymphoma (cHL) Urothelial carcinoma metastatic nonsmall cell lung cancer (NSCLC)

04/09/2014

27/09/2006 29/04/2010

27/10/2015 25/03/2011

22/12/2014

17/05/2016

18/05/2016 18/10/2016

174

Chapter 6

4. Immunotherapy agents activating immune cells 4.1 Immune checkpoint inhibitors Immune checkpoint inhibitors offer a promising treatment for different cancers because they offer a remarkable and lasting treatment response in some patients [30]. The mechanism of action for immune checkpoint inhibitors is directed towards suppressing the immune system caused by cancer [4]. This action occurs at the immune checkpoints where tumours can inhibit T-cell activity and the body’s immune system [31]. Thus, immune checkpoint inhibitors play a crucial role in blocking T-cell checkpoint receptors, thus giving long-term antitumour responses [4]. The use of immunomodulatory monoclonal antibodies is to block immune checkpoint receptors and stop the checkpoint ligand-receptor interactions [32]. Some immunomodulatory monoclonal antibodies that inhibit immune checkpoint receptors are cytotoxic T lymphocyte-associated antigen 4 (CTLA-4) and PD-1 or IC ligands, such as PD-L1 [33]. They play an essential function in regulating the immune response. CTLA-4 is crucial for the activation of T cells in the lymphatic tissue. In contrast, PD-1 modulates the activation of T cells in peripheral tissues that include the TME, resulting in the downregulation of T-cell effectors and cell death [34]. When CTLA4 is manifested on the T-cell surface, there is a competition between CTLA4 and the CD28, the stimulatory receptor for binding to the ligands, namely CD80/CD86 [35]. Because CTLA-4 has a greater affinity for binding to B7 than CD28, the binding to B7 prevents the costimulation of T cell-mediated by CD28, inhibiting T cell activation [36]. It is also essential for the function of regulatory T cells (Tregs), which rely on each other to sustain their activity in suppressing the immune response because the deficiency of CTLA-4 is associated with the development of profound systemic autoimmune diseases [37]. The success of Ipilimumab, an anti-CTLA-4 monoclonal antibody, resulted in its approval in 2015 by the FDA for its use in advanced melanoma [37-38]. The PD-1 receptor is expressed on dendritic cells (DCs), natural killer cells (NK), B cells, T cells, macrophages and monocytes [39]. It binds to two types of ligands: PD-L1 and PD-L2. PD-L1 is expressed on various cells, such as epithelium, T and B cells, muscle mesenchymal stem cells, cancer cells, macrophages and dendritic cells. In contrast, PD-L2 is expressed in immune-related cells, such as mast cells, dendritic cells and macrophages [40]. When PD-1 and its ligand PD-L1 interact in normal conditions,

Emerging Nanotechnologies for Cancer Immunotherapy

175

cytotoxic activity will be downregulated to maintain immune homeostasis. Some agents that inhibit either PD-1 or PD-L1 are under clinical investigation. These agents are two monoclonal antibodies, namely nivolumab and pembrolizumab, approved in 2014 by the FDA for advanced or unresectable melanoma therapy [41]. A study was conducted to address the issue of offtarget toxicity induced by CTLA-4 and PD-1 antibodies in their standard delivery. One way to solve the issue was to administer anti-CTLA-4 intratumorally with mesoporous silica nanoparticles [42]. This administration was said to have a high loading efficiency and increased antitumour efficacy compared to intraperitoneal administration of the soluble antibody in a murine melanoma model. This is most likely attributed to the controlled release of antibodies from an in situ depot [43].

5. Adoptive cell transfer Adoptive cell transfer is a certain type of cell-based anticancer immunotherapy that usually comprises: 1. the pool of circulating or tumour-infiltrating lymphocytes; 2. the selection/modification/expansion/activation ex vivo; and 3. the re-administration to patients after the destruction of lymphocytes and T cells caused by irradiation before starting immunotherapy (pre-conditioning lymphodepletion) and in combination with immunostimulatory agents [44]. The infusion of T cells can be either allogeneic or autologous [45]. Allogenic haematopoietic stem cell transplantation (HSCT) triggered an immune response based on allogeneic variations in peptide/HLA complex expression or minor histocompatibility antigens when infused into cancer patients from a healthy donor. Antitumour activity was observed in patients with haematological malignancies after the infusion of genetically modified autologous or allogeneic T cells [4]. There are various kinds of adoptive cell therapy, such as culturing tumourinfiltrating lymphocytes, wherein the tumour is directly obtained, isolated, and one specific T-cell or clone is expanded [46]. Another technique includes using T cells engineered in vitro to recognise and attack tumours. This technique is termed chimeric antigen receptor T-cell (CAR T-cell) therapy [47]. Additionally, recognising tumour by the chimeric antigen receptor is not compromised by the changes in the tumour MHC expression or processing of epitope or by TCR complex dissociation on responding T cells. The affinity for the interaction between the antibody and antigen is

176

Chapter 6

higher than the natural binding of the T-cell receptor (TCR) with MHC/peptide. The benefit of using CAR T-cell therapy is that they do not depend on the binding with MHC/peptide. Hence, they are not restricted by a patients’ HLA type [48]. The earliest approval for the disease using a CAR T-cell therapy was for young adult and paediatric patients with a relapsed and refractory B-cell acute lymphoblastic leukaemia. The first few CAR Tcell therapies approved by FDA are axicabtagene ciloleucel and tisagenlecleucel [49]. Adjuvant drug-loaded nanoparticles are used to modify the therapeutic cells’ surfaces to improve the efficacy of adoptive cell transfer. For example, the surfaces of CD8+ T cells can be conjugated with maleimidemodified synthetic nanoparticles via sulfhydryl groups exposed by cell surface proteins [50]. Nanoparticles encapsulated in cytokines that stimulate T cells, such as IL-15/IL-15Ra and IL-21, can also signal T-cell proliferation in situ. This technique has resulted in the proliferation of potent transferred T cells. It also eradicated tumours of metastatic melanoma in mice, whereas the co-administration of T cells with cytokines alone showed no eradication of tumours [43].

6. Cancer vaccines Vaccines in cancer therapy aim to enhance the patients’ antitumour T cells’ ability against tumour antigens [51]. Antigens must be selected appropriately to develop an effective cancer vaccine. New strategies should also be formulated to combat immune evasion and suppression mechanisms, provoking robust effector and memory T-cell responses [28]. Cancer vaccines can comprise peptides, tumour cells, immune cells or dendritic cells. Cancer vaccines may contain complete cancer cells, a fraction of cancer cells or purified antigens that can amplify the immune response against cancer cells [4].

6.1 Peptide-based vaccines Peptide vaccines stimulate an immune response against a single tumour antigen, which is expressed in association with HLA molecules on the surface of tumour cells [4]. Peptide-based vaccines usually target dendritic cells and other antigen-presenting cells (APCs). It works by administering full-length recombinant tumour-associated antigen (TAA) or peptides to patients with cancer via intramuscular, intradermal or subcutaneous routes together with one or more immunostimulatory agents, generally known as

Emerging Nanotechnologies for Cancer Immunotherapy

177

adjuvants. The adjuvant is normally used to promote the maturation of dendritic cells [44]. The mechanism of a peptide-based vaccine is that when a TAA is injected into patients with cancer, it will bind with the restricted MHC molecule that is expressed in APCs. After intracellular processing, the peptide/MHC complex is delivered to the cell surface. This delivery is then recognised by the TCR on the surface T cells, eventually causing the stimulation of T lymphocytes. Thus, a specific immune response is elicited against tumours when a peptide cancer vaccine is used [52].

6.2 Tumour cell-based vaccines Tumour cell-based vaccines use the whole tumour cells to provide immunogenic material sources. HLA-type restrictions do not regulate these vaccines compared to peptide-based vaccines. These vaccines can be used to exhibit a wide range of defence. Tumour cell-based vaccines are presented either in autologous or allogeneic methods. Autologous means that the tumour cells are used from the person who receives the vaccine, whereas allogeneic is where the tumour cells are used from another patient. After the tumour cells are taken, irradiation is performed for the immunisation of the cells, followed by administering the vaccine either in combination with an adjuvant such as granulocyte-macrophage colony-stimulating factor alone. One example of a tumour cell-based vaccine is M-vax (AVAX Technologies), which was proven effective in targeting melanoma, renal cell carcinoma and acute myeloid leukaemia in clinical trials [4].

6.3 Immune or dendritic cell-based vaccine Dendritic cells play an essential role in developing innate and adaptive immunity. They are referred to as essential targets for immunotherapy. It is one of the subsets of APCs that work by phagocytosing the antigens from their environment and is involved in damage-associated molecular patterns (DAMPs). When APCs are stimulated, they will deliver peptide antigens on MHCs to other immune cells, causing the secretion of effector cytokine (CD4+ T cell) and cytotoxic T lymphocyte (CTL) responses (CD8+ T cell) that are responsible for antitumour activity [21]. Dendritic cells can be used for vaccination against cancer by several means: 1) capturing of nontargeted peptide/protein and nucleic acids-based vaccines in vivo; 2) direct coupling of vaccines that comprises antigens to anti-dendritic cells antibodies; or 3) vaccines comprising ex vivo–generated dendritic cells that are loaded with antigens [53].

178

Chapter 6

Sipuleucel-T, which the FDA approved in 2011, demonstrated an improvement in the survival of patients with prostate cancer following a phase III trial. It was the earliest dendritic cell-based vaccine that received the first approval in cancer therapy [54]. It is prepared from a patient’s own (autologous) peripheral blood mononuclear cells separated by leukapheresis. Then, it is delivered to a central facility where the cell preparation is enriched for dendritic cells by gradient centrifugation. Following that, the dendritic cells are co-cultured for 36–44 hours with PA2024, which is a recombinant fusion protein comprising prostatic acid phosphatase (PAP) and granulocyte-macrophage colony-stimulating factor (GM-CSF) [55]. Finally, the cells are harvested and brought to the patient's treatment facility for administration by intravenous infusion [56].

7. Nanoparticles in cancer therapy Nanotechnology has witnessed rapid development because it resolves the problems linked to conventional drug therapeutics, including low therapeutic index, lack of targeting capability, water-solubility, systemic toxicity, and non-specific distribution [57]. Nanoparticles are generally in the size range of 1–100 nanometres. They can be utilised in a targeted drug delivery approach that specifically targets a drug or drug carrier to reduce systemic toxic effects [58]. Additionally, they can support therapeutic agents to pass through biological barriers, facilitate molecular interactions and identify changes in the molecules [57]. After binding nanoparticles to the receptors, they can undergo receptormediated endocytosis or phagocytosis by the cells, leading to the release of the encapsulated drug to the inside of the cell cytoplasm. Nanoparticles offer a greater surface area wherein their properties, such as electronic, magnetic and biological properties, can be modified compared to microparticles. Some existing nanotechnology-based drug delivery systems, including liposomes, dendrimers, polymeric micelles, nanospheres, nanotubes and nanocapsules, are available for cancer treatment. These methods are already marketed and/or under research and evaluation [3]. DOXIL (liposomal doxorubicin) and Abraxane (albumin-bound paclitaxel) are examples of nanotechnology-based formulations in use [59]. Various types of nanoparticles are combined with the immunotherapeutic agents to effectively traverse the biological membrane for destroying the cancer cells (Figure 6-3).

Emerging Nanotechnologies for Cancer Immunotherapy

Figure 6-3: Nanoparticles in immunotherapy Created with BioRender.com

179

180

Chapter 6

7.1 Liposomes Liposomes comprise a phospholipid bilayer surface that is surrounded by an aqueous core. They are used as drug carriers in gene therapy and the targeted delivery of natural or synthetic chemotherapeutics. The phospholipid bilayer of liposomes forms sealed spherical vesicles and drugs, either hydrophobic or hydrophilic. It can be encapsulated by cell internalisation or diffusion [60]. It has been used clinically to improve drug delivery to the tumour sites, reduce the side effects caused by chemotherapy or antimicrobial therapies, and improve specificity to the site of injuries [5]. The delivery of therapeutic agents to the target is made possible by encapsulating drugs into liposomes and preventing uptake by the reticuloendothelial system. The liposomal encapsulation of doxorubicin has positively changed both pharmacology and pharmacokinetics, resulting in enhanced drug delivery to the tumour sites and reduced side effects compared to conventional treatments [61]. Doxorubicin, a PEGylated liposomal formulation, has been proven to be successful. It is used in head and neck cancer, metastatic breast cancer, and ovarian cancer [62]. However, attempts of using liposome-based immunotherapy have not been successful. However, the macrophages activation by liposome-enclosing immunomodulators in mice having metastatic tumours has shown a positive effect in inhibiting tumour growth at the primary site and in metastasis [63].

7.2 Dendrimers Dendrimers are globular, nanometric, hyperbranched and monodisperse polymer carriers, with specific molecular weight, size, shape and properties of the host-guest entrapment. Dendrimers with a molecular weight greater than 40 KDa remained in the blood longer than polymers with lower molecular weight [62]. Dendrimers can interact with cell membranes, organelles and proteins. The cationic surface of dendrimers and cell lipid bilayer interaction will facilitate an enhanced permeability and reduce the biological membrane integrity [64]. The mechanism causing the leakage of cytosol proteins can be determined by identifying the interaction between dendrimers and the cell membrane. The permeability and the intracellular delivery of agents can also be facilitated by an interaction between the cationic charge of the end groups and the anionic charge of the membrane. The high-charge density of cationic dendrimers interacting with the membrane may cause the membrane’s integrity and leak critical intracellular components, resulting in toxicity and cell death [5].

Emerging Nanotechnologies for Cancer Immunotherapy

181

The interaction of dendrimers with distinct drug classes can be made possible by forming physical and chemical bonds, which can be used to integrate hydrophobic or hydrophilic molecules inside their vacant cavities through non-bonding interactions. Another alternative is when a complex can be formed by attaching the drug molecule to its periphery due to the electrostatic interactions or conjugation between the drug and the dendrimers [5]. The properties of polyamidoamine (PAMAM), such as biocompatibility, structural control, functionality and ease of synthesis, make it suitable in in nanotechnology and immunotherapy [21]. The conjugation of hydroxyl-terminated G4 PAMAM with paclitaxel, a chemotherapy drug, through a union with succinic acid, exhibited a significant rise in the anticancer activity compared to the free drug [65]. Immunodendrimers were used to treat ovarian cancer, where they were conjugated with monoclonal antibody K1 (mAbK1) to the half-generation polypropylene imine (PPI) dendrimers. Moreover, paclitaxel (PTX) was encapsulated into the hydrophobic nanodomains of PPI dendrimers. The mAbK1-PPI-PTX treatment in the ovarian cancer model reduced the tumour volume, and the animal’s survival was extended. This survival was accomplished through enhanced drug uptake in the tumour [66].

7.3 Gold nanoparticles Gold nanoparticles have become the focus of biomedical research because of their physical and chemical properties, such as shape, carrier capabilities, surface area and biocompatibility. However, these nanoparticles also have several disadvantages, such as low efficiency in encapsulation, poor storage stability and slow endosomal escape, limiting the use [5]. Gold nanoparticles are presently discovered as potential agents for drug delivery in incorporating drugs into the tumour cells. These metal nanoparticles were supposed to be conjugated with surface ligands that can direct them to tumour cells to spare the healthy cells while destroying the cancer cells. The conjugation of gold nanoparticles to PEG and the conjugation with specific antibodies that bind biomarkers expressed on tumour cells were portrayed by Huang et al. (2008) as the two methods of choice for targeting the tumour cells [67]. An example is the CYT-6091, 27-nm citrate-coated gold nanoparticles with thiolated PEG and tumour necrosis factor–Į (TNF-Į). Gold nanoparticles were the first nanoparticle therapy that entered an early phase of the clinical trials and revealed the dual effects of enhancing tumour targeting and toxicity [68]. Moreover, gold nanoparticles can also be used in cancer vaccines to efficiently deliver antigens to dendritic cells for cancer therapy because of their easy size control [69].

182

Chapter 6

8. Nanotechnology to enhance cancer immunotherapy 8.1 Nanoparticle-based therapeutic cancer vaccines Cancer vaccines aim to co-deliver the tumour-associated antigens, immune adjuvants, such as CpG, monophosphoryl lipid A or polyinosinic: polycytidylic acid [Poly(i:C)] and deliver neoantigens to professional APCs to trigger an immune response. Because the current vaccine strategies are less effective, the delivery of vaccines is also ineffective. Additionally, vaccines administered intramuscularly or subcutaneously present a challenge: they can spread to peripheral blood vessels because of their large size, causing them not to be transmitted effectively to lymphoid organs. Thus, using nanoparticles ranging between 20 and 200 nm for vaccine delivery will enhance their delivery of antigen to the APCs, producing immune responses to immune cells because they can drain freely to the lymphatic nodes [70-71]. One way to improve the target-specific delivery of nanocarrier is by targeting moieties. A therapeutic vaccine usually comprises an antigen and an adjuvant [22]. Luis et al. demonstrated that targeting antigens and adjuvants to dendritic cells plays an essential role in efficiently activating dendritic cells and producing a strong cytotoxic T-cell response. Their study showed that PEG-coated PLGA nanoparticles were encapsulated with ovalbumin, the modal antigen, toll-like receptor (TLR) ligand, and polyinosinic:polycytidylic acid combination [Poly(i:C)]. Subsequently, the PEGylated PLGA nanoparticles were conjugated with monoclonal antibodies against the surface of dendritic cells to increase the internalisation of nanoparticles. Using dendritic cell-targeted PLGA nanoparticles, subcutaneous vaccination in mice, produced a more potent CD8+ T-cell response. Nanoparticle vaccines that target dendritic cells through functionalisation with CD40 produce efficient CD8+ T-cell responses specific to the antigens [70]. An example of the nanoparticle-based adjuvant is short, single-stranded DNA molecules, CpG oligonucleotides. They can stimulate dendritic cells by binding the TLR-9 within the phagosome. The CpG oligonucleotides are coated with antigen-loaded nanoparticles, such as cationic gelatin-based nanoparticles, generating an antigen-specific T-cell response and providing a protective antitumoral immunity in a murine model of melanoma [21]. Nanoparticles in delivering antigens and adjuvants can prevent degradation and efficiently facilitate the targeted dendritic cell uptake. The use of a nanovaccine could also induce the activation of dendritic cells and initiate

Emerging Nanotechnologies for Cancer Immunotherapy

183

cellular and humoral immunity in vivo, thus indicating a great potential in augmenting the immunogenicity of cancer vaccines and leading to a stronger T-cell response [22].

8.2 Nanoparticles for T-cell activation Artificial APCs (aAPCs), a promising technology, have been developed for cancer immunotherapy [72]. The artificial antigen-presenting cells (aAPCs) that are synthesised using microparticles or nanoparticles can function in the same manner similar to the natural APCs in the activation of T cells’ responses [73]. Nanoparticle-based aAPCs bind to MHC loaded with specific peptides and a ligand or activating antibody for CD28 to nanoparticles such as iron oxide or quantum dots. These nanoparticle-based aAPCs can trigger antigen-specific T-cell proliferation by activating the T-cell receptor (TCR) and the co-stimulatory-stimulatory receptor from peripheral blood T cells. A study has demonstrated iron oxide-based nanoparticle APCs as a strategy to improve T-cell activation further [70]. The coupling of maleimidefunctionalised lipid nanoparticles to the T-cell surface before adoptive transfer to allow the direct delivery of phosphatase antagonist to the synapse of T cells, causing an intense T-cell expansion at the tumour site enhanced the survival of mice with prostate cancer [74].

8.3 Nanoparticles for the modulation of the tumour microenvironment The tumour causes the formation of an immunosuppressive TME because it can promote the growth of cancer cells [74]. Thus, tumour microenvironment modulation provides a significant approach to cancer immunotherapy [75]. Immunosuppressive T cells, such as Tregs, can hinder the activity of antitumour T-effector cells [76]. It inhibits immune cells in the tumour microenvironment and deteriorates the activity of anticancer immune response. An antitumour immunity can be provoked by suppressing or eliminating the Tregs. One standard method in cancer immunotherapy is manipulating the function of Treg by using a checkpoint blockade such as anti-CTLA-4 [72].

Chapter 6

Antigen-loaded nanoparticles (NPs)

Types

OVA-specific CTL response led to a reduction in tumour burden in E.G7-OVA tumour-bearing mouse mode.

Polymer-modified OVA-loaded liposomes, which becomes highly unstable below pH 6, where OVA can be released directly in endosomes.

Whole-cell lysate derived from patients with head and neck squamous cell carcinoma encapsulated in PLGA NP

Carcinoembryonic antigen (CEA) conjugated to inorganic iron oxide-zinc oxide (Fe3 O4–ZnO) core-shell NPs.

Outcome of study OVA-PMASH hydrogel internalisation by mouse APC resulted in OVA-specific CD4+ and CD8+ T cells. Following the intravascular vaccination of mice, CD4+ and CD8+ T cells showed 6-fold and 70-fold higher activation, respectively, compared with an equivalent amount of OVA administered alone. These NPs enabled real-time monitoring by magnetic resonance imaging. Mice immunised with NP-Ag-treated DCs demonstrated enhanced cytotoxic T lymphocyte-mediated responses, thereby delaying tumour growth and increasing survival rates. Ag-loaded NP delivered to patient-derived DCs led to stimulation of CD8+ T cells. Ќ ,)1Ȗ and Ў IL10 were observed in 80% of patients.

Nanocarrier system Layer-by-layer-assembled disulphide crosslinked poly(methacrylic acid (PMASH) hydrogel encapsulating ovalbumin (OVA)

Table 6-3 Different approaches of nanoparticle-based immunotherapy [79].

184

Antibody/ligand-coated NP for active targeting

Adjuvant/immunostimulantcoated NPs

Cytokine-loaded NPs

NPs (developed from polyethyleneimine and C32 SRO\ ȕ-amino ester)) encasing TLR agonists (CpG or poly I:C) and a plasmid (pSP-D-CD40L)-expressing CD40 ligand. OVA encapsulated PLGA NPs with lipidPEG complexed with humanised targeting Ab hD1 (DC-restricted CLRDC-SIGN). NPs were encapsulated with poly (I:C) and resiquimod (R848) as adjuvants.

Immunostimulatory peptide (Hp91) derived from an endogenous protein (HMGB1) encapsulated in or conjugated on the surface of PLGA NPs.

Polybutyl cyanoacrylate (PBCA) NPs loading TGF-ȕ antisense oligodeoxynucleotide (ODN) used to treat glioblastoma brain tumour in Fischer rats Ultrasmall gold NPs conjugated with CpG ODN, 15 nm in diameter, compared with administration of CpG alone.

IL-1-loaded NPs infused onto T cells ex vivo and reintroduced in mouse tumours.

185

DC-specific antibody/ligand-coated carriers achieved active targeting of DCs. NP induced CTL responses at a 100-fold lower dose of adjuvant than that administered in soluble form in mice.

Enhanced CpG macrophage stimulation High infiltration of DCs and macrophages at the tumour site, resulting in tumour inhibition and prolonged survival in mice. When encapsulated in or conjugated on the surface of NP, Hp-91 was found to be 5-fold and 20-fold more potent, respectively, than in the free form. Due to their DC-activating potential, Hp-91-NPs are promising delivery vehicles for the treatment of cancer. CD40L and TLR agonists act synergistically, resulting in tumour-free survival in NP-treated groups versus control

Ќ T-cell proliferation and survival within the tumour, amplifying the antitumour response as compared with a systemic cytokine administration. Ў TGF-ȕ levels, Ќ activated CD25+ T cells. Survival rates are higher in NP-immunised rats than in untreated rats.

Emerging Nanotechnologies for Cancer Immunotherapy

Adjuvant-Antigen codelivery with particle-based carriers

NPs with a combination of immunomodulator and drugs

186

Coencapsulation of monophosphoryl lipid A (MPLA) with either Ag-OVA or BLP25.

OVA and poly (I:C) conjugated with CTAB and encapsulated in pH-sensitive polyketide (PK3) microparticles.

OVA and poly (I:C) or CpG coadministered in PLGA microspheres

Doxorubicin-loaded NPs were developed using an immunotherapeutic self-organising acetylated fucoidan (AcFu) polymer. OVA and poly (I:C) or CpG coadministered in microspheres compared with incomplete Freund’s adjuvant

pleomorphic lipopolysaccharide (P-LPS) and Paclitaxel (PTX) encapsulated in PLGA NPs.

Chapter 6

Single vaccination in mouse models of EG-7 thymomas and MO-5 melanomas resulted in high titers of OVA-specific IgG1 and IgG2a and CTLmediated killing for up to 21 days postimmunisation Eightfold higher IFN-Ȗ CD8+ T-cell than control in melanoma mice models following a single immunisation. Secretion of IL-2 by CD8+ T cells enhanced by more than 6-fold. 30% and 25% higher secretion of cytokines TNF-Į and IFN-Ȗ respectively, as compared with the control group. Maturation of DCs was enhanced when induced by MPLA in PLGA NP compared with induction in a soluble form. MPLA coadministered with Ag led to Ќpro-inflammatory IL-6, IL-12, and TNF-Ș cytokine expression, and Ў IL-13 and IL-4 expression

The mean tumour volume of TLNP-treated mice was found to be 40% less than in animals treated with PTX and P-LPS alone. Higher infiltration of APCs (macrophages and DCs) and T cells was observed in the tumour microenvironment. AcFu-NP-potentiated secretion of TNF-Į and GMCSF in Raw264.7 macrophages

Gold NPs coated with RFP (as model Ag) and CpG were evaluated in RFP-expressing melanoma tumour models.

Acid-degradable hydrogel encapsulating OVA and CpG

187

Ќ Ag-specific T cells and greater survivability of mice upon therapeutic immunisation. Ag-specific T cells demonstrated 20% higher efficacy in lysing target cells in the OVA-CpG-NP treatment groups as against OVA-NP administered with CpG in soluble form Lymph node targeting resulted in the interaction of NPs with DC inducing a potent CTL response and Th1-driven Ab secretion with a significant antitumour response.

Emerging Nanotechnologies for Cancer Immunotherapy

188

Chapter 6

Several therapeutic approaches utilise nanoparticles to inhibit the growth of the tumour and modulate the immune cells in the tumour microenvironment [72]. One of the approaches is to use nanoparticles that can target the costimulatory receptors, such as the glucocorticoid-induced tumour necrosis factor receptor (GITR). GITR is expressed mostly by CD4+ Tregs. he stimulation of GITR has been presented to increase antitumour activity by diminishing the Treg lineage's stability within tumours, thus decreasing the suppression of the immune system [77]. An in vivo experiment demonstrated that the selective internalisation of polyethylene glycol–modified singlewalled carbon nanotubes (PEG–SWCNTs) with GITR ligands were internalised by Treg via receptor-mediated endocytosis and delivered into the cytoplasm/nucleus in vivo, hence providing innovative immunotherapy against cancer [78].

9. Application of nanotechnology in cancer targeting The design of nanoparticles can help target the desired cells via different modifications such as altering the chemical and physical properties, shape and size. The nanotechnology-based targeted delivery system comprises three critical elements: (i) an apoptosis-inducing agent (anti-cancer drug), (ii) a targeting moiety-penetration enhancer, and (iii) a carrier. Nanotechnology can target cancer cells through active or passive targeting [3].

9.1 Active targeting Active targeting is defined as the interactions between the drug-drug carrier and the target cells, which is frequently accomplished by interactions of the specific ligand-receptor or by recognising antibody-antigen for the intracellular localisation of the drug [80]. Cancer cells possess unique features that can be easily distinguished from healthy cells at the molecular level. The binding of the complementary ligands on the surface of the nanoparticles enables them to target the cancer cells specifically. Once they bind with receptors, receptor-mediated endocytosis or phagocytosis by cells occurs rapidly and led to the cell internalisation of the encapsulated drug [81]. This approach is most useful for a controlled drug release where the drug is either released into the intracellular or extracellular compartment [61]. Some examples of ligands are proteins, nucleic acids, small molecules such as vitamins, peptides, immune checkpoints, cytokines and antibodies. In contrast, the target molecules comprise lipids, sugars or proteins that can be

Emerging Nanotechnologies for Cancer Immunotherapy

189

found in organs affected by disease or on the cell surface. The two main aspects for evaluating the effectiveness of an active targeting system are the delivering capacity and the targeting specificity. The delivery capacity is associated directly with the material of nanoparticles and their structure [81]. In contrast, the ligand-functionalised nanoparticle biodistribution regulates the targeting specificity and determines how the nanoparticle system and conjugated ligand interact with the cells and off-target molecules. Some examples that use this active immunotherapy are anticancer vaccines that include peptides, dendritic cell-based and allogeneic whole-cell vaccines, oncolytic viruses and immune checkpoint inhibitors [82]. Strategies using nanotechnology-based therapies, such as cancer vaccines, have been discovered for drug delivery. These therapies can overcome biological barriers and modulate the intracellular trafficking of therapeutic payload. These nanoparticles-based systems display a better possibility for the site-selective delivery through the binding between the nanoparticle surface and recognition ligands that can improve nanoparticle endocytosis, thus affecting their intracellular trafficking and resulting in prolonged effects [83].

9.2 Passive targeting Passive targeting refers to drug accumulation in the areas around the tumour with leaky vasculature. It is also known as the enhanced permeation and retention effect [68]. Because cancer cells start to undergo apoptosis, it causes them to draw nutritious agents aberrantly through the blood vessels, leading to the formation of broad and leaky blood vessels around the cells by angiogenesis [81]. The development of leaky blood vessels is attributed to the apparent abnormalities of the basement membrane and the descending number of pericytes lining the rapidly proliferating endothelial cells. Hence, there is an increase in permeability of molecules to pass through the vessel wall into the interstitium surrounding tumour cells. Emerging nanoparticles make it easy to target certain parts of the capillary endothelium and accumulate the drug within a specific organ to eradicate the tumour cells by convection or passive diffusion [3]. For nanoparticles to pass through the pores easily, the size must be lesser than the pores (200–780nm) [3]. The lymphatic vessels in normal tissues are drained by the constant flow of extracellular fluid, allowing the renewal and constant draining of interstitial fluid and the recycling of solutes and colloids back into the circulation. However, because the tumour’s lymphatic system is malfunctioning, only a minimal uptake of interstitial fluid is

190

Chapter 6

accomplished. Thus, the colloids cannot return to circulation using convective forces. Molecules smaller than 4 nm can still return to the circulation and be reabsorbed. However, macromolecules and nanoparticles cannot diffuse back into the circulation due to their greater hydrodynamic radii. Therefore, nanoparticles that have spread to the perivascular space will not be cleared effectively, causing them to accumulate in the tumour interstitium [81]. Because of the characteristics of enhanced permeability and retention effect that the tumour cells possess, the targeted delivery of anticancer agents is easily permitted. [57]. A few examples of passive immunotherapy include transfusing cytokines, monoclonal antibodies and adoptive cell transfer of immune cells pre-activated for tumour lysis [82].

10. Conclusion and future perspectives Although chemotherapy and radiotherapy have improved cancer treatment, the emergence of immunotherapeutic approaches somehow provides a more favourable method for cancer, especially those using immune cell-mediated therapy [83]. It is critical to understand the immunology of the tumour, particularly the functions of the immunosuppressive tumour microenvironment and tumour antigens in cancer because these methods can help in the successful targeting of cancer immunotherapy. To further enhance immunotherapy delivery, various nanoparticles have been used as a strategy to improve immunotherapeutic efficacy. Nanoparticle-based immunotherapy is believed to have great potential in improving cancer immunotherapy. The emergence of nanoparticle delivery systems has allowed effective vehicles to deliver tumour-associated antigens and immune-stimulatory molecules to APCs and the tumour microenvironment to weaken the signals from the immunosuppressive tumour, leading to improved outcomes. They are much more effective when compared to non-nano particle-based therapeutics [21]. The use of nanoparticles in immunotherapy thus improved the therapy's efficacy compared to non-nanoparticle-based therapy. Hence, the availability of various types of nanoparticles can deliver significant development in the future because they can be used for vehicle delivery and targeting. By modifying their size, shape and surface functionalisation, the delivery can improve pharmacokinetics, biodistribution, stability and toxicity profiles in clinical trials.

Emerging Nanotechnologies for Cancer Immunotherapy

191

References 1.

2. 3. 4. 5.

6. 7.

8.

9.

10.

N. Dan, S. Setua, V. K. Kashyap, S. Khan, M. Jaggi, M. M. Yallapu, & S. C. Chauhan, Antibody-drug conjugates for cancer therapy: Chemistry to clinical implications. Pharmaceuticals (Basel, Switzerland), 11 (2018). https://doi.org/10.3390/ph11020032. NIH, PubMed Health - National Library of Medicine. PubMed Health Glossary, (2015) 1. K. B. Sutradhar & M. L. Amin, Nanotechnology in cancer drug delivery and selective targeting. ISRN Nanotechnology, 2014 (2014) 1–12. https://doi.org/10.1155/2014/939378. C. L. Ventola, Cancer Immunotherapy, Part 1: Current strategies and agents. P & 7 A Peer-reviewed Journal for Formulary Management, 42 (2017) 375–383. A. Jurj, C. Braicu, L.-A. Pop, C. Tomuleasa, C. D. Gherman, & I. Berindan-Neagoe, The new era of nanotechnology, an alternative to change cancer treatment. Drug Design, Development and Therapy, 11 (2017) 2871–2890. https://doi.org/10.2147/DDDT.S142337. T. Saleh & S. A. Shojaosadati, Multifunctional nanoparticles for cancer immunotherapy. Human Vaccines & Immunotherapeutics, 12 (2016) 1863–75. https://doi.org/10.1080/21645515.2016.1147635. H. Deng & Z. Zhang, The application of nanotechnology in immune checkpoint blockade for cancer treatment. Journal of Controlled RHOHDVH Official Journal of the Controlled Release Society, 290 (2018) 28–45. https://doi.org/10.1016/j.jconrel.2018.09.026. A. Konstorum, A. T. Vella, A. J. Adler, & R. C. Laubenbacher, Addressing current challenges in cancer immunotherapy with mathematical and computational modelling. Journal of the Royal Society, Interface, 14 (2017). https://doi.org/10.1098/rsif.2017.0150. T. O. Sullivan, R. Saddawi-konefka, W. Vermi, C. M. Koebel, C. Arthur, J. M. White, R. Uppaluri, D. M. Andrews, S. F. Ngiow, M. W. L. Teng, M. J. Smyth, R. D. Schreiber, & J. D. Bui, Cancer immunoediting by the innate immune system in the absence of adaptive immunity. 209 (2012) 1869–1882. https://doi.org/10.1084/jem.20112738. D. Mittal, M. M. Gubin, R. D. Schreiber, & M. J. Smyth, New insights into cancer immunoediting and its three component phases-elimination, equilibrium and escape. CurrentOopinion in Immunology, 27 (2014) 16–25. https://doi.org/10.1016/j.coi.2014.01.004.

192

11. 12. 13. 14. 15. 16. 17.

18.

19. 20.

21.

22.

Chapter 6

A. Passardi, M. Canale, M. Valgiusti, & P. Ulivi, Immune checkpoints as a target for colorectal cancer treatment. (2017). https://doi.org/10.3390/ijms18061324. G. P. Dunn, L. J. Old, & R. D. Schreiber, The immunobiology of cancer immunosurveillance and immunoediting. Immunity, 21 (2004) 137–48. https://doi.org/10.1016/j.immuni.2004.07.017. M. D. Vesely & R. D. Schreiber, Cancer Immunoediting: Antigens, mechanisms and implications to cancer immunotherapy. 1284 (2014) 1–5. https://doi.org/10.1111/nyas.12105.Cancer. G. P. Dunn, L. J. Old, & R. D. Schreiber, The three Es of cancer immunoediting. Annual Review of Immunology, 22 (2004) 329–360. https://doi.org/10.1146/annurev.immunol.22.012703.104803. M. Aris & M. Barrio, Lessons from cancer immunoediting in cutaneous melanoma. 2012 (2012). https://doi.org/10.1155/2012/192719. D. S. Chen & I. Mellman, Oncology meets immunology: The cancer immunity cycle. Immunity, 39 (2013) 1–10. https://doi.org/10.1016/j.immuni.2013.07.012. J. M. Pitt, A. Marabelle, A. Eggermont, J.-C. Soria, G. Kroemer, & L. Zitvogel, Targeting the tumour microenvironment: removing obstruction to anticancer immune responses and immunotherapy. Annals of Oncology, 27 (2016) 1482–1492. https://doi.org/10.1093/annonc/mdw168. M.-J. Tsai, W.-A. Chang, M.-S. Huang, & P.-L. Kuo, Tumour microenvironment: A new treatment target for cancer. ISRN Biochemistry, 2014 (2014) 1–8. https://doi.org/10.1155/2014/351959. D. Hanahan & L. M. Coussens, Accessories to the crime: Functions of cells recruited to the tumour microenvironment. Cancer Cell, 21 (2012) 309–322. https://doi.org/10.1016/J.CCR.2012.02.022. C. Q.-M. Yuan Yao, Jiang, Yu-Chen Sun, Chong-Kui, Role of the tumour microenvironment in tumour progression and the clinical applications (Review). Oncology Reports, 35 (2016) 2499–2515. https://doi.org/10.3892/or.2016.4660. H. Qiu, Y. Min, Z. Rodgers, L. Zhang, & A. Z. Wang, Nanomedicine approaches to improve cancer immunotherapy. Wiley Interdiscip Rev Nanomed Nanobiotechnol, 9 (2017). https://doi.org/10.1002/wnan.1456. H. Qian, B. Liu, & X. Jiang, Application of nanomaterials in cancer immunotherapy. Materials Today Chemistry, 7 (2018) 53–64. https://doi.org/10.1016/J.MTCHEM.2018.01.001.

Emerging Nanotechnologies for Cancer Immunotherapy

23.

24. 25.

26. 27.

28. 29. 30. 31.

32.

33.

193

F. Duan, X. Feng, X. Yang, W. Sun, Y. Jin, H. Liu, K. Ge, Z. Li, & J. Zhang, A simple and powerful co-delivery system based on pHresponsive metal-organic frameworks for enhanced cancer immunotherapy. Biomaterials, 122 (2017) 23–33. https://doi.org/10.1016/J.BIOMATERIALS.2017.01.017. L. Milling, Y. Zhang, & D. J. Irvine, Delivering safer immunotherapies for cancer. (2017) 79–101. https://doi.org/10.1016/j.addr.2017.05.011.Delivering. V. Sathyanarayanan & S. S. Neelapu, Cancer immunotherapy: Strategies for personalization and combinatorial approaches. Molecular Oncology, 9 (2015) 2043–2053. https://doi.org/10.1016/j.molonc.2015.10.009. A. M. Scott, J. P. Allison, & J. D. Wolchok, Monoclonal antibodies in cancer therapy. Cancer immunity, 12 (2012) 14. A. A. Argyriou & H. P. Kalofonos, Recent advances relating to the clinical application of naked monoclonal antibodies in solid tumours. Molecular medicine (Cambridge, Mass.), 15 (2009) 183–91. https://doi.org/10.2119/molmed.2009.00007. J. A. Marin-acevedo, A. E. Soyano, B. Dholaria, K. L. Knutson, & Y. Lou, Cancer immunotherapy beyond immune checkpoint inhibitors. (2018). https://doi.org/10.1186/s13045-017-0552-6. H. Song & G. Sgouros, Radioimmunotherapy of solid tumours: Searching for the right target. Current drug delivery, 8 (2011) 26– 44. R. W. Jenkins, D. A. Barbie, & K. T. Flaherty, Mechanisms of resistance to immune checkpoint inhibitors. British Journal of Cancer, 118 (2018) 9–16. https://doi.org/10.1038/bjc.2017.434. J. Dine, R. Gordon, Y. Shames, M. K. Kasler, & M. Barton-Burke, Immune checkpoint inhibitors: An innovation in immunotherapy for the treatment and management of patients with cancer. Asia-Pacific Journal of Oncology Nursing, 4 (2017) 127–135. https://doi.org/10.4103/apjon.apjon_4_17. H. O. Alsaab, S. Sau, R. Alzhrani, K. Tatiparti, K. Bhise, S. K. Kashaw, & A. K. Iyer, PD-1 and PD-L1 Checkpoint signaling inhibition for cancer immunotherapy: Mechanism, combinations, and clinical outcome. Frontiers in Pharmacology, 8 (2017) 561. https://doi.org/10.3389/fphar.2017.00561. D. M. Francis & S. N. Thomas, Progress and opportunities for enhancing the delivery and efficacy of checkpoint inhibitors for cancer immunotherapy. 1 (2017) 33–42. https://doi.org/10.1016/j.addr.2017.04.011.Progress.

194

34. 35.

36.

37.

38. 39. 40. 41.

42.

43.

44.

Chapter 6

M. Terranova-, S. Thomas, & P. N. Munster, Epigenetic modifiers in immunotherapy: A focus on checkpoint inhibitors. 8 (2016) 705– 719. R. J. Torphy, R. D. Schulick, & Y. Zhu, Newly emerging immune checkpoints: promises for future cancer therapy. International Journal of Molecular Sciences, 18 (2017). https://doi.org/10.3390/ijms18122642. F. A. Schildberg, S. R. Klein, G. J. Freeman, & A. H. Sharpe, Coinhibitory pathways in the B7-CD28 ligand-receptor family. Immunity, 44 (2016) 955–72. https://doi.org/10.1016/j.immuni.2016.05.002. H. Tsai & P. Hsu, Cancer immunotherapy by targeting immune checkpoints: Mechanism of T cell dysfunction in cancer immunity and new therapeutic targets. (2017) 1–8. https://doi.org/10.1186/s12929-017-0341-0. R. Wang & H. Y. Wang, Immune targets and neoantigens for cancer immunotherapy and precision medicine. Nature Publishing Group, 27 (2016) 11–37. https://doi.org/10.1038/cr.2016.155. Y. Dong, Q. Sun, & X. Zhang, PD-1 and its ligands are important immune checkpoints in cancer. Oncotarget, 8 (2017) 2171–2186. https://doi.org/10.18632/oncotarget.13895. J. Dunn & S. Rao, Epigenetics and immunotherapy: The current state of play. Molecular Immunology, 87 (2017) 227–239. https://doi.org/10.1016/J.MOLIMM.2017.04.012. J. Gong, A. Chehrazi-Raffle, S. Reddi, & R. Salgia, Development of PD-1 and PD-L1 inhibitors as a form of cancer immunotherapy: a comprehensive review of registration trials and future considerations. Journal for Immunotherapy of Cancer, 6 (2018) 8. https://doi.org/10.1186/s40425-018-0316-z. C. Lei, P. Liu, B. Chen, Y. Mao, H. Engelmann, Y. Shin, J. Jaffar, I. Hellstrom, J. Liu, & K. E. Hellstrom, Local release of highly loaded antibodies from functionalized nanoporous support for cancer immunotherapy. Journal of the American Chemical Society, 132 (2010) 6906–7. https://doi.org/10.1021/ja102414t. Y. Fan, J. Moon, Y. Fan, & J. J. Moon, Nanoparticle drug delivery systems designed to improve cancer vaccines and immunotherapy. Vaccines, 3 (2015) 662–685. https://doi.org/10.3390/vaccines3030662. L. Galluzzi, E. Vacchelli, J. B. Pedro, A. Buqué, L. Senovilla, E. E. Baracco, N. Bloy, F. Castoldi, J. Abastado, P. Agostinis, & N. Ron, Classification of current anticancer immunotherapies. 5 (n.d.).

Emerging Nanotechnologies for Cancer Immunotherapy

45.

46.

47.

48. 49.

50.

51. 52.

53.

195

C. C. Kloss & M. V Maus, Adoptive T-cell transfer: Harnessing immune cells to combat disease. eLS (Chichester, UK: John Wiley & Sons, Ltd, 2015), pp. 1–9. https://doi.org/10.1002/9780470015902.a0026238. S. M. Fernandez-Poma, D. Salas-Benito, T. Lozano, N. Casares, J.I. Riezu-Boj, U. Mancheño, E. Elizalde, D. Alignani, N. Zubeldia, I. Otano, E. Conde, P. Sarobe, J. J. Lasarte, & S. Hervas-Stubbs, Expansion of Tumour-Infiltrating CD8+ T cells Expressing PD-1 improves the efficacy of adoptive T-cell therapy. Cancer Research, 77 (2017) 3672–3684. https://doi.org/10.1158/0008-5472.CAN-170236. S. J. Oiseth & M. S. Aziz, Cancer immunotherapy: A brief review of the history, possibilities, and challenges ahead. Journal of Cancer Metastasis and Treatment, 3 (2017) 250. https://doi.org/10.20517/2394-4722.2017.41. S.R. Jackson, J., Yuan, R.M, & Teague, Targeting CD8+ T-cell tolerance for cancer immunotherapy. Immunotherapy, 6 (2014) 833– 852. https://doi.org/10.2217/imt.14.51 K. M. Mahadeo, S. J. Khazal, H. Abdel-Azim, J. C. Fitzgerald, A. Taraseviciute, C. M. Bollard, P. Tewari, C. Duncan, C. Traube, D. McCall, M. E. Steiner, I. M. Cheifetz, L. E. Lehmann, R. Mejia, J. M. Slopis, R. Bajwa, P. Kebriaei, P. L. Martin, J. Moffet, J. McArthur, D. Petropoulos, J. O’Hanlon Curry, S. Featherston, J. Foglesong, B. Shoberu, A. Gulbis, M. E. Mireles, L. Hafemeister, C. Nguyen, N. Kapoor, K. Rezvani, S. S. Neelapu, & E. J. Shpall, Management guidelines for paediatric patients receiving chimeric antigen receptor T cell therapy. Nature Reviews Clinical Oncology, 16 (2019) 45–63. https://doi.org/10.1038/s41571-018-0075-2. M. T. Stephan, S. B. Stephan, P. Bak, J. Chen, & D. J. Irvine, Synapse-directed delivery of immunomodulators using T-cellconjugated nanoparticles. Biomaterials, 33 (2012) 5776–87. https://doi.org/10.1016/j.biomaterials.2012.04.029. H. T. Marshall & M. B. A. Djamgoz, Immuno-oncology: Emerging targets and combination therapies. 8 (2018) 1–29. https://doi.org/10.3389/fonc.2018.00315. W. Obara, M. Kanehira, T. Katagiri, R. Kato, Y. Kato, & R. Takata, Present status and future perspective of peptide-based vaccine therapy for urological cancer. Cancer Science, 109 (2018) 550–559. https://doi.org/10.1111/cas.13506. K. Palucka & J. Banchereau, Dendritic-cell-based therapeutic cancer vaccines. Immunity, 39 (2013) 38–48.

196

54.

55.

56.

57.

58.

59.

60.

61.

62.

Chapter 6

https://doi.org/10.1016/j.immuni.2013.07.004. C. S. Higano, P. F. Schellhammer, E. J. Small, P. A. Burch, J. Nemunaitis, L. Yuh, N. Provost, & M. W. Frohlich, Integrated data from 2 randomized, double-blind, placebo-controlled, phase 3 trials of active cellular immunotherapy with sipuleucel-T in advanced prostate cancer. Cancer, 115 (2009) 3670–3679. https://doi.org/10.1002/cncr.24429. P. F. Mulders, M. De Santis, T. Powles, & K. Fizazi, Targeted treatment of metastatic castration-resistant prostate cancer with sipuleucel-T immunotherapy. Cancer Immunology, IPPXQRWKHUDS\ࣟ CII, 64 (2015) 655–63. https://doi.org/10.1007/s00262-015-1707-3. J. Obeid, Y. Hu, & C. L. Slingluff, Vaccines, adjuvants, and dendritic cell activators - Current status and future challenges. Seminars in Oncology, (2015). https://doi.org/10.1053/j.seminoncol.2015.05.006. Y. Yao, Y. Zhou, L. Liu, Y. Xu, Q. Chen, Y. Wang, S. Wu, Y. Deng, J. Zhang, & A. Shao, Nanoparticle-based drug delivery in cancer therapy and its role in overcoming drug resistance. Frontiers in Molecular Biosciences, 7 (2020). https://doi.org/10.3389/fmolb.2020.00193. S. Senapati, A. K. Mahanta, S. Kumar, & P. Maiti, Controlled drug delivery vehicles for cancer treatment and their performance. Signal Transduction and Targeted Therapy, 3 (2018) 7. https://doi.org/10.1038/s41392-017-0004-3. R. Wang, P. S. Billone, & W. M. Mullett, Nanomedicine in action: An overview of cancer nanomedicine on the market and in clinical trials. Journal of Nanomaterials, 2013 (2013) 1–12. https://doi.org/10.1155/2013/629681. P. Yingchoncharoen, D. S. Kalinowski, & D. R. Richardson, Lipidbased drug delivery systems in cancer therapy: What is available and what is yet to come. Pharmacological Reviews, 68 (2016) 701–87. https://doi.org/10.1124/pr.115.012070. V. Sanna, N. Pala, & M. Sechi, Targeted therapy using nanotechnology: Focus on cancer. International Journal of Nanomedicine, 9 (2014) 467–83. https://doi.org/10.2147/IJN.S36654. M. Lorscheider, A. Gaudin, J. Nakhlé, K.-L. Veiman, J. Richard, & C. Chassaing, Challenges and opportunities in the delivery of cancer therapeutics: update on recent progress. Therapeutic Delivery, 12 (2021) 55–76. https://doi.org/10.4155/tde-2020-0079.

Emerging Nanotechnologies for Cancer Immunotherapy

63.

64.

65.

66.

67.

68. 69.

70.

71.

72.

197

P. K. Gautam, P. Deepak, S. Kumar, & A. Acharya, Role of macrophage in tumour microenvironment: Prospect in cancer immunotherapy. European Journal of Inflammation, 11 (2013) 1– 14. https://doi.org/10.1177/1721727X1301100101. K. Madaan, S. Kumar, N. Poonia, V. Lather, & D. Pandita, Dendrimers in drug delivery and targeting: Drug-dendrimer interactions and toxicity issues. Journal of Pharmacy & Bioallied Sciences, 6 (2014) 139–50. https://doi.org/10.4103/0975-7406.130965. R. I. Castro, O. Forero-Doria, L. Guzmán, R. I. Castro, O. ForeroDoria, & L. Guzmán, Perspectives of dendrimer-based nanoparticles in cancer therapy. Anais da Academia Brasileira de Ciências, 90 (2018) 2331–2346. https://doi.org/10.1590/0001-3765201820170387. H. Yang, Targeted nanosystems: Advances in targeted dendrimers for cancer therapy. 1DQRPHGLFLQH Nanotechnology, Biology, and Medicine, 12 (2016) 309–16. https://doi.org/10.1016/j.nano.2015.11.012. J. P. M. Almeida, E. R. Figueroa, & R. A. Drezek, Gold nanoparticle mediated cancer immunotherapy. Nanomedicine Nanotechnology, Biology, and Medicine, 10 (2014) 503–14. https://doi.org/10.1016/j.nano.2013.09.011. Hala Gali-Muhtasib, Racha Chouaib, Nanoparticle drug delivery systems for cancer treatment. 2020 by Jenny Stanford Publishing, Singapore. S. Carabineiro, Carabineiro, & S. A. Correia, Applications of gold nanoparticles in nanomedicine: Recent advances in vaccines. Molecules, 22 (2017) 857. https://doi.org/10.3390/molecules22050857. S. D. Jo, G.-H. Nam, G. Kwak, Y. Yang, & I. C. Kwon, Harnessing designed nanoparticles: Current strategies and future perspectives in cancer immunotherapy. Nano Today, 17 (2017) 23–37. https://doi.org/10.1016/J.NANTOD.2017.10.008. Y.-M. Park, S. J. Lee, Y. S. Kim, M. H. Lee, G. S. Cha, I. D. Jung, T. H. Kang, & H. D. Han, Nanoparticle-based vaccine delivery for cancer immunotherapy. Immune Network, (2013). https://doi.org/10.4110/in.2013.13.5.177. W. Park, Y.-J. Heo, & D. K. Han, New opportunities for nanoparticles in cancer immunotherapy. Biomaterials Research, 22 (2018) 24. https://doi.org/10.1186/s40824-018-0133-y.

198

73. 74.

75.

76.

77.

78.

79. 80.

81.

82.

Chapter 6

S. Shukla & N. F. Steinmetz, Emerging nanotechnologies for cancer immunotherapy. Experimental Biology and Medicine, 241 (2016) 1116–1126. https://doi.org/10.1177/1535370216647123. L. Jeanbart & M. A. Swartz, Engineering opportunities in cancer immunotherapy. Proceedings of the National Academy of Sciences of the United States of America, 112 (2015) 14467–72. https://doi.org/10.1073/pnas.1508516112. C. Devaud, L. B. John, J. A. Westwood, P. K. Darcy, & M. H. Kershaw, Immune modulation of the tumour microenvironment for enhancing cancer immunotherapy. Oncoimmunology, 2 (2013) e25961. https://doi.org/10.4161/onci.25961. S. L. Topalian, C. G. Drake, & D. M. Pardoll, Immune checkpoint blockade: A common denominator approach to cancer therapy. Cancer Cell, 27 (2015) 450–61. https://doi.org/10.1016/j.ccell.2015.03.001. D. Schmid, C. G. Park, C. A. Hartl, N. Subedi, A. N. Cartwright, R. B. Puerto, Y. Zheng, J. Maiarana, G. J. Freeman, K. W. Wucherpfennig, D. J. Irvine, & M. S. Goldberg, T cell-targeting nanoparticles focus delivery of immunotherapy to improve antitumour immunity. Nature Communications, 8 (2017) 1747. https://doi.org/10.1038/s41467-017-01830-8. K. H. Son, J. H. Hong, & J. W. Lee, Carbon nanotubes as cancer therapeutic carriers and mediators. International Journal of Nanomedicine, 11 (2016) 5163–5185. https://doi.org/10.2147/IJN.S112660. S. Bhaskar & M. S. Singh, Nanocarrier-based immunotherapy in cancer management and research. ImmunoTargets and Therapy, 3 (2014) 121. https://doi.org/10.2147/ITT.S62471. E. Panzarini, V. Inguscio, B. Tenuzzo, E. Carata, L. Dini, E. Panzarini, V. Inguscio, B. A. Tenuzzo, E. Carata, & L. Dini, Nanomaterials and autophagy: New insights in cancer treatment. Cancers, 5 (2013) 296–319. https://doi.org/10.3390/cancers5010296. N. Bertrand, J. Wu, X. Xu, N. Kamaly, & O. C. Farokhzad, Cancer nanotechnology: The impact of passive and active targeting in the era of modern cancer biology. Advanced Drug Delivery Reviews, 66 (2014) 2–25. https://doi.org/10.1016/J.ADDR.2013.11.009. N. E. Papaioannou, O. V. Beniata, P. Vitsos, O. Tsitsilonis, & P. Samara, Harnessing the immune system to improve cancer therapy. Annals of Translational Medicine, 4 (2016). https://doi.org/10.21037/10188.

Emerging Nanotechnologies for Cancer Immunotherapy

83.

199

J. Conniot, J. M. Silva, J. G. Fernandes, L. C. Silva, R. Gaspar, S. Brocchini, H. F. Florindo, & T. S. Barata, Cancer immunotherapy: Nanodelivery approaches for immune cell targeting and tracking. Frontiers in Chemistry, 2 (2014) 105. https://doi.org/10.3389/fchem.2014.00105.

CHAPTER 7 CURRENT DEVELOPMENT IN NANOMATERIALS FOR NUCLEIC ACID DELIVERY IN CANCER THERAPY SHEBA RANI DAVID*1 AND RAJAN RAJABALAYA2 1

School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2

*Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract RNA interference (RNAi) is a gene silencing mechanism currently used to target cancer-related genes. The rapid development of nanotechnology has facilitated the availability of various non-viral carriers that can be used for gene delivery. Importantly, naked DNA/RNA must overcome various barriers to exert the gene silencing activity in the target cells. The non-viral carriers are inorganic nanomaterials, carbon-based nanomaterials, proteins and peptide nanomaterials, lipid-based nanomaterials, polymer-based nanomaterials and dendrimers. These non-viral vectors offer many advantages, such as high transfection efficiency, reduced toxicity and improved endosomal escape. This chapter discusses the current development in nanomaterials for nucleic acid delivery in cancer therapy.

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

201

Abbreviations AAV AuNPs Bcl-2 CNTs DOPC dsRNA EGFP EPR LUC miRNA MNPs mRNA ncRNA PEG PEI PEIs PLGA PPI RES rHDL RISC RNAi shRNA siRNA SLNs TERT VEGF

Adeno-associated virus Gold nanoparticles B-cell lymphoma 2 Carbon nanotubes 1,2-dioleoyl- sn-glycero-3-phosphatidylcholineDouble-stranded RNA Enhanced green fluorescent protein Enhanced permeability and retention Luciferase MicroRNA Magnetic nanoparticles Messenger RNA Non-coding RNA Polyethylene glycol Polyethyleneimine Polyethyleneimines Polylactic acid-co-glycolic acid Polyproylenimine Reticuloendothelial system Reconstituted high-density lipoprotein RNA-induced silencing complex RNA interference Small hairpin RNA Small interfering RNA Solid lipid nanoparticles Telomerase reverse transcriptase Vascular endothelial growth factor

1. Introduction 1.1 The advances of nanotechnology The rapid advancement in nanotechnology has guided the development of nanoparticles for cancer treatment [1]. Typically, nanoparticles have a diameter ranging between 1 nm and 100 nm with at least one dimension of less than 100 nm, exhibiting an improved permeability and retention (EPR) effect. Hence, they circulate longer in the tumour site under in vivo conditions. Additionally, particles of 10–100 nm in size are non-toxic to

202

Chapter 7

mammalian cells, and nanoparticles bigger than 100 nm or smaller than 10 nm can get entrapped in the reticuloendothelial region of the immune system and other interstitial spaces of the body [2]. Nanoparticles can be divided into various categories depending on their morphology and physical and chemical properties. Nanoparticles can be easily engineered to the desired shape, size and surface charge. Examples of some nanoparticles are carbon-based nanoparticles, metal nanoparticles, ceramic nanoparticles, semiconductor nanoparticles, polymeric nanoparticles and lipid-based nanoparticles [3].

1.2 What is gene therapy Gene therapy is a technique used to prevent and treat various diseases, including cancer, by introducing exogenous nucleic acids, DNA or RNA into the host’s cells. There are two different types of systems used for nucleic acid delivery: viral vector and non-viral systems. In 2012, Glybera (alipogene tiparvovec), the first gene therapy used to treat lipoprotein lipase deficiency, was introduced. It represented DNA delivery with the help of viruses such as an adeno-associated virus (AAV) and gamma-retroviruses. However, despite a high transfection efficiency, the viral vector system is commonly associated with the activation of the immune system, which can decrease the effectiveness of gene delivery. The non-viral vector system is used to deliver exogenous nucleic acids to target the host’s cells to avoid the frequent activation of the immune system (Figure 7-1). The advances in nanotechnology have helped overcome some of the limitations of using nonviral vectors by increasing their efficiency in transfection, enhancing their ability to cross over multiple biological barriers and escape from endosomal degradation after endocytic uptake [4]. Cancer is characterised by genetic mutations that cause uncontrolled cell division, tumour growth, invasion and metastasis [5]. Cancer is one of the main targets of RNAi-based therapy. Oncogenes, tumour-suppressor genes and other genes associated with tumour growth are the primary targets for gene silencing by the RNAi-based therapy due to its significant effect, high specificity, minor side effects and easy synthesis [6]. RNAi, a member of non-coding RNA, is a post-translational gene regulation technology of gene therapy that leads to gene silencing associated with messenger RNA (mRNA). Fire and Mello first discovered RNAi in 1998. RNAi inhibits the expression of genes triggered by the genome origin microRNA (miRNA), small interfering RNA (siRNA) and double-stranded small hairpin RNA (shRNA) [7]. Table 7-1 shows that each non-coding RNA has different nucleotides in length and characteristics.

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

203

Figure 7-1 shows the entry of non-viral vectors into the cells via endocytosis and their passage to the nucleus through the cytoplasm Created with BioRender.com

Chapter 7

204

Table 7-1: Different classes of non-coding RNA (ncRNA) in mammalian [6-7]. Classes of ncRNA

Common abbreviations

Number of nucleotides/base pairs in length 21–22 nucleotides

Small interfering RNAs

SiRNA

MicroRNAs

miRNAs

20–23 nucleotides

Small hairpin RNAs

shRNAs

80 base pairs

Characteristics Produced from the cleavage of doublestranded RNA (dsRNA) by Dicer. SiRNAs form complexes with RISC, which involves in gene silencing. Bind completely to mRNA and result in its degradation or translational inhibition Produced from the cleavage of imperfect RNA hairpins by Dicer and drosha. miRNAs associated with RISC, which involves in post-transcriptional gene regulation Bind partially to mRNA and result in its degradation or translational inhibition Safer than siRNA and shRNA Produced from the cleavage of dsRNA by Dicer SiRNAs form complexes with RISC, which involves in gene silencing Bind completely to mRNA and result in its degradation or translational inhibition

SiRNA is a double-stranded RNA molecule with 21–23 nucleotides in length and two nucleotides overhanging at the 3’-ends [8]. The siRNA gene silencing mechanism works in the starting stage and the effecting stage.

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

205

SiRNA is derived from the first cleavage of double-stranded RNA by the enzyme Dicer in the starting stage. The double-stranded RNA can originate from bacterial DNA, viral genome, and synthetic RNA produced using bioinformatic data [6]. Then, each double-stranded RNA is split into guide and passenger strands by helicase in the effecting stage. Endogenous endonucleases will degrade the passenger strand, whereas the guide strand will be incorporated into the RNA-induced silencing complex (RISC). In the RISC, the guide strand of SiRNA will then bind with a complementary sequence in the messenger mRNA, which causes post-translational gene silencing [9]. RNAi includes siRNA, shRNA and miRNA, which possess the exact gene silencing mechanism except that only siRNA and shRNA bind completely to their target mRNA. In contrast, miRNA only binds partially complementary to its target mRNA. Generally, SiRNA is more suitable for cancer therapy because it is easier to synthesise. However, the naked SiRNA is extremely small and can easily escape from the body because of the high renal clearance rate. Therefore, attaching nanoparticles to SiRNA will increase the EPR effect, causing it to retain longer in the tumour region. Various delivery systems, such as inorganic nanomaterials (gold and magnetic nanoparticles), carbon-based nanoparticles (fullerenes, carbon nanotubes and graphene), proteins and peptide nanomaterials (albumin and elastin), lipid-based nanomaterials (liposome and gemini surfactants), polymer-based nanomaterials (polyethyleneglycol (PEG), polyethyleneimine (PEI) and dendrimers, are used for gene delivery.

2. Barriers or challenges in the delivery of non-viral vectors Numerous barriers block the siRNA molecules from reaching the target cells and delivering therapeutic outcomes. Hence, several steps are required for the siRNA delivery from the administration site to the target site. Various barriers, such as vascular and cellular barriers, must be overcome to ensure siRNA delivery to target cells. The naked siRNA can transfect tumours but will normally undergo enzymatic degradation in the systemic circulation, causing the ineffective delivery of therapeutic genes to the target sites. The major challenge is that siRNA is negatively charged, which causes repulsion from the similarly negatively charged cell surface membrane and results in poor cell uptake. Hence, ‘packaging’ of the naked DNA or RNA with a suitable delivery system tends to protect the siRNAs and reach the desired site of action to deliver their therapeutic action in the form of gene silencing [10].

206

Chapter 7

Most siRNA-based drugs for cancer therapy are not available orally because the RNAi drug must survive the acidic environment of the digestive tract, particularly the stomach, and pass through the intestinal epithelium into the circulation. Subcutaneous injection is another administration route having the ability to avoid the first-pass metabolism in the liver and directly enter the blood circulation via capillaries or lymphatic drainage. However, the lipophilicity and size of the formulation are a challenge that must be considered as they all can affect the ability of the drug to reach circulation. It is also reported that subcutaneous administration applies to 60-90 nm liposomal nanoparticles. This is due to many factors that must be considered while using subcutaneous injection. Thus, only intravenous and infusion injections are the most commonly used routes for administering the siRNAbased drug [11]. Naked siRNA is not regularly applied in systemic delivery because the enzyme nucleases rapidly degrade this siRNA. The half-life of naked siRNA lies between several minutes and 1 hour. Additionally, the kidneys also play a significant role in siRNA clearance. Multiple studies in animals have demonstrated that the biodistribution of siRNA exhibits the most uptake in the kidney. The use of delivery systems or carriers protects siRNA from enzymatic degradation. After systemic administration, siRNA is taken up by the reticuloendothelial system (RES) comprising phagocytic cells such as monocytes and macrophages, whose functions are to clear foreign substances [9]. Further, naked siRNA does not easily cross the negatively charged cell surface membrane by passive diffusion due to its large size, high molecular weight and negative phosphate. Hence, the coating of siRNA with a positively charged carrier can promote the binding of the positively charged siRNA-carrier complex to the anionic cell membrane and enter the cells by endocytosis via the formation of endocytic vesicle [12]. Off-target effects of siRNA are defined as the adverse effects arising from the modulation of other targets. These effects may be biologically related or completely unrelated to the target of interest [13]. This result can complicate the interpretation of siRNA's therapeutic effects, resulting in gene expression changes and unwanted toxic effects. Further, a high level of siRNAs can result in an inflammatory response by activating Toll-like receptors and producing cytokines. TLRs, which are involved in siRNA recognition, are TLR3, TLR7 and TLR8 [9,14]. Hence, four key requirements for a successful siRNA delivery are as follows: 1) Protect siRNA from enzymatic degradation and rapid renal clearance; 2) High loading of siRNA without any toxicity effect; 3) Promote cellular uptake by escaping from endosomal

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

207

uptake; and 4) Transport siRNA to the desired sites and avoid non-specific delivery [15].

3. Types of nanomaterial in gene delivery (delivery system) and their characteristics A broad range of nanoparticle systems is used for siRNA-based gene silencing in cancer treatment. siRNA could either be included in the core of the nanoparticles or as a surface-attached molecule around the nanoparticles via non-covalent or covalent forces. The surface-attached siRNA may be prevented from degradation by chemical modification or adding a layer of neutral polymers such as polyethyleneglycol (PEG). Some of the nanoparticles used for gene delivery are presented below. The advantages of using nanoparticles for gene delivery are prolonged half-life in blood, improved pharmacokinetics, and increased EPR effect [10]. Delivery systems are also used to specifically recognise the target cells, escape from nuclease degradation and selectively release siRNA into the cytoplasm [8].

3.1 Inorganic nanomaterials Inorganic nanomaterials have been used to deliver genes because of their unique properties, such as biocompatibility, non-immunogenicity, nontoxicity and easier production. Inorganic nanoparticles also serve as a crucial platform for gene delivery in cancer therapy due to their small size and high surface-area-to-volume ratio, which are beneficial in siRNA loading at a high rate. Some examples of inorganic nanomaterials are gold and magnetic nanoparticles. Magnetofection uses DNA-coated superparamagnetic iron oxide nanoparticles for gene delivery in the presence of a magnetic field. After introducing magnetoreception, several systems were developed where magnetic nanoparticles made of iron oxide are complexed into viral or nonviral vectors to improve the transfer of the genes into targeted tissues [16]. In vivo, the magnetic fields, which are targeted to a specific site, can increase transfection efficiency and locate the therapeutic gene to a specific site within the body. Particles carrying the therapeutic gene are typically injected intravenously. External magnets with high-gradient capture the particles as they pass through the bloodstream. The captured particles then stay at the target site until the tissue takes them up. The nanoparticles used to deliver genes are small enough to enter the cells and nucleus while escaping the endosomal degradation [17]. For example, monocytes can be removed from the patient. They can thus be transfected with magnetic

208

Chapter 7

nanoparticles coated with the therapeutic gene and re-injected back into the patient. The application of a magnetic field close to/on the tumour will increase the adhesion effect of monocytes to tumour vasculature and invade tumour cells. However, in vivo magnetoreception also has several challenges, such as lower transfection efficiency, high clearance of iron oxide particles, and increased safety concerns regarding the accumulation of iron oxide in cells, particularly in multi-dosing experiments [16]. Recently, gold nanoparticles (AuNPs) are a reliable delivery system for siRNA due to their physicochemical properties, such as shape, surface area, amphiphilicity, easy surface functionalisation and biological non-toxicity. Attaching polyvalent siRNA to gold nanoparticles through thiol groups can increase the half-life and stability of free siRNA [10]. The high reduction potential of gold (Au) keeps the gold nanoparticles (AuNPs) together while circulating in the bloodstream. Importantly, the addition of PEG molecules to gold nanoparticles results in extended retention and makes nanoparticles avoid opsonisation and stealth [2]. AuNPs were used to deliver B-cell lymphoma 2 (Bcl-2) or VEGF siRNAs to a human cervical carcinoma cell line for silencing the enhanced green fluorescent protein (EGFP) and luciferase (LUC) reporter genes [14]. However, gold nanoparticles have disadvantages, such as low encapsulation efficiency, slow endosomal escape and poor storage stability [7].

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

209

Figure 7-2: Potential role of magnetic nanoparticles (MNPs) in improving the delivery of a monocyte-based gene to cancer cells. The patient’s monocytes are transfected with a loaded therapeutic gene. These are combined with sterile MNPs and re-injected into the patient. Applying a magnetic field near/on the tumour leads to increased infiltration of MNPs in the tumour. Created with BioRender.com

210

Chapter 7

3.2 Carbon-based nanoparticles Carbon nanotubes are considered the potential vehicles or carriers for the delivery of nucleic acids because of their attractive physical and chemical properties. Carbon nanotubes (CNTs) could easily cross the cell surface membrane and directly translocate into the cytoplasm of target cells. [18]. They can be either single-walled carbon nanotubes of a diameter between 0.4 and 3 nm or multi-walled carbon nanotubes, which can be modified with additional functional groups for loading nucleic acids [4]. The incubation of siRNA, which inhibits telomerase reverse transcriptase (TERT) to positively charged single-walled carbon nanotubes with tumour cells, was effective in siRNA delivery and reducing the expression of TERT in tumour cells [14]. However, the positive charge of carbon nanotubes is toxic in mammalian cells. Hence, modifying carbon nanotubes with polymers is effective for gene delivery [2]. A fullerene is a form of carbon nanoclusters with a unique hydrophobic spherical structure. Some properties of fullerenes are as follows: high chemical reactivity, redox property and sensitivity toward the light (photosensitivity). Amphipathic fullerene nanostructures can deliver genes to form a complex with the genes of interest. Fullerenes are generally made multifunctional by synthesising derivatives of cationic molecules, such as aminofullerenes, poly-N-dimethylfulleropyrrolidinium, and fullerene epoxide tetra(piperazino), which could effectively deliver the gene of interest under in vivo conditions. Further, fullerenes can protect the nucleic acids from degradation by forming a protective sheath until the nucleic acids are released into the cytoplasm via the degradation of fullerenes or the loss of binding ability with the nucleic acids [2].

3.3 Proteins and peptide nanomaterials Proteins are well favoured for gene delivery because they are biodegradable, biocompatible, and have low toxicity. Protein and peptides nanomaterials can condense nucleotides via electrostatic interactions between negatively charged nucleic acids and positively charged amino acids. Gelatin is one of the most widely used proteins due to the ease of modification and the low cost of production. Gelatin is formed by collagen hydrolysis and comprises peptides of different lengths. For example, the use of psi(RFP)-tGel nanoparticles can induce gene silencing in RFP/B16F10 melanoma cells resulting in an 80% reduction of RFP mRNA expression [4]. However, proteins-based nanomaterials also have several negative aspects, such as

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

211

difficulty controlling their molecular size and slow biodegradability, which may cause systemic toxicity [19].

3.4 Lipid-based nanomaterials (liposome and Gemini surfactants) Lipid-based nanoparticles, such as liposomes, solid lipid nanoparticles (SLNs) and reconstituted high-density lipoprotein (rHDL) nanoparticles, have been developed for the systemic delivery of DNA or RNA into tumours [20]. Liposomes were first discovered in the 1960s by Bangham et al. In the 1970s; liposomes were introduced as drug delivery vehicles [21]. The main advantage of liposomes as the delivery system is reducing side effects at the injection site. They are also proven to be a highly efficient delivery system of siRNA because they can be formulated to approximately 100 nm and their by-product are not toxic to living tissues [22]. Neutral liposomes are less toxic to the living tissue than cationic lipids. However, neutral liposomes have lower entrapment efficiency due to the wear interactions between negatively charged nucleic acids and neutral lipids during formulation [18]. The efficiency of gene delivery depends on the liposomal charge and size, degree of PEGylation, type of targeting ligand and tissue targeting [10]. The cationic lipids constitute a cationic head group, a connecting linker and a lipophilic group. The cationic head group, the tail group comprising the carbon chain, and the linker’s nature can affect siRNA vehicles' toxicity and transfection efficiency [12]. The lipid encapsulation process of the SiRNA involves mixing the siRNA solution with cationic liposomes to form lipoplexes through the binding of the cationic liposomes with the negatively charged SiRNA, which prevents siRNA from degradation by nucleases, encourages uptake into target cells via endocytosis, increase the release of siRNA from endo-lysosomal compartments, and facilitate the accumulation of siRNA in the cytoplasmic matrix [10–12]. A study reported that 1,2dioleoyl-sn-glycero-3-phosphatidylcholine-(DOPC)-encapsulated siRNA against Ephrin Type-A receptor 2 (EohA2) oncoprotein was effective (65%) in reducing the expression of EphA2 after 48 hours of a single-dose administration in the orthotopic model [14]. Further, the PEGylated liposome can protect siRNA from the reticuloendothelial system and also help to delay the binding of opsonin proteins to the surface of liposomes, allowing them to circulate longer and accumulate in the tumour region [10].

212

Chapter 7

3.5 Polymer-based nanomaterials (polyethyleneglycol [PEG] and polyethyleneimine [PEI]) The use of polymer-based nanoparticles in drug and gene delivery increases day by day. Polymers are recognised as an efficient delivery system because they are easily modified and suitable for drug loading. Most commonly used polymers are polyethyleneglycol (PEG), polyethyleneimine and poly-llysine, poly(lactic-co-glycolic) acid. The interaction between the positively charged polymers and the negatively charged nucleic acids leads to the formation of nanosized complexes known as polyplexes. The properties of polyplex, such as size, structure and surface charge, prevent nucleic acid degradation by blocking the entry of nucleases into the enclosed nucleic acid drugs [21]. Polyethyleneglycol (PEG) can prolong systemic circulation, improving the efficiency of delivering therapeutic genes. PEG is also commonly used as polymers because it can reduce immunogenicity, avoid aggregation and prevent therapeutic loss due to endosomal degradation [5]. Polyethyleneimines (PEIs) are the cationic polymer commonly used for gene delivery due to their high transfection efficiency. The strong cationic nature of polyethyleneimines (PEIs) obtained from the presence of numerous amine groups enables it to interact with siRNA; however, the formation of siRNA-PEI complexes depends on the molecular weight of PEI and the ratio of free amino groups in PEI to phosphate groups in siRNA known as the N/P ratio [10]. A branched PEI can increase transfection efficiency, whereas a high molecular weight PEI tends to have a high toxicity effect [12]. The inclusion of PEI helps siRNA to overcome endosomal entrapment during transfection, allowing the siRNA to escape from the acidic environment of endo-lysosomes. Thus, PEI enhances the transfection efficiency of siRNA [10]. For example, the intraperitoneal administration of PEI–siRNA targeting the c-erbB2/neu (HER-2) receptor resulted in the receptor’s downregulation and a significant decrease in tumour growth [12]. However, PEI also has some significant toxicity issues, including the damage to the membrane and activation of a mitochondriamediated apoptotic programme [18]. PLGA is a biodegradable polymer approved by US Food and Drug Administration [18]. Polylactic acid-co-glycolic acid (PLGA) is beneficial for siRNA delivery because it is considered a safe, biocompatible, and biodegradable polymer [7]. Further, PLGA has lower toxicity than cationic polymers and lipids [12]. However, PLGA causes a slow release of encapsulated siRNA, resulting in low efficiency. A novice technology known as particle replication in non-wetting templates (PRINT) has been

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

213

used to modify PLGA-based nanoparticles for better gene delivery to overcome this problem [10]. PLGA modification is done by incorporating PEI into PLGA nanoparticles, which showed complete siRNA loading, prevented siRNA degradation by nuclease, and increased gene silencing activity in the cells compared to PEI only [12].

3.6 Dendrimers Dendrimers are highly branched polymers that consist of a central core molecule. Dendrimers having hydrophilic group end groups are soluble in polar solvents, whereas dendrimers ending in hydrophobic groups are soluble in non-polar solvents. Multiple functional groups, such as -OH, COOH and -NH2, are present in dendrimers, allowing the conjugation of various ligands and small molecules (transferrin, biotin, folic acid) antibodies against tumour-associated antigens. The amines inside dendrimers act as a proton sponge that enables endosomal escape and nucleic acid delivery into the cytoplasm [23]. Poly-amidoamine (PAMAM) is a commonly used dendrimer [7]. The most commonly used dendrimers used for gene delivery are PAMAM dendrimers due to the ease of synthesis and commercial availability [18]. The PAMAM dendrimers are made up of ethylenediamine and ammonia core with three to four branching points and are relatively non-toxic. PAMAM is a cationic dendrimer that can interact effectively with the anionic nucleic acids via electrostatic forces of attraction and condense it to form dendriplexes. Despite their benefits in increasing the cellular uptake of siRNA, dendrimers are found to cause cellular toxicity. Fortunately, the surface modification with agents like PEG to PAMAM dendrimers can improve the efficiency in delivering genes and reduce the overall toxicity [15]. The other dendrimers used to deliver genes are polyproylenimine (PPI) units, with butylenediamine used as the core molecule. These dendrimers have several unique chemical and physical properties, differentiating them from other polymers used for gene delivery. [23].

214

Chapter 7

Table 7-2: Summary of the advantages and disadvantages of siRNA delivery system/carriers Carriers Liposomes

Advantages Cationic lipids: Biocompatible, Biodegradable and easy to synthesise Neutral lipids: low toxicity

Proteins

Biodegradable Biocompatible Low toxicity

Carbon nanotubes

Cationic nature of CNTs can easily bind with negatively charged siRNA or other nucleic acids Ease of surface modification Easy to control its size Non-toxic

Gold nanoparticles

Magnetic nanoparticles

Used for both diagnostic and therapeutic purposes High transfection efficiency

Polymer-based nanoparticles

PEI: high transfection efficiency PLGA: Biocompatible and Biodegradable

Dendrimers

High surface area can carry a high concentration of siRNA

Disadvantages Cationic lipids: High amount of liposomes exhibit toxicity Cationic and neutral lipids: lack of size control Difficult to control their molecular size Slow biodegradable, which may cause systemic toxicity Non-biodegradable

Low encapsulation efficiency Slow endosomal escape Poor storage stability Rapid systemic clearance of iron oxide particles Accumulation of iron oxide in cells, particularly in multidosing experiments, is toxic to cells PEI: non-degradability, cytotoxicity PLGA: low transfection efficiency Positively charged dendrimers interact with negatively charged cell surfaces, leading to haematological toxicity

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

215

4. Clinical application of nanoparticles-based RNAi therapy Despite the significant attention received in this field, most of the siRNA non-viral delivery systems for the treatment of cancer are still in preclinical studies. In May 2008, the first human-phase I trial of CALAA-01 was initiated in patients with solid tumours, resulting in the successful delivery of nanoparticles and a decrease of the corresponding mRNA in tumour biopsies. In July 2009, the Atu027 was opened to patients with advanced solid tumours because preclinical studies showed that the repeated IV administration of Atu027 resulted in the gene silencing of protein kinase N3 expression in rats, mice and non-human primates. It also showed tumour growth inhibition and prevented lymph node metastasis in orthotopic mouse models of pancreatic and prostate cancer [12]. DCR-MYC is being used to suppress Myc oncoprotein. A dose-escalation trial investigated the clinical activity of DCR-MYC in patients with advanced solid tumours and multiple myeloma. This trial showed that at different dose levels, DCR-MYC provides excellent therapeutic responses [14]. Additionally, DOPCencapsulated siRNA liposomes targeting the EphA2 oncoprotein were highly effective in reducing the expression of EphA2 after intravenous or intraperitoneal administration of the DOPC-encapsulated siRNA in mouse models of various human cancers [18]. S12GD LODER is a polymericbased matrix with a siRNA against a mutated KRAS oncogene. The overexpression of mutated KRAS oncogene causes more than 90% of human pancreatic ductal carcinoma. The phase I trial of TKM-080301, a lipid nanoparticle with siRNA against polo-like kinase I for a patient having cancer with hepatic metastases, was completed in July 2016 [1].

EphA2

Liposomes (Neutral liposomes)

Biodegradable polymeric matrix

Lipid nanoparticles

Lipid-based nanoparticles

Cyclodextrin nanoparticle contained polymer SiRNAlipoplex

Carrier

IV infusion

Hepatic intraarterial injection

IV infusion

IV infusion

IV infusion

Delivery route IV injection

Chapter 7

Solid tumours; multiple myeloma; Cancer with hepatic metastases Pancreatic Ductal Adenocarcinoma Pancreatic Cancer Solid tumours

Advanced solid cancer

Solid tumours, Cancer

Disease

I

I

I

I

I

I

Phase

Recruiting

Completed

Completed

Terminated

Completed

Terminated

Status

NCT01591356

NCT01188785

NCT02191878

NCT02110563

NCT00938574

Clinical trials identifier NCT00689065

RRM2: M2 subunit of ribonucleotide reductase, PKN3: Protein kinase N3, VEGF: Vascular endothelial growth factor, KSP: kinesin spindle protein, MYC-, PLK-1: polo-like kinase 1, EphA2: Eph receptor A

siRNAEphA2DOPC

M.D. Anderson Cancer Center

PLK-1

TKM080301

KRASG12D

MYC

DCRMYC

siG12D LODER

PKN3

Target site RRM2

Atu027

Name of drug CALAA01

Silenseed, Ltd.

Silence Therapeutics GmBH Dicerna Pharmaceutical, inc Tekmira Pharmaceutical

Calando Pharmaceuticals

Manufacturer

Table 7-3: SiRNA-based drug for cancer therapy

216

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

217

5. Conclusion In addition to surgery, chemotherapy and radiation methods, RNAi can be considered a promising new cancer therapy approach because it promotes post-transcriptional gene silencing. Due to major restrictions of naked siRNA of low transfection efficiency and fast degradation in the cellular cytoplasm, siRNA needs to be complexed with suitable vehicles or carriers to optimise the siRNAs delivery to their target sites. Although greater advances have been made for an efficient siRNA delivery, several problems associated with the safety, effectiveness and ease of manufacturing or production must be considered when selecting suitable siRNA carriers. Future studies must focus on the safety profiles of various carriers following in vivo delivery of siRNA to develop more effective delivery systems for the RNAi-based cancer therapeutics.

References 1.

2. 3. 4. 5. 6. 7. 8.

M. Díaz & P. Vivas-Mejia, Nanoparticles as drug delivery systems in cancer medicine: Emphasis on RNAi-containing nanoliposomes. Pharmaceuticals, 6 (2013) 1361–1380. https://doi.org/10.3390/ph6111361. S. Singh, Nanomaterials as Non-viral siRNA delivery agents for cancer therapy. %LR,PSDFWVࣟ BI, 3 (2013) 53–65. https://doi.org/10.5681/bi.2013.007. I. Khan, K. Saeed, & I. Khan, Nanoparticles: Properties, applications and toxicities. Arabian Journal of Chemistry, (2017). https://doi.org/10.1016/j.arabjc.2017.05.011. M. Riley & W. Vermerris, Recent advances in nanomaterials for gene delivery—A review. Nanomaterials, 7 (2017) 94. https://doi.org/10.3390/nano7050094. F. Maiyo & M. Singh, Selenium nanoparticles: Potential in cancer gene and drug delivery. Nanomedicine, 12 (2017) 1075–1089. https://doi.org/10.2217/nnm-2017-0024. B. Mansoori, S. S. Shotorbani, & B. Baradaran, RNA interference and its role in cancer therapy. Advanced Pharmaceutical Bulletin, 4 (2014) 313–321. https://doi.org/10.5681/apb.2014.046. I. Fernandez-Piñeiro, I. Badiola, & A. Sanchez, Nanocarriers for microRNA delivery in cancer medicine. Biotechnology Advances, 35 (2017) 350–360. https://doi.org/10.1016/j.biotechadv.2017.03.002. H. J. Kim, A. Kim, K. Miyata, & K. Kataoka, Recent progress in development of siRNA delivery vehicles for cancer therapy.

218

9. 10.

11. 12. 13. 14.

15.

16. 17.

18.

19.

Chapter 7

Advanced Drug Delivery Reviews, 104 (2016) 61–77. https://doi.org/10.1016/j.addr.2016.06.011. C. Xu & J. Wang, Delivery systems for siRNA drug development in cancer therapy. Asian Journal of Pharmaceutical Sciences, 10 (2015) 1–12. https://doi.org/10.1016/j.ajps.2014.08.011. A. Babu, R. Muralidharan, N. Amreddy, M. Mehta, A. Munshi, & R. Ramesh, Nanoparticles for siRNA-based gene silencing in tumour therapy. IEEE Transactions on NanoBioscience, 15 (2016) 849–863. https://doi.org/10.1109/TNB.2016.2621730. D. Haussecker, Current issues of RNAi therapeutics delivery and development. Journal of Controlled Release, 195 (2014) 49–54. https://doi.org/10.1016/j.jconrel.2014.07.056. J. Wang, Z. Lu, M. G. Wientjes, & J. L. S. Au, Delivery of siRNA therapeutics: Barriers and carriers. AAPS Journal, 12 (2010) 492– 503. https://doi.org/10.1208/s12248-010-9210-4. D. G. Rudmann, On-target and off-target-based toxicologic effects. Toxicologic Pathology, 41 (2013) 310–314. https://doi.org/10.1177/0192623312464311. G. Mahmoodi Chalbatani, H. Dana, E. Gharagouzloo, S. Grijalvo, R. Eritja, C. D. Logsdon, F. Memari, S. R. Miri, M. R. Rad, & V. Marmari, Small interfering RNAs (siRNAs) in cancer therapy: A nano-based approach. International Journal of Nanomedicine, 14 (2019) 3111–3128. https://doi.org/10.2147/IJN.S200253. P. Kesharwani, S. Banerjee, U. Gupta, M. C. I. Mohd Amin, S. Padhye, F. H. Sarkar, & A. K. Iyer, PAMAM dendrimers as promising nanocarriers for RNAi therapeutics. Materials Today, 18 (2015) 565–572. https://doi.org/10.1016/J.MATTOD.2015.06.003. M. Alsaggar & D. Liu, Physical methods for gene Transfer. Adv. Genet. (Academic Press Inc., 2015), pp. 1–24. https://doi.org/10.1016/bs.adgen.2014.10.001. S. Majidi, F. Zeinali Sehrig, M. Samiei, M. Milani, E. Abbasi, K. Dadashzadeh, & A. Akbarzadeh, Magnetic nanoparticles: Applications in gene delivery and gene therapy. Artificial Cells, Nanomedicine, and Biotechnology, 44 (2015) 1–8. https://doi.org/10.3109/21691401.2015.1014093. J.-M. Lee, T.-J. Yoon, & Y.-S. Cho, Recent developments in nanoparticle-based siRNA delivery for cancer therapy. BioMed Research International, 2013 (2013). https://doi.org/10.1155/2013/782041. D. Verma, N. Gulati, S. Kaul, S. Mukherjee, & U. Nagaich, Protein based nanostructures for drug delivery. Journal of Pharmaceutics,

Current Development in Nanomaterials for Nucleic Acid Delivery in Cancer Therapy

20.

21.

22.

23. 24.

219

2018 (2018) 1–18. https://doi.org/10.1155/2018/9285854. Y. Xiao, K. Shi, Y. Qu, B. Chu, & Z. Qian, Engineering nanoparticles for targeted delivery of nucleic acid therapeutics in tumour. Molecular Therapy - Methods & Clinical Development, 12 (2019) 1–18. https://doi.org/10.1016/j.omtm.2018.09.002. T. Kafshdooz, L. Kafshdooz, A. Akbarzadeh, Y. Hanifehpour, & S. W. Joo, Applications of nanoparticle systems in gene delivery and gene therapy. Artificial Cells, Nanomedicine and Biotechnology, 44 (2016) 581–587. https://doi.org/10.3109/21691401.2014.971805. W. Hasan, K. Chu, A. Gullapalli, S. S. Dunn, E. M. Enlow, J. C. Luft, S. Tian, M. E. Napier, P. D. Pohlhaus, J. P. Rolland, & J. M. DeSimone, Delivery of multiple siRNAs using lipid-coated PLGA nanoparticles for treatment of prostate cancer. Nano Letters, 12 (2012) 287–292. https://doi.org/10.1021/nl2035354. C. Dufès, I. F. Uchegbu, & A. G. Schätzlein, Dendrimers in gene delivery. Advanced Drug Delivery Reviews, 57 (2005) 2177–2202. https://doi.org/10.1016/j.addr.2005.09.017. K. Tatiparti, S. Sau, S. Kashaw, & A. Iyer, siRNA Delivery strategies: A comprehensive review of recent developments. Nanomaterials, 7 (2017) 77. https://doi.org/10.3390/nano7040077.

CHAPTER 8 NANOPARTICLES AS DRUG DELIVERY SYSTEM IN MELANOMA SHEBA RANI DAVID*1 AND RAJAN RAJABALAYA2 1

School of Pharmacy, University of Wyoming, Laramie, WY 82071, USA PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2

*Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract For decades, melanoma has always been reported as one of the top lethal cancers worldwide. Until now, many effective drug delivery systems, such as nanoparticles, have been available to manage melanoma. Different types of nanoparticles are used to treat melanoma, and their effects are comparable with the rest of the drug delivery systems. Effective drug delivery systems can be achieved by considering the nature of the nanoparticles and their target. Because of different melanomas, the nanoparticles used as a drug delivery system can be customised by conjugating them with the corresponding ligand that could bind to the target site. Apart from this flexibility, nanoparticles can be easily adjusted and modified to enhance the permeability, transport, drug release and cellular targeting throughout the delivery pathway. The general factors of nanoparticles affecting the drug delivery pattern are their particle characteristics, cellular

Nanoparticles as Drug Delivery System in Melanoma

221

targeting and toxicity. Several studies have supported nanoparticles’ characteristics, which contribute to effective drug delivery. However, most of the studies did not sufficiently establish the toxicity effect of nanoparticles. This chapter’s limitations include a small number of references and studies used to support this literature topic.

Abbreviations CA4P CM DiI DOPC DSAA HHV HNE HPV IL-2 MAPK MC1R MM MSH MTIC NM NP PACM-ECD PAR-1 PEG PLGA Rb siRNA SSM TMZ UM UV WHO

Combretastatin phosphate Cutaneous melanoma Dioctadecyl-3, 3, 3ƍ, 3ƍ-tetramethyl-indocarbocynine perchlorate 2 dioleoyl-sn-glycero-3-phosphatidylcholine N,N-distearyl-N-methyl-N-2- (Nƍ-arginyl) aminoethyl ammonium chloride Human herpes virus 4-hydroxynonenal Human papillomavirus Interleukin-2 Mitogen-activated protein kinase Melanocortin 1 receptor Mucosal melanoma Alpha-melanocyte-stimulating hormone 5-(3-dimethyl-1-triazenyl) imidazole-4-carboxamide Nodular melanoma Nanoparticles E-cyclodextrin-poly(4-acryloylmorpholine) conjugate Protease-Activated-Receptor-1 Polyethyleneglycol Poly(lactic-co-glycolic) acid Retinoblastoma Short interfering RNA Super spreading melanoma Temozolomide Uveal melanoma Ultraviolet World Health Organization

222

Chapter 8

1. Introduction Melanoma is the cancer of melanocytes wherein these pigment-producing cells have transformed into their malignant form [1]. Melanocytes are developed from the ectoderm or, more specifically, the pluripotent neural crest stem cells during the embryo stage of human life [1–2]. The function of melanocytes is to produce brown-black or yellow-red pigment known as melanin, which helps absorb harmful ultraviolet (UV) light [2]. In the epidermis of the skin, melanocytes are also present in the eye, mucous membrane, oesophagus, and meninges, where the sites are concentrated with pigment. Therefore, melanoma is further classified into cutaneous, uveal and mucosal melanoma, which arise from melanocytes in the epidermis, ocular stroma and mucous membrane, respectively. Regardless of the areas affected, all the melanocytes are prone to transforming into malignant ones [1]. Because the pathogenesis of melanoma is well established, various treatment methods were designed to target the abnormalities and the malignancy of melanomas. The standard treatments of melanoma are chemotherapy, immunotherapy and targeted therapy. Nanoparticles are one of the possible candidates to be utilised as drug carriers in the treatment of melanoma to improve the effectiveness of these treatments. Different types of nanoparticles, such as liposomes, solid lipid nanoparticles, polymeric nanoparticles, quantum dots, and carbon nanotubes, are considered suitable for melanoma. Nevertheless, many aspects need to be considered while designing the nanoparticles. The surface characteristics of nanoparticles play a significant role in determining the effectiveness of drug delivery in the tumour microenvironment as those in melanoma. Several studies have been selected to assess the capability of nanoparticles as drug delivery systems in different cases of melanoma. This literature review investigates whether the surface characteristics, drug delivery and release patterns are suitable to be applied in melanoma.

Nanoparticles as Drug Delivery System in Melanoma

223

2. Stages of melanoma

Figure 8-1: The progressive stages of melanoma from normal skin into Stage I till IV. Created with BioRender.com

224

Chapter 8

2.1 Cutaneous melanoma Cutaneous melanoma (CM) is the most common type of melanoma [1]. CM is referred to as ‘melanoma skin cancer’ or ‘malignant melanoma’ [3]. CM is recognised as the most fatal and destructive type among all the skin cancer types [4]. According to the World Health Organization (WHO), the annual incidents of melanoma skin cancer are estimated to be 132,000 cases globally and account for 1.6% of all diagnosed cancers [5]. Karimkhani et al. revealed that the nations with the highest melanoma burden are New Zealand, Australia and Europe from 21 regions selected in the study [6]. The ‘burden’ mentioned in the context refers to the quantifiable degree of health loss caused by the parameters, such as diseases, injuries and risk factors, which are usually measured and estimated by a scientific system known as the Global Burden of Disease Study [6]. Chang et al. (2017) claimed that melanoma skin cancer has been neglected and inadequately treated in several Asian countries such as Taiwan, despite the cases being comparatively less than those in Western countries like Europe. The survival rate of skin melanoma in Asian countries is relatively low and constantly dropping, especially in Taiwan [7]. Besides that, the incidence rate is also rising globally [1]. The mortality rate of melanoma is profoundly high due to late diagnosis because symptoms only arise when melanoma has reached the advanced stage and metastasised [8]. These studies show that CM still imposes a tremendous burden on the affected populations despite improvements in prevention, diagnosis, and treatment. The causes of CM can be classified as genetic and environmental [1]. The genetic cause of CM is associated with the gene mutation of the protein receptors involved in CM, which is the melanocortin 1 receptor (MC1R). MC1Rs are found on the cell surface of melanocytes. They produce pigment through a signalling cascade, which is then activated by the alpha-melanocyte-stimulating hormone (MSH). The pigmentation of the skin has an absolute and significant impact that can potentially cause the skin to become malignant. Additionally, the gene mutation of the MC1R gene contributes to the decreased pigmentation of the skin and increased skin sensitivity to UV light [1]. The environmental factor greatly responsible for developing melanoma is UV light exposure. The UV light, more specifically the UV-B light, radiates from the sun and is more concentrated along the equator. Therefore, a higher number of skin melanoma cases are found in the equatorial regions [1]. Most of the risk factors that could develop skin melanoma are associated with the degree of sun exposure, such as occasional childhood sun exposure and individuals who experienced severe sunburns more than five times. The other risk factors include individuals with xeroderma pigmentosum, a

Nanoparticles as Drug Delivery System in Melanoma

225

genetic condition due to the inability to repair the skin’s damaged DNA by UV light. The use of a sunbed, also known as a solarium, can cause CM for individuals, especially those under the age of 35 years. The UV-A irradiated sunbed can increase the risk of developing CM by 75%. Additionally, individuals with a history of skin melanoma, those with immunosuppression who had organ transplantation, patients with other skin malignancies like non-melanoma skin cancers, and individuals with high naevus counts are at high risk of developing cancer CM [1]. CM can be classified further into four different basic types, and each of them presents different clinical manifestations [8]. The most common and popular type is super spreading melanoma (SSM), which accounts for 60– 70% of all CM cases and affects young people most of the time [8–9]. SSM is present as an asymmetric, irregular plague with a pigment that presents a variety of colours ranging from black to red, brown, blue and white [9]. It can be found on any body part with a diameter of greater than 0.5 cm [8–9]. In men, SSM is common on the trunk, whereas in women, it is usually found on their legs. However, SSM can be noticed on the upper back in both genders. Second, nodular melanoma (NM) is generally diagnosed when the affected area becomes a visible dark brown or black, dome-shaped bump. It accounts for 15–30% of CM cases [8]. The nodule bleeds and ulcerates in NM. At the beginning of NM development, the nodule has an irregular edge and presents with a mixture of different colours, which are usually absent. This is why NM is often diagnosed late after becoming visible with a welldefined border and symmetrical shape once the nodule has turned malignant [9]. NM usually occurs in the trunk, upper and lower limbs, and scalp, mainly affecting the elderly [8]. Another type of CM that usually affects the elderly is Lentigo maligna melanoma (LMM), accounting for 5–15% of all CM cases. It is presented with multiple colours of pink, blue, white and grey pigments and has an irregular and notched border of 1–20 cm or greater [9]. LMM is commonly caused by chronic sun exposure, which damages the skin irradiated by UV light [8]. Finally, the type of CM that accounts for 2– 8% is Acral lentiginous melanoma (ALM), also known as malignant melanoma under the nail [9]. It is found commonly in populations with dark skin, such as among African-Americans and Asians [8]. It is presented with brown or black discolouration underneath the fingernails, on the palms of the hands, on the soles of the feet, or on the mucous membrane [8–9]. The malignancy in ALM can advance more quickly than in SSM and LMM [8].

226

Chapter 8

2.2 Uveal melanoma Uveal melanoma (UM) is recognised as the most common eye malignancy. It accounts for 3–5% of all melanoma cases [1,10]. The specific affected area could be the iris, ciliary body or choroid in naturally pigmented areas of the eye [10]. The incidence of UM is more frequent in the white population of the Western countries like the USA, compared to the South African black population and Asian population from Far East regions. Although the number of cases has been consistent throughout the years, it is concerning that the survival rate of UM did not improve. UM's most significant risk factor is the exposure to sunlamps in the eyes and the mutation of the BAP1 tumour-suppressor gene [1,10]. Some studies reviewed by Ali et al. claimed that several factors were associated with the risk of developing UM, such as individuals having light skin and iris colour and chronic lifetime exposure to the UV-B light [1]. The clinical manifestation of UM is the development of darkly pigmented nevi in the uvea or fundus of the affected eye that will grow towards the vitreous space with the shape of a mushroom [5,10]. In UM, the usual symptoms are blurred vision and seeing flashing lights and shadows, whereas sometimes UM is asymptomatic upon diagnosis [5,10].

2.3 Mucosal melanoma Mucosal melanoma (MM) develops on the mucous membrane where the sites are unexposed to UV radiation. MM accounts for 1.5% of all melanoma cases, and the incidence rate has remained stable annually for a few decades. It is suspected that the development of MM has a relationship with the innate immune system because the mucosal membrane is usually acting as the critical barrier against foreign microbes. Besides that, the melanocytes residing in the mucosal membrane functioned as the antigenpresenting cells in the event of immune responses. They played a neuroendocrine role at the leptomeningeal sites of the brain and spinal cord. The affected areas are the head, neck, genital tract in females and anorectal region [1]. MM can be classified according to the area of the mucous membrane affected. Thus, the subtypes of MM are conjunctival, sinonasal, laryngeal, primary pulmonary, gastrointestinal, oral, oesophageal, small bowel, colon, gall bladder, vaginal, urethral, urine bladder and penile melanoma [11]. The melanoma tumour appears to be flat or slightly elevated from the mucosal surface in the beginning. Eventually, the tumour will become darker in colour when the malignancy of MM is advanced and will invade the underlying tissues [12]. As MM is not induced by the sun, the

Nanoparticles as Drug Delivery System in Melanoma

227

possible causal factors could be the human papillomavirus (HPV) and human herpesvirus (HHV), which commonly contribute to oral melanoma. Aside from viruses, carcinogens such as formaldehyde and tobacco can induce oral melanoma [1].

3. Melanomagenesis Because the treatment for melanomas is primarily being designed and targeted at a molecular level, a good understanding of the pathways involved in melanoma is essential. The pathways are almost similar in the pathogenesis of most melanomas. However, it could be different in the case of MM. In CM and UM, the activated pathways are mostly induced by UV radiation. The common pathways in melanoma are mitogen-activated protein kinase (MAPK), PI3K/AKT, which involves protein RAS, c-KIT, CDK4 or CDK 6, GNAQ or GNA11, and MITF. Figure 8-1 shows the MAPK pathways. The genetic mutation of the MAPK pathway is commonly present in most tumour types. The pathway's cascade of proteins activation in this key signalling promotes cell growth and survival. In normal conditions, MAPK will be eventually inhibited directly by the proteins such as the CRAF protein [1]. In CM, the proteins that are usually mutated to cause the deregulation of the MAPK pathway are NRAS and BRAF [1–4]. The MAPK pathway in CM can be induced by UV radiation, especially UV-B, at a chronic exposure [13]. In this case, the skin's cell surface receptors, such as the epidermal growth factor receptor (EGFR), will be activated and then activate the MAPK pathway [13]. Aside from the proteins involved in the MAPK pathway, the mutation of protein regulator p16, which normally acts as the negative regulator in the cell cycle, will deregulate the cell cycle and potentially stimulate malignancy in melanocytes [1]. In UV, melanogenesis could be highly caused by disrupting the function of tumour-suppressor genes such as retinoblastoma (Rb), cyclin D4 and p53 [1]. Additionally, the mutations in GNAQ and GNQ11 genes, which encode the alpha subunit of the heterotrimeric cell surface of G proteins, may also be responsible for melanogenesis [1,4]. GNAQ is also correlated with the MAPK pathway (Figure 8-2) [4]. The gene mutation that codes for the BRCA1-associated protein-1 (BAP1) has been discovered in most patients with metastasising UM. Moreover, there is a significant increase in vascular endothelial growth factors (VEGFs), specifically VEGF-A levels in the aqueous humour of eyes, in many cases of UM [1]. Compared to CM and UM, melanogenesis occurs predominantly due to the mutation of the proto-

228

Chapter 8

oncogene protein, c-KIT, which will subsequently activate the MAPK and phosphoinositide 3-kinase (PI3K) pathways [1,4]. Similarly to MAPK, the PI3K pathway could also promote cell proliferation, migration, differentiation and survival of melanocytes [1].

4. Treatment of melanoma Various treatments can target dysfunctional pathways, knowing how these mutations are associated with melanomas. The treatments are summarised in Table 8-1.

5. Nanoparticles Nanoparticles (NP) have been widely used in many clinical indications because they possess different properties by modifying them. NPs have sizes ranging between 10 nm and 1000 nm. They can usually be found in various products such as cosmetics and foods. One of the most frequent clinical applications of NPs is in cancer treatment. As a form of drug carrier, NPs can benefit in several ways. The advantages of NPs are improved bioavailability and prolonged stay within the body. Additionally, they are known to specifically target the site for the therapeutic action [16]. In addition to these advantages, NPs require low doses and have minimum toxicity. They also promise protection from the severe side effects of NPs attaching to non-target tissue [17]. NPs can encapsulate, dissolve and/or entrap the active ingredient of nanomolecular size before delivering the active ingredients to the targeted site [16]. There are many different types of NPs available as drug carriers in therapeutic, diagnostic and theranostic aspects. Table 8-2 depicts the types and characteristics of different types of NPs.

Photodynamic therapy

Chemotherapy

Electrochemotherapy

N/A

N/A

Temozolide

N/A

N/A

Examples Dacarbazine

Class N/A

Table 8-1: The treatment for melanoma.

Suitable palliative treatment choice for patients with Stage III/IV metastatic CM

Adjuvant therapy

For CM

High electric impulses facilitate the drug delivery to the cells

Technique that combines with cytotoxic drugs: bleomycin and cisplatin

For advanced CM and metastatic UM

For metastatic CM, UM melanoma, and MM Prodrug of active metabolite of dacarbazine

Standard of care in melanoma

Description Alkalyting agent

Nanoparticles as Drug Delivery System in Melanoma

[14]

[14]

[14–15]

Ref. [11,14,15]

229

Immunotherapy

230 N/A

N/A N/A

Ontak

Interferon D-2b

Peginterferon D-2b

Interleukin-2

Treg inhibition

Increase antitumoral immunity

Inhibit antitumoral immune responses

Activate Teffs

For metastatic melanoma

Increase number of effector T cells and Tregs

Adjuvant therapy for Stage III melanomas

Dose-dependent proapoptotic effect

Inhibit the proliferation of melanoma cells

Exerts antiangiogenic, antiproliferative and antitumor activities

Stimulates the immune system to produce multiple cell types such as natural killer cells and B lymphocytes

Interfere with viral replication

Chapter 8

[14]

[14]

[14]

[14]

Biochemotherapy

Nivolumab

Talimogene laherpatepvec

Programmed cell death protein / PD-1 ligand blockade

Oncolytic virus therapy Gp 100 peptide vaccine Adoptive T-cell therapy Combination of dacarbazine, cisplatin and vinblastine

Ipilimumab

Cytotoxic T lymphocyteassociated antigen 4 blockade

[14]

Infusion with a high concentration of melanoma-specific T cells

[14]

[14–15]

Enhance cytotoxic T lymphocytes reactivity

Combination of immunotherapy and chemotherapy

[14]

[14–15]

[14–15]

For metastatic CM and UM Modified herpes simplex virus type 1

Reduce tumour progression by mediating immune response and inducing antitumor activity

Inhibits the binding of ligands PD-LI and PDL2 to the PD-1 receptor

For advanced CM and metastatic UM

Enhance the production of pro-inflammatory T-cell cytokine

Blocking the inhibitory effect

Nanoparticles as Drug Delivery System in Melanoma 231

Targeted therapy

232 Vemurafenib Trametinib

Imatinib Bevacizumab

Combination of PI103 and rapamycin as the PI3K inhibitor and mTOR inhibitor, respectively Ribociclib Iapatinib

BRAF inhibitor

MEK inhibitors

C-KIT inhibitors

VEGF inhibitors

PI3K-AKTmTOR pathway inhibitors

Cyclin-dependent kinase inhibitors

ErbB4 inhibitor

Reduction in cell growth

Inhibits CDK4 and thus inhibits the cellular proliferation

Downregulate the pathway

Inhibit tumour growth

Neutralise VEGF

c-KIT mutations in MAPK and PI3K/AKT pathways

For metastatic UM with BRAF mutations

For unresectable or metastatic CM with BRAF mutations

For unresectable or metastatic melanomas with BRAF mutations

Chapter 8

[14]

[14]

[14]

[14]

[14]

[14–15]

[14]

N/A

N/A

Type Liposomes

Has a good biocompatibility profile

Rapid clearance and thus short circulating halflife

Simplest form is a phospholipid bilayer surrounding an aqueous core

Can integrate target ligands into the liposomal NPs

Most easily synthesised NPs

Diameters of 40 to 50 nm

Characteristics Platforms for drug delivery of high toxicity and low bioavailability drugs

Table 8-2: Different types of nanoparticles

Radiotherapy

Materials Phospholipid bilayer

Example Liposomal irinotecan (Onivyde“) for treatment of pancreatic cancer

Provides loco-regional control of head/neck (MM) melanomas

Improves local control of sinonasal (MM) melanoma

Nanoparticles as Drug Delivery System in Melanoma

Ref [18–20]

[11]

233

Solid Lipid NPs

Polymeric NPs

234

Can produce on a large scale

Easy to sterilise

Enhanced stability and thus, greater control in drug delivery and less chance of drug leakage

Selective on cancer cells without disrupting normal cells Average size of less than 500 nm

Size can be maintained through internal body circulation

Improve chemotherapy effectiveness by giving a consistent drug delivery rate

Broad application

Simple synthesis

Simplest form of NPs

Chapter 8

Natural polymers: Chitosan Protein Polysaccharides

Synthetic polymers: Polyethyleneglycol (PEG) Poly(lactic-coglycolic) acid (PLGA) Polylactic acid PEGylated gamma E1a (PLEGRIDY“) for treatment of relapsing multiple sclerosis

[11]

[18–21]

Carbon nanotubes

Dendrimers

Polymeric micelles

Less toxic

Biocompatible and hydrophilic

Able to modify into various functions

Consists of a hexagonal arrangement of carbon atoms

Tube made of the graphene sheet

Drug-dendrimer complex is stable in both phosphate buffer saline and aqueous solutions

Easy to conjugate and modify for different functions

Mono disperse synthetic macromolecule

Capable of carrying less aqueous soluble drugs

Has a hydrophobic inner core and hydrophilic external surface

Inherent and modifiable

Controlled delivery of hydrophobic drugs achieved by its self-assembled amphiphiles

Nanoparticles as Drug Delivery System in Melanoma

Carbon nanotubeMTX

PEG-pluronicDOX

Topical micellar formulation of estradiol (Estrasorb•) for severe vasomotor symptoms of menopause

[16]

[7]

[18–19]

235

Quantum dots

236

Consists of a large surface area for attachment of active material and can develop multifunctional properties

Size range of 2 to 10 nm

Basic semiconductors

Chapter 8 [16–20]

Nanoparticles as Drug Delivery System in Melanoma

237

6. Factors affecting nanoparticles 6.1 Particle size Several factors are considered while designing effective NP delivery systems. The size of NPs can affect the pharmacokinetic parameters of NPs, such as biodistribution, elimination and clearance. The biodistribution is affected by the size of NPs concerning how they will interact with the cells and adhere to certain biological complexes in the body [16]. Concerning elimination, some NPs are designed to avoid and delay its degradation in the body [16,22]. The clearance of NPs is associated with their capability to avoid physiological systems that are normally ready to erase and degrade foreign substances. The systems are the mononuclear phagocytic system (MPS) and the reticuloendothelial system. It has been proven that NPs with sizes larger than 100 nm can potentially be targeted by protein opsonin in the blood serum and proceed to be degraded by the MPS. The process in which the opsonin protein binds substances is known as opsonisation, which constitutes a part of the NPs. The avoidance of opsonin with a smaller size of NPs (i.e., less than 100 nm) can thus prolong the time in the blood circulation [16]. However, the size of NPs should not be too small to encourage rapid leakage into the blood capillaries. To avoid NPs being caught by the reticuloendothelial system in the spleen and the liver, NPs delivered by injection are far more effective than the conventional form of NPs. Nevertheless, it still needs to depend on adjusting the size and surface features [22].

6.2 Particle charge The surface charge coated around the surface of NPs affects the action of NPs and their delivery through the cellular membranes. The surface charge of NPs will repel each other evenly in suspensions formulated with NPs. Thus, the nanosuspension can be stabilised, and aggregations can be prevented. Mucoadhesion and retention of the NPs can be enhanced by the NPs wherein their surface charge is modified to become positively charged. The mucus membrane occupies the negatively charged polyelectrolyte properties, attracting the oppositely charged NPs. Several anionic polymers incorporated on polymeric NPs could display a disadvantageous characteristic. The NPs will be repelled from the mucous membrane and unable to proceed for cellular uptake. In contrast, positively charged NPs can help promote the NPs’ attraction and adhesion on the membrane for extensive cellular uptake by endocytosis or other mechanisms [16].

238

Chapter 8

6.3 Hydrophilicity of particle NPs made of hydrophilic materials such as polyethylene glycol (PEG) tend to be able to escape the capture by macrophages. PEG helps to avoid opsonisation by repelling the plasma protein. This characteristic is somewhat similar to that of particle charge [22].

6.4 Particle shape Depending on the types of particles used, the shapes of the NPs may alter the period of their residence in the blood circulation and the degree of cell uptake. Polymeric micelles with longer micelle lengths and smaller body spheres could provide greater cellular uptake and efficiency in delivering the drug. The net length of NPs can affect the cellular adhesion on the targeted area; the longer the particle length, the lower the degree of NPs binding onto the targeted membrane. Therefore, the shape of NPs can determine the therapeutic outcomes of the drug incorporated within it [16].

6.5 Cell targeting Membrane receptors and other big complex biological molecules are the most common targets for NPs. Normally, biological processes occur when a complement and suitable ligand act against the target receptors. Therefore, the receptor interactions and NP valence capacitance need to be considered beforehand while designing an NP. During a cellular uptake, the proteins caveolae and clathrin will play an essential role to facilitate the endocytosis mechanism wherein NPs have their specificity towards them . Caveolaemediated endocytosis is usually facilitated in the event of carcinogenic cellular uptake, which should not be happening in normal healthy cells. Considering this outcome, NP can be designed to target the cancerous cells without disrupting the structure and functions of the normal cells. This action can prevent unwanted cytotoxicity in the normal cells [16]. Such a mechanism is also labelled as passive targeting, where the permeability and retention effect are enhanced [22].

7. Transport of NPs When NPs interact with the cellular plasma membrane, endocytosis or passive diffusion are used as uptake methods. Endocytosis comprises two different categories: pinocytosis and phagocytosis. Pinocytosis is also referred to as receptor-mediated endocytosis and macropinocytosis.

Nanoparticles as Drug Delivery System in Melanoma

239

Receptor-mediated endocytosis involves the protein clathrin predominantly, where the receptors are simultaneously engulfed when attached to its ligand (i.e., attached to the NPs) [16]. This mechanism is also particularly known as active targeting [22]. Thus, clathrin-protein-coated pits are formed, which is the classical form in receptor-mediated endocytosis. In the case of caveolae-mediated endocytosis, the molecules first bind onto the caveolae surface before being engulfed. However, unlike clathrin-mediated endocytosis, this kind of endocytosis did not separate the protein from the vesicle following endocytosis. Macropinocytosis works by engulfing larger solute macromolecules with a size of around 5 P m. The cellular uptake that involves phagosomes is known as phagocytosis, where phagosomes usually ingest and transport large microparticles for elimination [16]. Figure 8-2 shows the multistage processes of endocytosis and passive diffusion.

8. NPs used in melanoma Currently, FDA-approved NPs are not available for detecting and managing melanoma. However, several studies have been made to assess the efficiency of NPs as drug delivery systems in the treatment and diagnosis of melanoma.

240

Chapter 8

Figure 8-2: Different cellular uptake mechanisms of nanoparticles thorough nonphagocytic internalisation pathways: Macropinocytosis (A), clathrin-mediated endocytosis (B) and caveolae-mediated endocytosis (C). Created with BioRender.com

Nanoparticles as Drug Delivery System in Melanoma

241

8.1 Studies on NPs used in the diagnosis of CM Zheng H, Chen G, DeLouise LA and Lou Z (2010) investigated an in vitro method of conjugating PEG-COOH capped highly fluorescent NPs quantum dots with an antibody that attaches with the adaptor molecule (i.e., streptavidin). They investigated to detect the CD146 complex on cultured and live cells of CM [23]. CD146 is an antigen that is strongly involved in the progression of CM. Its overexpression is common in malignant melanocytes. The magnificent photostability and improved brightness over the fluorescein isothiocyanate (FITC) (i.e., the fluorescent substance during the imaging process) injected cells of QDs labels have demonstrated that it can be an effective tool in cells imaging [23]. Al-Jamal et al. (2008) investigated another in vivo investigation using QD in imaging as multimodal nanoparticles. The functionalised-quantum-dotliposome (f-QD-L) hybrid NPs were formed by encapsulating poly(ethylene glycol)-coated QDs in various aqueous lipid bilayer vesicles, resulting in the formation of functionalised-quantum-dot-liposome (f-QD-L) hybrid NPs [24]. f-QD-L hybrids can rapidly accumulate in the tumour and remain there for a reasonably long time during the imaging process. Al-Jamal et al. performed the imaging on mice with C57B16 solid tumour of CM implanted subcutaneously in them. The effectiveness of f-QD-L was compared with controlled dioctadecyl-3, 3, 3ƍ,3ƍ-tetramethyl-indocarbocyanine perchlorate (DiI)-labelled liposomes. Both f-QD-L and DiI-liposomes were injected intratumorally. The result demonstrated that intratumoral binding, retention and distribution profiles were identical [24].

8.2 Studies on NPs used in the treatment of CM Nab-Paclitaxel is one of the often used treatment choices for metastatic breast cancer. NPs albumin-bound paclitaxel (Nab-PTX) was investigated in a group of patients with previously treated and untreated CM for three to four weeks during a Phase II clinical trial described by E. Hersh et al. (2010). The results demonstrated that both tested populations were had high tolerance levels and effective with Nab-PTX. The response rate, progression-free survival and survival were compared with the standard dacarbazine therapy and other combination therapies and those used in the previously treated patients with CM [25]. Mitrus et al. (2009) investigated a combination treatment of a vasculardisruptive medication with a liposomal version of a chemotherapeutic agent. Combretastatin phosphate (CA4P) and doxorubicin were used as the

242

Chapter 8

vascular-disruptive drug and chemotherapeutic [26]. Combretastatin exerts anti-vascular actions that inhibit the polymerisation of tubulin and thus prevent cancer cell division. Doxorubicin worked by intercalating into the DNA helix, arresting the cell cycle by inhibiting topoisomerase II. The drug was administered intratumorally into 6-to-8-week-old C57B1/6 mice that were inoculated with B16-F10 cells. Compared to the monotherapies, this combination therapy showed more significant inhibition in CM cells proliferation and growth [26]. This result shows that liposomal NPs could assist in preventing the escape of the drug from the leaky capillary with its size of about 100 nm. Aside from that, the liposomal NP, in this case, enhanced the drug’s permeability and its retention time in the tumour. Villares et al. (2008) investigated short interfering RNA (siRNA) integrated with liposomal NPs. The overexpression of the thrombin receptor, proteaseactivated-receptor-1 (PAR-1), is usually found in metastatic CM cell lines and patients with metastatic lesions. PAR-1 activation could promote the adhesion, invasion, and angiogenesis of malignant cells. To suppress PAR1 activation, PAR-1siRNA is introduced to the affected site systemically. In this case, PAR-1 siRNA is incorporated to a neutral liposome 1,2 dioleoylsn-glycero-3-phosphatidylcholine (DOPC) forming the PAR-1-siRNADOPC complex. This complex was then inoculated into the mouse model with the most metastatic CM cell lines (i.e., A375SM). The result showed effectiveness in silencing the PAR-1 activation. Thus, there were significant reductions in angiogenesis by IL-8 and VEGF and the invasion factors such as the MMPs involved in the CM tumour progression. Besides that, compared with viruses (i.e., shRNA), liposomal NPs were considered safer as drug carriers [27]. Chen et al. (2010) had also conducted a study on siRNA; however, they used cationic lipid N, N-distearyl-N-methyl-N-2- (Nƍ-arginyl) aminoethyl ammonium chloride (DSAA) as NPs [28]. The siRNA was directed at the tumour location in their study, which had c-Myc overexpression. C-Myc is involved in cancerous cells, such as malignant melanocytes, proliferation, growth and survival. DSAA was believed to be a suitable drug delivery of c-Myc siRNA. It consisted of a head group of arginine residue. The ligand used was anisamide, which was then bound to the sigma 1 receptor of melanoma cells. The c-Myc siRNA with NP DSAA was applied on cultured B16F10 murine melanoma cells that had expressed the sigma 1 receptor. Compared with c-Myc siRNA alone, the result with targeted DSAA NPs had shown effectiveness in inhibiting the growth of the tumour. DSAA could trigger the cells’ apoptosis, and the antiapoptotic protein Bcl-2 presented in B16F10 melanoma cells was decreased and downregulated

Nanoparticles as Drug Delivery System in Melanoma

243

[28]. This action was attributed to the induction of reactive oxygen species by DSAA. Therefore, c-Myc siRNA with DSAA NPs can be considered therapeutically effective in treating melanoma. Tran et al. (2008) conducted an in vitro study on targeted therapy against v600E BRAF and Akt3 [29]. The proteins BRAF and Akt3 are highly involved in the activation and deregulation of the MAPK pathway (Figure 8-1). Tran et al. designed cationic liposome NPs loaded with siRNA that were characterised to formulate a topical delivery system. With the application of low-frequency ultrasound, the results showed effectiveness in delaying melanocytic lesion development in the skin tested and suppressing the potential melanoma metastases. These liposome NPs showed improved skin permeation through the skin barrier. Additionally, an ultrasound enhanced siRNA penetration through the skin [29]. Thus, the combination of two drug delivery systems (i.e., NPs and ultrasound) seems favourable for the topical delivery of anticancer treatment in CM. Pizzimenti et al. (2013) investigated a polymeric derivative of -cyclodextrinpoly(4-acryloylmorpholine) conjugate (PACM-CD) employed as NPs in the delivery of4-hydroxynonenal (HNE). HNE powerfully exerts anticancer activities, such as inhibiting cell proliferation and promoting cell apoptosis in many cancerous cells. It is usually toxic in most biological cells. HNE was incorporated into PACM-ECD NPs and delivered to prepare threedimensional skin reconstruction of A375 melanoma cells. Before the HNE/PACM-ECD complex was introduced to the 3D melanoma skin structure in the in vitro studies with cultured cells of A375 human CM tumour cell line, the complex had shown the ability to internalise within the melanoma cells rapidly. The A375 melanoma cell growth inhibition by the HNE/PACM-ECD complex was clearly shown when applied in the incubated 3D melanoma skin structure compared to free HNE [30]. This action must be attributed to PACM-ECD NPs' capability to enhance and speed up the delivery of HNE into the A375 CM cells. Interleukin-2 (IL-2) as immunotherapy is another antitumor treatment option for melanoma. IL-2 stimulates immune responses, producing more natural killer cells and B cells at the malignant sites. Yao et al. (2011) designed the polycationic vector (H1) as the NPs to carry human IL-2 plasmid polyplexes, forming the H1/pIL-2 polyplexes. H1 comprises polyethyleneimine attached to E-cyclodextrin and conjugated with folic acid. The in vitro study with mouse B16-F1 melanoma cells showed a reduction of B16-F1 melanoma cell growth [31]. In contrast, in the in vivo study where H1/pIL-2 polyplexes were injected subcutaneously on the left and

244

Chapter 8

right flank of the 6–9-week-old c57BL/6N female mice, the result showed effectiveness in suppression of the melanoma tumour growth compared with the control (i.e., H1/pEGFP). Additionally, there is an increase in the accumulation of natural killer cells and B cells in the injected site. In comparison with the previous study by the same authors that used recombinant adenoviruses expressing IL-2 (rAdv-IL-2), H1/pIL-2 had maintained the level of tumour suppression after the injection as the polyplexes (i.e., the NPs) were not attacked by the specific host antibodies [31]. Therefore, this result has proven that NPs could protect, prolong and stabilise the residence of the drug in the body for a reasonable period during the treatment. A recent study conducted by Clemente et al. (2018) used solid lipid NPs (SLN) as the drug carrier for an anticancer drug: Temozolomide (TMZ). Clemente et al. hypothesised that with SLN used as the drug carrier, the conversion of TMZ to 5-(3-dimethyl-1-triazenyl) imidazole-4-carboxamide (MTIC) can be prevented, and its therapeutic efficacy will be improved. MTIC was found to be responsible for the side effects such as toxicity in the liver, heart and pulmonary systems and myelosuppression when TMZ was given at a therapeutic level dose. Besides that, SLN is believed to effectively carry the drug through biological barriers and increase drug stability in different biological environments. The in vitro experiment on the cultured mouse B16-F10 melanoma cells had evidenced the capability of SLN to reduce the side effects due to the hydrolysis of TMZ to MTIC. Compared with TMZ alone, SLN-TMZ had significantly decreased the Ki67 expression in the tumour cells. Additionally, there was also an increase in TH17/Treg cell ratio as the result of increased expression of IL-17A. In contrast, the expression of IL-10 remained undisturbed [32]. This result shows that SLN-TMZ can be a potential candidate for the treatment of melanoma.

8.3 Studies of NPs used in UM Yang et al. (2018) published a review that used HPV as the NPs in the treatment of UM. The complex is known as AU-011, which comprises HPV conjugated to an infrared-activated photodynamic dye. Currently, this complex is in phase Ib/II of its clinical trial for patients diagnosed with UM, more specifically, small primary choroidal melanoma. AU-011 is administered intravitreally in the affected eye. Its viral ligands bind selectively to the cancerous cells presented with the overly expressed and modified heparin sulphate proteoglycans [15]. This destroyed the tumour cell membrane as the ophthalmic laser was activated. The viral NPs, in this

Nanoparticles as Drug Delivery System in Melanoma

245

case, improve the binding specificity onto the malignant cells, resulting in effective targeted delivery of the therapeutic substances.

9. Future directions for the research It is interesting to see how NPs could significantly overcome the difficulties experienced by typical standard conventional or systemic drugs. There are no contradictions in studies for NPs used in CM. All the outcomes are relatively similar: NPs incorporated therapy shows more effectiveness and improvement than free drugs or therapeutic substances (i.e., drugs or substances that are not bound to NPs). However, the studies had a lack of assessment of the toxicity of NPs, where toxicity is one of the critical parameters in designing safe and effective drug delivery NPs. Besides, the types of NPs studied in this literature review are not fully covered as only a few NPs were being scientifically studied. The route of administration applied to the study subjects are mostly intratumoral and a few other routes (i.e., topical and subcutaneous). A different route of administration could affect the analysis of the NPs characteristics in bringing out the desired effects. The targeted therapy can be improved with NPs, especially polymeric NPs, as they are selective towards several tumour microenvironments. According to Masood (2016), polymers used in polymeric NPs such as PLGA can target PI3K receptors [33]. As mentioned in the earlier section, melanoma can be caused by the deregulation of the PI3K/Akt pathway. This finding could be applied as potential targeted therapy with PI3K-AKT-mTOR pathway inhibitors incorporated in PLGA polymeric NPs as the drug carrier for a more accurate target. To prove the effectiveness of NPs and their safety, research more on studies that involve the toxicity of NPs in clinical uses. Researchers are encouraged to find good quality studies from other different qualified scientific databases for the future.

10. Conclusions Several types of nanoparticles have shown therapeutic and diagnostic potential in treating and diagnosing various forms of melanomas. There are more examples and research employing nanoparticles in cutaneous melanoma than in uveal and mucosal melanoma among the three forms of melanomas. On top of that, the treatment options for uveal and mucosal melanomas are minimal compared to cutaneous melanoma. This chapter reviews a number of research strategies on nanoparticles as drug carriers, which indicate promising and good results in melanoma therapy and

246

Chapter 8

diagnostics. With a solid understanding of melanoma's biology and pathophysiological aspects and the properties (i.e., the surface characteristics and delivery mechanisms) of various types of NPs, effective and safe NPs as drug delivery vehicles may be devised.

References 1. 2. 3.

4. 5.

6.

7.

8. 9.

Z. Ali, N. Yousaf, & J. Larkin, Melanoma epidemiology, biology and prognosis. European Journal of Cancer, Supplement, 11 (2013) 81–91. https://doi.org/10.1016/j.ejcsup.2013.07.012. G. J. Tortora & B. Derrickson, Principles of anatomy and physiology (2000). https://doi.org/10.1016/S0031-9406(05)60992-3. Zoe Apalla, Dorothée Nashan, Richard B. Weller, Xavier Castellsagué, Skin cancer: Epidemiology, disease burden, pathophysiology, diagnosis, and therapeutic approaches. Dermatology and Therapy, 7 (2017) 5–19. https://doi.org/10.1007/s13555-016-0165-y. M. A. Davies & J. E. Gershenwald, Targeted therapy for melanoma: A primer. Surgical Oncology Clinics of North America, 20 (2011) 165–180. https://doi.org/10.1016/j.soc.2010.09.003. C. Pandiani, G. E. Béranger, J. Leclerc, R. Ballotti, & C. Bertolotto, Focus on cutaneous and uveal melanoma specificities. Genes and Development, 31 (2017) 724–743. https://doi.org/10.1101/gad.296962.117. C. Karimkhani, A. C. Green, T. Nijsten, M. A. Weinstock, R. P. Dellavalle, M. Naghavi, & C. Fitzmaurice, The global burden of melanoma: Results from the Global Burden of Disease Study 2015. British Journal of Dermatology, 177 (2017) 134–140. https://doi.org/10.1111/bjd.15510. J. W. C. Chang, J. Guo, C. Y. Hung, S. Lu, S. J. Shin, R. Quek, A. Ying, G. F. Ho, H. S. Nguyen, B. Dhabhar, V. Sriuranpong, M. L. Tiambeng, N. Prayogo, & N. Yamazaki, Sunrise in melanoma management: Time to focus on melanoma burden in Asia. AsiaPacific Journal of Clinical Oncology, 13 (2017) 423–427. https://doi.org/10.1111/ajco.12670. A. Tarbuk, A. M. *UDQFDULü & M. Šitum, Skin cancer and UV protection. Autex Research Journal, 16 (2016) 19–28. https://doi.org/10.1515/aut-2015-0050. A. H. Alasadi & B. Alsafy, Diagnosis of malignant melanoma of skin cancer types. International Journal of Interactive Multimedia and Artificial Intelligence, 4 (2017) 44.

Nanoparticles as Drug Delivery System in Melanoma

10.

11.

12.

13.

14. 15.

16.

17.

18.

19.

247

https://doi.org/10.9781/ijimai.2017.458. B. A. Krantz, N. Dave, K. M. Komatsubara, B. P. Marr, & R. D. Carvajal, Uveal melanoma: Epidemiology, etiology, and treatment of primary disease. Clinical ophthalmology (Auckland, N.Z.), 11 (2017) 279–289. https://doi.org/10.2147/OPTH.S89591. L. H. Mikkelsen, A. C. Larsen, C. von Buchwald, K. T. Drzewiecki, J. U. Prause, & S. Heegaard, Mucosal malignant melanoma – A clinical, oncological, pathological and genetic survey. Apmis, 124 (2016) 475–486. https://doi.org/10.1111/apm.12529. G. Malaguarnera, R. Madeddu, V. E. Catania, G. Bertino, L. Morelli, R. E. Perrotta, F. Drago, M. Malaguarnera, & S. Latteri, Anorectal mucosal melanoma. Oncotarget, 9 (2018) 8785–8800. https://doi.org/10.18632/oncotarget.23835. C. López-Camarillo, E. A. Ocampo, M. L. Casamichana, C. PérezPlasencia, E. Álvarez-Sánchez, & L. A. Marchat, Protein kinases and transcription factors activation in response to UV-radiation of skin: Implications for carcinogenesis. International Journal of Molecular Sciences, 13 (2012) 142–172. https://doi.org/10.3390/ijms13010142. B. Domingues, J. Lopes, P. Soares, & H. Populo, Melanoma treatment in review. ImmunoTargets and Therapy, 7 (2018) 35–49. https://doi.org/10.2147/ITT.S134842. J. Yang, D. K. Manson, B. P. Marr, & R. D. Carvajal, Treatment of uveal melanoma: Where are we now? Therapeutic Advances in Medical Oncology, 10 (2018) 175883401875717. https://doi.org/10.1177/1758834018757175. D. L. Cooper, C. M. Conder, & S. Harirforoosh, Nanoparticles in drug delivery: Mechanism of action, formulation and clinical application towards reduction in drug-associated nephrotoxicity. Expert Opinion on Drug Delivery, 11 (2014) 1661–1680. https://doi.org/10.1517/17425247.2014.938046. S. R. Mudshinge, A. B. Deore, S. Patil, & C. M. Bhalgat, Nanoparticles: Emerging carriers for drug delivery. Saudi Pharmaceutical Journal, 19 (2011) 129–141. https://doi.org/10.1016/j.jsps.2011.04.001. D. Bobo, K. J. Robinson, J. Islam, K. J. Thurecht, & S. R. Corrie, Nanoparticle-based medicines: A review of FDA-approved materials and clinical trials to date. Pharmaceutical Research, 33 (2016) 2373–2387. https://doi.org/10.1007/s11095-016-1958-5. S. Chowdhury, F. Yusof, W. W. A. W. Salim, N. Sulaiman, & M. O. Faruck, An overview of drug delivery vehicles for cancer treatment: Nanocarriers and nanoparticles including photovoltaic nanoparticles.

248

20. 21.

22.

23.

24.

25.

26.

27.

28.

Chapter 8

Journal of Photochemistry and Photobiology % Biology, 164 (2016) 151–159. https://doi.org/10.1016/j.jphotobiol.2016.09.013. A. H. Faraji & P. Wipf, Nanoparticles in cellular drug delivery. Bioorganic and Medicinal Chemistry, 17 (2009) 2950–2962. https://doi.org/10.1016/j.bmc.2009.02.043. S. Chaurasiya & V. Mishra, Biodegradable nanoparticles as theranostics of ovarian cancer: An overview. Journal of Pharmacy and Pharmacology, 70 (2018) 435–449. https://doi.org/10.1111/jphp.12860. K. Cho, X. Wang, S. Nie, Z. Chen, & D. M. Shin, Therapeutic nanoparticles for drug delivery in cancer. Clinical Cancer Research, 14 (2008) 1310–1316. https://doi.org/10.1158/1078-0432.CCR-071441. H. Zheng, G. Chen, L. A. DeLouise, & Z. Lou, Detection of the cancer marker CD146 expression in melanoma cells with semiconductor quantum dot label. Journal of Biomedical Nanotechnology, 6 (2010) 303–311. https://doi.org/10.1166/jbn.2010.1136. W. T. Al-Jamal, K. T. Al-Jamal, P. H. Bomans, P. M. Frederik, & K. Kostarelos, Functionalized-quantum-dot-liposome hybrids as multimodal nanopartides for cancer. Small, 4 (2008) 1406–1415. https://doi.org/10.1002/smll.200701043. E. M. Hersh, S. J. O’Day, A. Ribas, W. E. Samlowski, M. S. Gordon, D. E. Shechter, A. A. Clawson, & R. Gonzalez, A phase 2 clinical trial of nab-Paclitaxel in previously treated and chemotherapy-naive patients with metastatic melanoma. Cancer, 116 (2010) 155–163. https://doi.org/10.1002/cncr.24720. I. Mitrus, A. Sochanik, T. &LFKRĔ & S. Szala, Combination of combretastatin A4 phosphate and doxorubicin-containing liposomes affects growth of B16-F10 tumours. Acta Biochimica Polonica, 56 (2009) 161–165. G. J. Villares, M. Zigler, H. Wang, V. O. Melnikova, H. Wu, R. Friedman, M. C. Leslie, P. E. Vivas-Mejia, G. Lopez-Berestein, A. K. Sood, & M. Bar-Eli, Targeting melanoma growth and metastasis with systemic delivery of liposome-incorporated protease-activated receptor-1 small interfering RNA. Cancer Research, 68 (2008) 9078–9086. https://doi.org/10.1158/0008-5472.CAN-08-2397. Y. Chen, S. R. Bathula, Q. Yang, & L. Huang, Targeted nanoparticles deliver siRNA to melanoma. Journal of Investigative Dermatology, 130 (2010) 2790–2798. https://doi.org/10.1038/jid.2010.222.

Nanoparticles as Drug Delivery System in Melanoma

29.

30.

31.

32.

33.

249

M. A. Tran, R. Gowda, A. Sharma, E.-J. Park, J. Adair, M. Kester, N. B. Smith, & G. P. Robertson, Targeting V600EB-Raf and Akt3 Using nanoliposomal-small interfering RNA inhibits cutaneous melanocytic lesion development. Cancer Research, 68 (2008) 7638– 7649. https://doi.org/10.1158/0008-5472.CAN-07-6614. S. Pizzimenti, E. Ciamporcero, P. Pettazzoni, S. Osella-Abate, M. Novelli, C. Toaldo, M. Husse, M. Daga, R. Minelli, A. Bisazza, P. Ferruti, E. Ranucci, M. Grazia Bernengo, C. Dianzani, F. Biasi, R. Cavalli, & G. Barrera, The inclusion complex of 4-hydroxynonenal with a polymeric derivative of ȕ-cyclodextrin enhances the antitumoural efficacy of the aldehyde in several tumour cell lines and in a three-dimensional human melanoma model. Free Radical Biology and Medicine, 65 (2013) 765–777. https://doi.org/10.1016/j.freeradbiomed.2013.06.035. H. Yao, S. S. Ng, L.-F. Huo, B. K. C. Chow, Z. Shen, M. Yang, J. Sze, O. Ko, M. Li, A. Yue, L.-W. Lu, X.-W. Bian, H.-F. Kung, & M. C. Lin, Effective melanoma immunotherapy with interleukin-2 delivered by a novel polymeric nanoparticle. Molecular Cancer Therapeutics, 10 (2011) 1082–1092. https://doi.org/10.1158/15357163.MCT-10-0717. N. Clemente, B. Ferrara, C. L. Gigliotti, E. Boggio, M. T. Capucchio, E. Biasibetti, D. Schiffer, M. Mellai, L. Annovazzi, L. Cangemi, E. Muntoni, G. Miglio, U. Dianzani, L. Battaglia, & C. Dianzani, Solid lipid nanoparticles carrying temozolomide for melanoma treatment. Preliminary in vitro and in vivo studies. International Journal of Molecular Sciences, 19 (2018). https://doi.org/10.3390/ijms19020255. F. Masood, Polymeric nanoparticles for targeted drug delivery system for cancer therapy. Materials Science and Engineering C, 60 (2016) 569–578. https://doi.org/10.1016/j.msec.2015.11.067.

CHAPTER 9 ADVANCES IN LASERS AND NANOPARTICLES IN TREATMENT AND TARGETING OF EPITHELIAL-ORIGINATED CANCERS RAJAN RAJABALAYA*1 AND STACY DAVID2 1

PAPRSB Institute of Health Sciences, Universiti Brunei Darussalam, Brunei Darussalam 2 Department of Biology, Indiana University-Purdue University Indianapolis, IN 46202, USA. *Corresponding author: Dr Rajan Rajabalaya Senior Assistant Professor PAPRSB Institute of Health Sciences Universiti Brunei Darussalam Jalan Tungku Link BE1410 Bandar Seri Begawan Brunei Darussalam Email: [email protected] Phone: + 6732460922

Abstract Lasers and nanoparticles are currently being used to treat epithelialoriginated cancer. Relevant studies have shown that combining laser and nanoparticles may improve patient outcomes in epithelial-originated cancer. However, only a few studies have explored the combination of lasers and nanoparticles. Photo-thermal properties of ultrafast lasers generate nanoparticles, which may offer a fascinating combined therapeutic effect. The purpose of this chapter is to review the laser-assisted treatment of melanoma and nanoparticle for the treatment of cutaneous squamous cell carcinoma as well as basal cell carcinoma by using various nanoparticlesbased drug therapy. Combination of laser with gold and silver nanoparticles (AuNP & AgNPs) for the treatment of epithelial-originated cancer are also

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

251

explored. This chapter discusses about interaction of lasers as well as mechanism of laser nanoparticles in the tissues.

Abbreviations 5-FU AgNPs anti-EGFR anti-HER2 anti-VEGF-A APC AuNP AVEX BCG CEA cSCC CTCs CYP Doxo DR5-NP Dtxl EAC EGFR FAP FDA HNPCC IFN LPS mAb NIRF NLC OS PDT PEG PFS PLGA PlGF PNP PpIX PPTT

5-fluorouracil Silver nanoparticles Anti-epidermal growth factor receptor Anti-human epidermal growth factor receptor 2 Anti-vascular endothelial growth factor-A Adenomatous polyposis coli Gold Nanoparticles Avastin in Elderly with Xeloda Bacillus Calmette-Guérin Carcinoembryonic antigen Cutaneous squamous cell carcinoma Circulating tumour cells Cytochrome Doxorubicin Death receptor 5-specific antibodies Docetaxel Oesophageal adenocarcinoma cancer Epidermal Growth Factor Receptor Familial adenomatous polyposis Food and Drug Administration Hereditary nonpolyposis colorectal cancer Interferon-alpha Lipopolysaccharides Monoclonal antibodies Near-infrared fluorescence Nanostructured lipid carriers Overall survival Photodynamic therapy Polyethylene glycol Progression-free-survival Polylactic glycolic acid Placental growth factor Polymeric nanoparticles Protoporphyrin IX Plasmonic photothermal therapy

252

Ptxl RES RESOLV RESS ROS RR SCC SLN UCP

Chapter 9

Paclitaxel Reticuloendothelial system Rapid expansion of a supercritical solution into a liquid solvent Rapid expansion of a supercritical solution Reactive oxygen species Resection rate Squamous cell carcinoma Solid lipid nanoparticle Up conversion particles

1. Introduction Cancer is a significant public health concern globally, evidenced by data stating this disease as Brunei’s leading cause of death and America’s second [1]. Significant progress has been achieved in recent years to understand the possible features of the development and treatment of cancer. Nevertheless, the medical management of cancer with its rising occurrence remains a challenge for the 21st century [2].

1.1 Laser-assisted treatment of melanoma Despite their history, lasers are widely recognised as a breakthrough technology worldwide because they have revolutionised modern medicine and dentistry, such as cancer diagnosis and treatment [3]. Surgery, radiotherapy, and chemotherapy are the most commonly used cancer treatments. Typically, radiation and chemotherapy are offered as alternatives when the surgical removal of tumours is not feasible. Although they effectively reduce tumour size, these therapies can also damage healthy cells and may not necrotise the entire tumour. The optimal cancer treatment should destroy local tumour cells with little or no damage to the normal tissue and cause metastatic tumours to relapse and prevent their recurrence [4]. The primary long-term cancer regulation is the host immune surveillance and protection system [5]. Laser immunotherapy can be an effective treatment for cancer [4]. Laser immunotherapy was first developed and introduced in 1997. It combines locally targeted phototherapy and immunological stimulation with immunoadjuvant [4–5]. Laser immunotherapy aims to destroy tumours directly at the treatment site and induce tumour-specific immune responses to the host [5].

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

Figure 9-1: Laser-assisted treatment of melanoma Created with BioRender.com

253

254

Chapter 9

Laser therapy is only indicated in particular cases and after collecting and preserving histological specimens for the accurate histological diagnosis of pigmented melanocytic lesions on actinically damaged skin.

1.2 Nanoparticles Nanoparticles are ultra-dispersed solid supramolecular structures with lengths ranging from 10 to 1,000 ȝm. Nanoparticles are well matched to biological molecules and structures found within living cells on the size scale of 1–200 nm. Cellular organelles are approximately 100–300 nm, whereas intracellular proteins and molecules are approximately 10–50 nm. Nanoparticles appear to have the right size range for molecular-level imaging and manipulation [6]. Nanoparticles typically have a relatively high cell absorption compared to microparticles. They are accessible to a broader range of cell and intracellular targets due to their small size and mobility [7]. Nanoparticles are sphere-like biocompatible materials comprising inert silica, metal or crystals and are a few nanometres in size [6]. Carbon nanotubes, aluminium, copper, gold, iron, silver, silica, zinc, zinc oxide and titanium dioxide nanoparticles are the most widely produced [8]. These are similar to biological molecules, such as enzymes, receptors and antibodies [6]. Nanoparticles may imitate or modify biological processes. After opening endothelium tight junctions with hyperosmotic mannitol, nanoparticles can cross the blood-brain barrier, providing a sustained supply of therapeutic agents for difficult-to-treat diseases such as brain tumours [7]. Targeted drug delivery has enormous potential in improving cancer treatment by selectively delivering drug doses at the tumour site that are therapeutically active and protect normal tissues [9–10]. Targeted therapies are generally more effective than conventional therapies and exert less undesirable adverse effects [11]. The targeted delivery can be achieved either by active or passive targeting. Active targeting requires the therapeutic agent to be established by combining the therapeutic agent or carrier system with a ligand-specific cell or tissue. In contrast, passive targeting is accomplished by integrating the therapeutic agent into a macromolecule or nanoparticle that passively reaches the target organ [7].

2. Epithelial cancer Cancers are categorised based on various aspects, such as cell origin type [3]. Carcinoma involves cancers of epithelial cells (e.g., epithelial squamous cell cancer) and internal organs (e.g., lung cancer) or glands (e.g., breast

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

255

cancer) [3]. Various epithelial cells can grow into different carcinoma types [6].

2.1 Types of epithelial cancer The common types of carcinoma are squamous cell carcinoma (SCC), adenocarcinoma, transitional cell carcinoma, and basal cell carcinoma (BCC) [6]. SCC is classified by location, often found in the skin, head and neck, oesophagus, lung and cervix, and more rarely in the pancreas, thyroid, bladder and prostate. SCC derived from the epithelial tissues can be categorised as stratified squamous epithelium [12]. Cutaneous squamous cell carcinoma (cSCC) is also known as the SCC of the skin [13]. Adenocarcinomas begin as adenomatous cells in glandular cells. The glandular cells produce fluids to maintain the tissue’s moisture [14]. Adenocarcinoma was once regarded as an extremely rare histological form of oesophageal cancer [15] . Cells that can stretch with the expansion of an organ expand are transitional cells and are made up of the transitional epithelium tissue [14]. Therefore, cancers that begin in these cells are called transitional cell carcinoma. One example is the upper urinary tract transitional cell cancer, a rare disease that is highly challenging for the urologist [16]. Basal cell carcinoma is a malignant skin epithelial neoplasm that often develops in chronic exposure to the sun [17].

2.2 Current treatment of cutaneous squamous cell carcinoma Generally, surgical therapy, such as surgical excision, is the most effective in treating cSCC. Due to the potential for recurrence and metastasis, the guiding principle has relatively few exceptions, especially in the case of high-risk cSCC. When surgical treatment is not possible or preferred, if tumours are at low risk, non-surgical methods could be considered with the assumption that the level of cure may be lower. Further research is needed to determine the comparative safety and effectiveness of non-operative cSCC therapies [18].

Chapter 9

Type of treatment Photodynamic therapy

PK Absorption: Oral bioavailability is 50-60% Metabolism: Following topical administration, synthesis into PpIX takes place in situ in the skin

Drug used

ALA

Reference [18]

Side effect Erythema Pain/burning Irritation Oedema Pruritus Exfoliation

MOA ALA is a source of photoactive porphyrins, which accumulate with a photodynamic lamp in the skin lesions where the drug is added. The accumulated photoactive porphyrins create a photodynamic reaction when exposed to light, resulting in a cytotoxic process based on the simultaneous presence of oxygen; the singlet oxygen can then react to form superoxide and hydroxyl radicals.

PD ALA metabolism is the first step in the biochemical pathway that leads to the synthesis of heme. ALA is not a photosensitiser, but a protoporphyrin IX (PpIX) metabolic precursor, which is a photosensitiser. ALA synthesis is typically tightly controlled by enzyme feedback inhibition ALA synthetase, probably by rates of intracellular heme. ALA bypasses this control point when it is delivered to the cell and accumulates PpIX, which is converted by ferrochelatase into heme by adding iron to the PpIX nucleus.

Table 9-1: Pharmacology of the current drug treatment of cutaneous squamous cell carcinoma

256

Topical therapies

Imiquimod

MAL

Absorption: Generally absorbed (as a cream through the body. Half-life: 20 hours (topical), 2 hours (subcutaneous)

Absorption: In vitro, the mean cumulative absorption by human skin after 24 hours was 0.26% of the dose received Porphyrins may accumulate intercellularly in the treated skin lesions after topical application of methyl aminolevulinate. Intracellular porphyrins, which include PpIX, are photoactive, fluorescent compounds and form singlet oxygen after light activation in the presence of oxygen, causing damage to cellular compartments, particularly mitochondria. Accumulated porphyrin light activation leads to a photochemical reaction and phototoxicity to the target cells exposed to light. Imiquimod is a regulator of the immune response, which functions as an agonist toll-like receptor 7. Imiquimod is especially useful in places where surgery or other treatments, especially the face and lower legs, may be difficult, complicated or otherwise undesirable.

This is the precursor of photoactive porphyrins that accumulate in the skin lesions where the drug has been applied and are subsequently illuminated with a narrow red light spectrum with a light dose of 37 J/cm² from an Aktilite CL128 lamp. When exposed to the light, the accumulated photoactive porphyrins trigger a photodynamic reaction, resulting in a cytotoxic process based on the simultaneous presence of oxygen, resulting in single oxygen. The singlet oxygen can then react to radicals of superoxide and hydroxyl. Interferon- Į induction.

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

Asthenia Headaches Increase risk of infection Nausea Pain

Skin burning Erythema Pruritus Skin or eyelid oedema

[20]

[19]

257

Chemotherapy

258 Bioavailability: 28% Urinary secretion: less than 10% Bound in plasma: 8-12% Clearance: 16±7 mL/min/kg Volume of distribution: 0.25±.120 L/kg Half-life: 11±4 min Peak concentration: 11.2μM

Urinary excretion: 23±9% Clearance: 6.3±1.2 mL/min/kg Volume of distribution: 0.28±0.07 L/kg

5-FU

Cisplatin

5-FU is an antimetabolite antineoplastic. Antimetabolites masquerades as purine and pyrimidinebecoming DNA’s building blocks. During the ‘S’ phase, these compounds are stopped from becoming integrated into DNA, preventing normal development and division. It also blocks an enzyme that converts the cytosine nucleotide into the derivative of deoxy. Therefore, DNA synthesis is prevented as fluorouracil prevents the incorporation of the thymidine nucleotide into the DNA strand. In the class of alkylating agents, cisplatin is an antineoplastic used to treat different forms of cancer. This is because of its ability to add alkyl groups to many electronegative groups under cell conditions, called alkylating agents. Cross-linking guanine

Chapter 9

Cisplatin enters the cells by an active Cu² + transporter, CTR1, and easily degrades the transporter. ATP7A and ATP7B copper carriers and multidrug resistance protein 1 deliberately extrude the compounds from the cell. Inside the cell, the three analogues, chloride, cyclohexane, and oxalate

5-FU interferes with the synthesis of DNA by blocking deoxyuridine acid methylation with thymidylate acid.

Anaemia Arrhythmias Bone marrow failure Electrolyte imbalance

Alopecia Diarrhoea Nausea Skin reaction Stomatitis Vomiting

[21]

[20]

Panitumumab

Half-life: 4-11 days Clearance: 4.9±1.4 mL/kg/day

Half-life: 0.53±0.10 hours Peak concentration: 2 hr: 3.4±1.1 μg/mL 7 hr: 1.0± μg/mL bases in double-helix DNA strings prevents tumour developments from directly attacking DNA. It prevents the strands from uncoiling and splitting. Since this is necessary for the replication of DNA, the cells can no longer divide. Further, these drugs add methyl or other alkyl groups to molecules where they do not belong, which prevents their proper use through base pairing and contributes to DNA miscoding. Alkylating agents are cell-cycle nonspecific. Panitumumab, a recombinant human IgG2 kappa monoclonal antibody, binds specifically to the human Epidermal Growth Factor Receptor (EGFR). While EGFR is expressed in normal cells, in many human cancers, including colon and rectum, the overexpression of EGFR Panitumumab directly binds to EGFR on normal tumour cells and prevents ligand binding for EGFR competitively. Non-clinical studies show that panitumumab binding to EGFR prevents ligandinduced receptor autophosphorylation and activation of receptorassociated kinases. This

ligands, are replaced by water molecules, resulting in a highly reactive and positively charged molecule. The drug’s aquated species then interacts with nucleophilic sites on DNA and proteins in the main cytotoxic reaction.

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

Rash Dermatological toxicity Severe infusion reaction

[22]

259

Chapter 9

results in cell growth is observed. EGFR’s inhibition, apoptosis interaction with its initiation, decreased pronormal ligands results in inflammatory cytokine and phosphorylation and vascular growth factor activation of a series of development, and intracellular proteins, internalization of EGFR. which will control the transcription of genes involved in cell growth and survival, motility, and proliferation. Signal transduction via EGFR leads to activation of the wild-type KRAS gene. Still, the existence of an activating somatic mutation of the KRAS gene within a cancer cell may result in dysregulation of signalling pathways and resistance to EGFR inhibitor therapy. PK: Pharmacokinetic, PD: Pharmacodynamic, MOA: Mechanism of action, ALA: 5-aminoleculinic acid, MAL: Methyl aminolevulinate, 5-FU: 5fluorouracil

260

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

261

Available data on the treatment of patients with remote metastatic cSCC are incomplete and confined to clinical trials in Phase II. Chemotherapy has shown efficacy, such as cisplatin as a single agent or combined with 5fluorouracil (5-FU). However, the findings have not been verified in larger cohorts. A phase III trial in the head and neck SCC found that adding panitumumab to the combination of cisplatin and 5-FU enhanced progression-free survival but not overall survival [18].

2.3 Current treatment of oesophageal cancer Initial treatment methods for oesophageal adenocarcinoma cancer (EAC) are determined by several factors, such as the tumour stage and level, the tumour's location, the patient’s comorbidities and age, and clinical therapy experience [15]. Because the systemic treatment for EAC is often not curative, patients who may be eligible for curative endoscopic or surgical therapy should be correctly identified. Multimodal approaches are used in patients with late-stage illnesses. However, the importance of this method has not been thoroughly elucidated. In the USA, neoadjuvant chemotherapy and internal beam radiotherapy are the most common approaches for treating locally advanced EAC. In most trials, 5-FU (Table 9-1) was administrated to patients who were initially administered with cisplatin (Table 9-1) or carboplatin (Table 9-2). The simultaneous dosage of internal beam radiation applied ranged from 35 to 45 Grey. Many patients with locally advanced neoplasia are treated with surgery for curative reasons and to alleviate dysphagia and improve the quality of life of patients who could not be healed [15].

2.4 Current treatment of upper urinary tract transitional cell cancer Antegrade installation of the bacillus Calmette-Guérin (BCG) vaccine or mitomycin C in the upper urinary tract by percutaneous nephrostomy via a three-valve device open at 20 cm (after complete tumour eradication) is feasible after kidney sparing or cancer in situ therapy. Retrograde installation is also used through a ureteric stent. However, it can also be risky during instillation/perfusion due to possible ureteric obstruction and consequent pyelovenous reflux. The reflux obtained for a double J-stent has been used, but it is not recommended because it does not always penetrate the renal pelvis [23].

Chapter 9

Drug used

PK

PD

Carboplatin

Urinary excretion: 77±5 % Bound in plasma: 0% Clearance: 1.5±0.3 mL/min/kg Volume of distribution: 0.24±0.03 L/kg Half-life: 2±0.2 hours Peak time: 0.5 hours Peak concentration: 39 ± 17 μg/mL

Carboplatin is an alkylating agent class antineoplastic used to treat different forms of cancer. This is due to their ability to add alkyl groups to many electronegative groups under conditions in cells; they are called alkylating agents. Cross-linking guanine bases in double-helix DNA strings prevents tumour development from directly attacking DNA. It prevents the strands from uncoiling and splitting. Since this is necessary for the replication of DNA, the cells can no longer divide. Further, these drugs add methyl or other alkyl groups to molecules where they do not belong, which prevents their proper use through base pairing and contributes to DNA miscoding. Alkylating agents are cell-cycle nonspecific PK: Pharmacokinetic, PD: Pharmacodynamic, MOA: Mechanism of action

Type of treatment Chemotherapy

Table 9-2: Pharmacology of the current drug treatment of EAC

262

Side effect Alopecia Anaemia Asthenia Cardiovascular disease

MOA Carboplatin enters the cells by an active Cu² + transporter, CTR1, and easily degrades the transporter. ATP7A and ATP7B copper carriers and multidrug resistance protein 1 deliberately extrude the compounds from the cell. Inside the cell, the three analogues, chloride, cyclohexane, and oxalate ligands, are replaced by water molecules, resulting in a highly reactive and positively charged molecule. The drug’s aquated species then interacts with nucleophilic sites on DNA and proteins in the main cytotoxic reaction.

[15]

Reference

PK -

Absorption: Erratic Half-life: 8-48 min

Drug used

BCG vaccine

Mitomycin C

Mitomycin is one of the oldest chemotherapy drugs that have been around for decades and have been in use. It is an antibiotic that has been shown to be effective in anti-tumours. Mitomycin prevents DNA synthesis selectively. The quality of guanine and cytosine is associated with the degree of cross-linking caused by mitomycin.

-

PD The M. bovis strain the vaccine is immunologically similar to M. tuberculosis and induced cell-mediated immunity and stimulated M. tuberculosis infection. Can convey effective immunity through stimulation of endogenous antibodies production. Mitomycin is activated in vivo as an alkylating agent that is bifunctional and trifunctional. Binding to DNA contributes to DNA synthesis and work crosslinking and inhibition. Mitomycin is phasenonspecific in the cell cycle.

MOA [23]

[23–24]

Skin reactions Intravenous use: bone marrow depression, cough, dyspnoea

Reference

263

Increased risk of infection osteitis

Side effect

PK: Pharmacokinetic, PD: Pharmacodynamic, MOA: Mechanism of action, BCG vaccine: Bacillus Calmette-Guérin vaccine

Types of therapy Adjuvant topical agents

Table 9-3: Pharmacology of the current drug treatment of upper urinary tract transitional cell cancer

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

Chapter 9

264

2.5 Current treatment of basal cell carcinoma BCC is a relatively common condition and is routinely administered in the care of outpatients. Early diagnosis is a prerequisite for a better prognosis based on sound knowledge and timely-coordinated and effective care. Despite the slow development and various therapeutic approaches such as cryotherapy, Mohs surgery and Roentgen therapy, BCC should not be underestimated. Most treatment methods do not automatically dissolve the risk of relapses. If the case is ignored, left unchecked or obtained, and no appropriate therapy is administered, BCC may kill the underlying tissues and spread metastases. The treatment includes 5-FU (Table 9-1), Imiquimod (Table 9-1) and interferon-alpha (Table 9-4) [24]. Table 9-4: Pharmacology of the current drug treatment for basal cell carcinoma Types of therapy

Dru g used

Pharmacokinetics

Mechanism of action

Side effect

Referenc e

Interferon -alpha (IFN)

IFN

Route of elimination: Reticuloendothelia l system, kidneys and liver.

Natural IFN includes the composition of several subtypes typical of the human body’s interferon. This is believed to result in a wider range of direct anti-viral and immunoregulator y activity, with the subtypes working synergistically to respond.

Alopecia Anaemia Anxiety Abnorma l appetite

[24]

3. Principles of nanotechnology Nanotechnology involves the manipulation of matter on a near atomic scale to produce new structures, devices, and materials. Nanotechnology deals with nanomaterials ranging from 1 to 100 nm in at least one dimension [25]. At the nanoscale, fundamental mechanical, electrical, optical, and other properties can significantly differ from their bulk material counterparts. Nanotechnology is expected to achieve significant progress in detecting treatment and prevention of diseases. Nanotechnology hopes to enhance

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

265

available treatment approaches by working at least on two fundamental levels: conferring new properties on a pharmaceutical and directly targeting the agent at the site of operation [26].

3.1 Nanomedicine nanoparticles for cancer therapy Nanomedicine aims to use nanomaterials properties and physical characteristics to diagnose and treat diseases at a molecular level. Nanomedicine is one of the most promising fields for potential new advances in medical engineering with the body at the cellular and molecular levels. This technology also revolutionises medical areas, such as monitoring, diagnosis, treatment and prevention, cell drug delivery, and positioning it as a revolution in medical science and health care [26]. The use of nanoparticles in cancer treatment has two main aspects: diagnosis and therapy. Ideal therapy nanoparticles can enhance the aggregation and release of active agents at the neoplastic site, increase therapeutic efficacy, and reduce the incidence and severity of side effects by reducing their positron in healthy tissues. Passive-targeted nanoparticles can increase aggregation at pathological sites, whereas active targeting increases the absorption by the neoplastic cells. Nanoparticles that react to the environment can improve therapeutic agents’ release rate. Nanoparticles could avoid healthy tissues by combining these strategies while killing tumour cells [27].

4. Principles of laser in medicine The laser application is focused on the interaction of laser radiation with biological tissues in medical treatment. Laser radiation can be included in a large category of electromagnetic radiation emitted by many radiation sources such as the sun, fire, bulbs, lamps, electrical discharge, and plasma. With the advancement of laser physics and the subsequent discovery of other aspects of laser technology, such as the generation of new wavelengths, radiation with various energy levels, high power and slight beam divergence, a new field of science has begun to develop with ‘laser medicine’ applications. The development of new methods of using lasers and the discovery of new forms of lasers inspired new medical treatments [28].

4.1 Interaction of lasers with tissues Human tissue is a heterogeneous material that comprises many different components. It can be divided into hard tissue, soft tissue and biological fluids using an elementary classification [28].

Chapter 9

266

4.1.1 Hard tissue These include bone, dental enamel or dentine, and calcified tissue plates. Hard tissue can be characterised by its water content and the hydroxyapatite OH radicals concerning its contact with radiation. 4.1.2 Soft tissue Soft tissues are mainly muscles, nervous tissue, skin and fatty tissue. Soft tissue can be opaque and transparent and comprise water, the main chromophore, a substance or tissue element that absorbs a particular radiation wavelength. Soft tissue can also be made up of many other molecules, such as melanin pigment, lipids and carbohydrates [28]. 4.1.3 Biological fluids These are predominantly expressed by blood, including water and haemoglobin proteins, leukocytes, thrombocytes and blood cells. The effects of tissue-radiation interactions are unique in each tissue due to the complexity of the various tissue types and their components. Molecules absorb radiation photons with specific wavelengths, and their behaviour after absorption is determined by interacting radiation’s output characteristics: its fluency, intensity, power and others [28].

5. The mechanism of laser Particulate contact with laser or ultrasonic radiation can result in localised cavitation that may mediate selective tumour therapy. Local heating of heavily absorbing particles by pulsed laser radiation may accomplish laserinduced cavitation. The cavitation caused by ultrasound is produced by the interaction of ultrasonic radiation with particles showing a lower cavitation threshold. The cavitation induced by the laser or ultrasound can perforate tumour blood vessel walls and cancer cell membranes and cause microconvection in the interstitial space [29].

5.1 Mechanism of laser and nanoparticle The penetration of anticancer drugs (especially macromolecular agents) from blood is limited because of physiological barriers, such as tumour capillary tumour walls and cancer cell membranes. Exogenous nano- or microparticles interacting with laser or ultrasonic radiation may increase the drug delivery in tumour cells because of cavitation induced by laser or

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

267

ultrasound [29]. The known photothermal therapy (PTT), converts laser light into heat that can target and kill tumour cells. Another technique, photodynamic therapy (PDT), uses laser light to generate reactive oxygen species (ROS), such as hydroxyl radicals, singlet oxygen, superoxide radicals, and hydrogen peroxide, which can wreak devastation on tumour cells. By combining laser light with nanomaterials, researchers have been able to deliver drugs to sites in the body. The surface of the nanoparticle can be modified to attach a photosensitive molecule to the surface. In combination of laser and nanoparticle to be successfully produce a sufficient ambient oxygen must be present to produce enough ROS to kill tumour cells. The nanoparticles can be used to deliver chemotherapeutic agents or antibiotics to the tumour site. When light is applied, generating ROS molecules in the tumour kills tumour cells and bacteria. The antibiotics can be released to prevent infection in the treated area. This method may provide extremely localised damage in a controlled way, ranging from a few nanometres to tens of microns (single size of cancer cell), without damaging the healthy tissue around it by carefully designing the laser wavelength, pulse duration, particle size and shape [30]. Plasmonic photothermal therapy (PPTT) is an approach that has been developed over the last decade. Researchers have shown nanoparticle resonance spectral tuning to the ‘therapeutic optical window’ (from 750 to 1,100 nm) and achieving a fair ratio between absorption and scattering efficiencies. Before the laser treatment, refractive index matching agents, such as poly (ethylene glycol) diacrylate, can be added to the tumour surface to minimise the backreflected light intensity from the surface layers of the skin [31].

5.2 Combination of lasers and nanoparticles for treatment of epithelial-originated cancer Novice methods, such as hormone therapy, photodynamic therapy (PDT), nanoparticle therapy and, increasingly, laser and nanoparticles combinations, have been applied through recent advances in medical science and new technology. PDT relies on oxygen availability in tumours. However, there is no restriction on using lasers and nanoparticles, and PDT can be used as an alternative. Gold and silver nanoparticles are highly regarded metallic nanoparticles in the biomedical field [3]. The unique characteristics of lasers, such as photothermal properties and the extremely small size of nanoparticles, provide a fascinating combination of therapeutic effects. Several nanoparticles have their unique properties and applications, such as nanoring, nanoshells, nanorods, nanopores, and

268

Chapter 9

nanowires. Different types of lasers are used based on the peak absorption of nanoparticles [3].

5.3 Gold nanoparticles (AuNP) in combination with laser The advent of nanotechnology revealed that gold’s physicochemical properties makes it a perfect material for nanoparticle fabrication. Because of its unique electronic, optical, thermal, chemical and biological properties, gold nanoparticles (AuNP) have attracted significant attention [32]. Because they absorb laser light effectively, are non-toxic, combine readily with proteins and antibodies, and have changeable optical characteristics, AuNPs are the best choice for photothermal sensitisation between nanostructures. Different types of nanoparticles were used for various experiments, such as silica gold nanoshells. Such nanoparticles are well absorbed in NIR spectra, producing the most conversion and the least light reflection in essential tissues. Several types of research works were conducted in this area [3].

5.4 Silver nanoparticles (AgNPs) use in combination with laser Silver nanoparticles (AgNPs) with multiple valence antimicrobial activity have been commonly used as antimicrobial agents in different industries. They have recently made their way into cancer therapies [3,37]. When tested on living cells, AgNPs captivatingly exhibited dual action. This inhibited the growth and division of tumour cells and their nuclei while being biocompatible with healthy ones. Other recent results demonstrated that AgNPs of different sizes could increase the magnetic thermo-sensitivity of glioma cells depending on their size [3].

Laser characteristics

NIR laser

Laser light (Med Art, Hvidovre, Denmark) for 2 minutes at 820 nm and 4 W/cm²

Target cells

Human MDAMB-231 cancer of the breast A2780, cancer cells of the ovaries

HER2-positive KB cells HER2-negative HeLaS3

Gold-silica nanoshell paired with nanobody anti-HER2 100 nm

Nanoparticle characteristics 1:3:1 (NP3) ratio of DOX@PEG-HAuNS

Table 9-5: Studies of combination of Au nanoparticles with laser

In vitro, on treatment with NIR laser, NP3 mediated PTA of both cancer cells and released doxorubicin. In vivo, NP3 showed slower blood clearance and higher tumour aggregation than free doxorubicin. Greater operation of anti-tumours. There has been a significant reduction in systemic toxicity. Enhanced the efficacy of antitumours. There was substantial cell death in the cultures of tumours in the KB No signs of cell damage or death in cell cultures of HeLaS3

Result

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

[3,34]

[3,33]

References

269

Laser characteristics

NIR

Near-infrared fluorescence MRI

Target cells

Mouse

Subcutaneous tumours in animal models of breast cancer

270

Gold nanoshells, spherical nanoparticles with silica base and shells of gold, nanoshells mixed with DNA

Nanoparticle characteristics 40-45 nm of gold colloidal nanospheres

Chapter 9

Through AuNP tumour aggregation, the use of T-cell chaperones for AuNP delivery can boost the efficacy of nanoparticle-based therapies and imaging applications. Nanoshells with double-stranded DNA also provide a way to deliver small molecules to cells.

Result

[3,36]

[3,35]

References

Ti: Sapphire laser Wavelength 800nm

NIR femtoseconds three specific laser average powers: 1.2,0.6, 0.12 mW

KB cancer cells, a cell line of the human epidermoid

Laser characteristics

Non-small cells of the lung (NCI-H460)

Target cells

Silver-dendrimer composite nanodevice

Nanoparticle characteristics Chitosan-coated silver nanotriangles

Result Biocompatible plasmonic nanoparticles with a disproportionate potential to the conducted in opposition to many cancers as active photo therapeutic sources Significant decrease in the breakdown threshold; selective promotion of optical breakdown intracellular laser-induced was observed

Table 9-6: Studies of the combination of Ag nanoparticles with lasers

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

[3,39]

[3,38]

References

271

5.5 Combination of Au/Ag with lasers

Chapter 9

Cells of SK-BR-3 (HER2/negative breast cancer cells) and H520 (HER2/negative lung cancer)

(NB-4) cells

Target cells

NIR region: 800nm 35 W/cm² Ti: Sapphire laser pulse for 7 min

8.5 x 10Ś W/m² laser exposure

Laser characteristics

Anti-EGFR-conjugated Nanostructures of $Xႂೋႃ$Jႂႆႀೋႃ with dendrite

Nanoparticle characteristics Au-Ag nanorods

Table 9-7: Studies of combinations of Au/Ag with lasers

The hollow $Xၵǿ $Jၵၼ nanostructured dendrites have the potential to kill cancer cells in photothermolysis.

Au-Ag nanorod combination provides targeted tumour cells with selective and effective photothermal killing. The tumour tissue will be selectively killed at laser levels that will not damage the normal tissue around it.

Result

[3,36]

[3,40]

References

Only two papers have addressed laser application with a mixture of Ag and Au nanoparticles, probably because both elements have similar advantages.

272

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

273

6. Diagnosis of epithelial cancer using laser Scientists have directed more attention towards near-infrared fluorescence (NIRF) imaging in recent years. Although light at the wavelength of 600– 1,100 nm has a high tissue penetration, the background of spontaneous fluorescence is strong. Most of the upconversion particles (UCP) research has been conducted exclusively using rare-earth metal particles. New molecular imaging examinations now suitable for NIRF imaging need to have high sensitivity, high fluorescence stability, good safety and high energy of light conversion, also known as the anti-Stokes effect. Chatterjee et al. published a report on NaYF4: YbUCP coated with Er used for tumour cell and small animal imaging. Their results showed good compatibility of these UCPs and observed no toxicity to stem cells of the bone marrow. Nonetheless, to detect cancer cells in general, the nanoparticles need to work on the surface, which is usually done by coupling antibodies. Though numerous studies have tried to create UCPs with rare-earth ions, the viability of tumour cells in vivo imaging still needs to be studied [41]. Coated UCPs continued to agglomerate and precipitate, whereas yields of UCPs were reduced in the aqueous phase. Tumours in tumour-bearing mice could thus be successfully identified. Nanoparticles can enter cells through passive absorption and active uptake pathway. Unfortunately, these entry patterns have significant limitations in cancer diagnosis due to the lack of tolerance of nanoparticles to the cancer cells. The nanoparticles must therefore be combined with molecules attacking cancer cells. Therefore, the biological function of the sample is modified after the coupling of the molecular probe with the nanoparticles [41]. Another diagnostic tool is an immunosensor. An important area of the immunosensor design is diagnosing cancer by identifying particular tumour biomarkers [42]. These biomarkers are a type of biochemical substance formed by human tumour tissues that may indicate the presence and growth of tumours in the body. Cancer biomarkers have important medical significance in the early screening of tumours, the assessment of the disease stage, the choice of appropriate therapy and the curative effect identification. Circulating tumour cells (CTCs) flow from the primary tumour into the vasculature and circulate through the bloodstream. Therefore, CTCs are seeds for the subsequent growth of disseminated tumour mass (metastasis) in remote vital organs, activating a process responsible for most cancerrelated deaths. The identification of CTCs can thus have crucial prognostic and therapeutic implications [42].

274

Chapter 9

7. Conclusion Advances in lasers and nanoparticles are essential for treating and targeting epithelial-originated cancer. However, they are only limited studies on this topic. Thus, further research concerning various types of nanoparticles for the application of lasers to treat epithelial-originated cancer is needed to improve the problems of cancer therapy. Moreover, in vivo and in vitro studies should be considered while investigating this issue.

References 1. 2.

3.

4.

5.

6.

7. 8.

R. L. Siegel, K. D. Miller, & A. Jemal, Cancer statistics, 2019. &$ A Cancer Journal for Clinicians, 69 (2019) 7–34. https://doi.org/10.3322/caac.21551. R. Baskar, K. A. Lee, R. Yeo, & K.-W. Yeoh, Cancer and radiation therapy: Current advances and future directions. International Journal of Medical Sciences, 9 (2012) 193–199. https://doi.org/10.7150/ijms.3635. R. Fekrazad, N. Naghdi, H. Nokhbatolfoghahaei, & H. Bagheri, The combination of laser therapy and metal nanoparticles in cancer treatment originated from epithelial tissues: A literature review. Journal of Lasers in Medical Sciences, 7 (2016) 62–67. https://doi.org/10.15171/jlms.2016.13. S. Long, Y. Xu, F. Zhou, B. Wang, Y. Yang, Y. Fu, N. Du, & X. Li, Characteristics of temperature changes in photothermal therapy induced by combined application of indocyanine green and laser. Oncology Letters, (2019). https://doi.org/10.3892/ol.2019.10058. X. Li, H. Le, R. F. Wolf, V. A. Chen, A. Sarkar, R. E. Nordquist, H. Ferguson, H. Liu, & W. R. Chen, Long-term effect on emt6 tumours in mice induced by combination of laser immunotherapy and surgery. Integrative Cancer Therapies, 10 (2011) 368–373. https://doi.org/10.1177/1534735410387597. R. Ankathil, V. S. Nair, A. Ravindran, & R. Ravindran, Conspectus on nanotechnology in oral cancer diagnosis and treatment. Multifunct. Syst. Comb. Deliv. Biosensing Diagnostics (Elsevier, 2017) 31–49. https://doi.org/10.1016/B978-0-323-52725-5.00003-4. R. Singh & J. W. Lillard, Nanoparticle-based targeted drug delivery. Experimental and Molecular Pathology, 86 (2009) 215–223. https://doi.org/10.1016/j.yexmp.2008.12.004. A. Tymoszuk & N. Miler, Silver and gold nanoparticles impact on in vitro adventitious organogenesis in chrysanthemum, gerbera and

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

9. 10. 11.

12. 13.

14.

15. 16.

17.

18.

275

Cape Primrose. Scientia Horticulturae, 257 (2019) 108766. https://doi.org/10.1016/j.scienta.2019.108766. J. K. Vasir & V. Labhasetwar, Targeted drug delivery in cancer therapy. Technology in Cancer Research & Treatment, 4 (2005) 363–374. https://doi.org/10.1177/153303460500400405. R. K. Jain, Delivery of molecular and cellular medicine to solid tumours. Microcirculation, 4 (1997) 1–23. https://doi.org/10.3109/10739689709148314. B. Bahrami, M. Hojjat-farsangi, H. Mohammadi, & E. Anvari, Nanoparticles and targeted drug delivery in cancer therapy. Immunology Letters, 190 (2017) 64–83. https://doi.org/10.1016/j.imlet.2017.07.015. A. Sánchez-Danés & C. Blanpain, Deciphering the cells of origin of squamous cell carcinomas. Nature Reviews Cancer, 18 (2018) 549– 561. https://doi.org/10.1038/s41568-018-0024-5. L. J. L. Forbes, F. Warburton, M. A. Richards, & A. J. Ramirez, Risk factors for delay in symptomatic presentation: A survey of cancer patients. British Journal of Cancer, 111 (2014) 581–588. https://doi.org/10.1038/bjc.2014.304. A. Van Keymeulen & C. Blanpain, Tracing epithelial stem cells during development, homeostasis, and repair. Journal of Cell Biology, 197 (2012) 575–584. https://doi.org/10.1083/jcb.201201041. J. H. Rubenstein & N. J. Shaheen, Epidemiology, Diagnosis, and management of esophageal adenocarcinoma. Gastroenterology, 149 (2015) 302–317. https://doi.org/10.1053/j.gastro.2015.04.053. M. M. Elawdy, Y. Osman, D. E. Taha, S. El-halwagy, M. A. Elhamid, & R. T. Abouelkheir, Long-term outcomes of upper tract urothelial carcinoma: A retrospective evaluation of single-center experience in 275 patients. Türk Üroloji Dergisi/Turkish Journal of Urology, 45 (2019) 177–182. https://doi.org/10.5152/tud.2019.02185. D. T. Netscher & M. Spira, Basal cell carcinoma: An overview of tumour biology and treatment. Plastic and Reconstructive Surgery, 113 (2004) 74e-94e. https://doi.org/10.1097/01.PRS.0000113025.69154.D1. J. Y. S. Kim, J. H. Kozlow, B. Mittal, J. Moyer, T. Olenecki, P. Rodgers, M. Alam, A. Armstrong, C. Baum, J. S. Bordeaux, M. Brown, K. J. Busam, D. B. Eisen, V. Iyengar, C. Lober, D. J. Margolis, J. Messina, A. Miller, S. Miller, E. Mostow, C. Mowad, K. Nehal, K. Schmitt-Burr, A. Sekulic, P. Storrs, J. Teng, S. Yu, C.

276

19.

20.

21.

22.

23.

24.

25.

26.

Chapter 9

Huang, K. Boyer, W. S. Begolka, & C. Bichakjian, Guidelines of care for the management of cutaneous squamous cell carcinoma. Journal of the American Academy of Dermatology, 78 (2018) 560– 578. https://doi.org/10.1016/j.jaad.2017.10.007. A. M. Martins, J. M. Marto, J. L. Johnson, & E. M. Graber, A review of systemic minocycline side effects and topical minocycline as a safer alternative for treating acne and rosacea. Antibiotics, 10 (2021) 757. https://doi.org/10.3390/antibiotics10070757. G. T. Prince, M. C. Cameron, R. Fathi, & T. Alkousakis, Topical 5fluorouracil in dermatologic disease. International Journal of Dermatology, 57 (2018) 1259–1264. https://doi.org/10.1111/ijd.14106. J. Fu, C. Li, Y. Liu, M. Chen, Q. Zhang, X. Yu, B. Wu, J. Li, L. Du, Y. Dang, D. Wu, M. Wei, Z. Lin, & X. Lei, The microneedles carrying cisplatin and IR820 to perform synergistic chemophotodynamic therapy against breast cancer. Journal of Nanobiotechnology, 18 (2020) 146. https://doi.org/10.1186/s12951020-00697-0. O. Bouché, M. Ben Abdelghani, J.-L. Labourey, S. Triby, R.-J. Bensadoun, T. Jouary, & G. Des Guetz, Management of skin toxicities during panitumumab treatment in metastatic colorectal cancer. World Journal of Gastroenterology, 25 (2019) 4007–4018. https://doi.org/10.3748/wjg.v25.i29.4007. M. Rouprêt, M. Babjuk, M. Burger, O. Capoun, D. Cohen, E. M. Compérat, N. C. Cowan, J. L. Dominguez-Escrig, P. Gontero, A. Hugh Mostafid, J. Palou, B. Peyronnet, T. Seisen, V. Soukup, R. J. Sylvester, B. W. G. van Rhijn, R. Zigeuner, & S. F. Shariat, European Association of Urology guidelines on upper urinary tract urothelial carcinoma: 2020 update. European Urology, 79 (2021) 62–79. https://doi.org/10.1016/j.eururo.2020.05.042. L. Dourmishev, D. Rusinova, & I. Botev, Clinical variants, stages, and management of basal cell carcinoma. Indian Dermatology Online Journal, 4 (2013) 12. https://doi.org/10.4103/2229-5178.105456. H. Calipinar & D. Ulas, Development of nanotechnology in the world and nanotechnology standards in Turkey. Procedia Computer Science, 158 (2019) 1011–1018. https://doi.org/10.1016/j.procs.2019.09.142. J. K. Patra, G. Das, L. F. Fraceto, E. V. R. Campos, M. del P. Rodriguez-Torres, L. S. Acosta-Torres, L. A. Diaz-Torres, R. Grillo, M. K. Swamy, S. Sharma, S. Habtemariam, & H.-S. Shin, Nano

Advances in Lasers and Nanoparticles in Treatment and Targeting of Epithelial-Originated Cancers

27.

28. 29.

30.

31.

32.

33.

34.

35.

277

based drug delivery systems: Recent developments and future prospects. Journal of Nanobiotechnology, 16 (2018) 71. https://doi.org/10.1186/s12951-018-0392-8. B. García-Pinel, C. Porras-Alcalá, A. Ortega-Rodríguez, F. Sarabia, J. Prados, C. Melguizo, & J. M. López-Romero, Lipid-based nanoparticles: Application and recent Advances in cancer treatment. Nanomaterials, 9 (2019) 638. https://doi.org/10.3390/nano9040638. H. Jelínková, Introduction: The history of lasers in medicine. Lasers Med. Appl. (Elsevier, 2013), 1–13. https://doi.org/10.1533/9780857097545.1. R. O. Esenaliev, I. V. Larina, K. V. Larin, M. Motamedi, & B. M. Evers, Mechanism of laser-induced drug delivery in tumours. In D.D. Duncan, J.O. Hollinger, & S.L. Jacques, eds., (2000), 188. https://doi.org/10.1117/12.388045. R. R. Letfullin, C. Joenathan, T. F. George, & V. P. Zharov, Laserinduced explosion of gold nanoparticles: Potential role for nanophotothermolysis of cancer. Nanomedicine, 1 (2006) 473–480. https://doi.org/10.2217/17435889.1.4.473. G. S. Terentyuk, I. L. Maksimova, N. I. Dikht, A. G. Terentyuk, B. N. Khlebtsov, N. G. Khlebtsov, & V. V. Tuchin, Cancer laser therapy using gold nanoparticles. Lasers Med. Appl. (Elsevier, 2013), 659–703. https://doi.org/10.1533/9780857097545.4.659. K. Kalimuthu, B. S. Cha, S. Kim, & K. S. Park, Eco-friendly synthesis and biomedical applications of gold nanoparticles: A review. Microchemical Journal, 152 (2020) 104296. https://doi.org/10.1016/j.microc.2019.104296. J. You, R. Zhang, G. Zhang, M. Zhong, Y. Liu, C. S. Van Pelt, D. Liang, W. Wei, A. K. Sood, & C. Li, Photothermal-chemotherapy with doxorubicin-loaded hollow gold nanospheres: A platform for near-infrared light-trigged drug release. Journal of Controlled Release, 158 (2012) 319–328. https://doi.org/10.1016/j.jconrel.2011.10.028. F. Sheikholeslami, R. Fekrazad, Rasaee, S. Ardestani, Hakimiha, Farokhi, & Kalhori, Treatment of oral squamous cell carcinoma using anti-HER2 immunonanoshells. International Journal of Nanomedicine, (2011) 2749. https://doi.org/10.2147/IJN.S24548. L. C. Kennedy, A. S. Bear, J. K. Young, N. A. Lewinski, J. Kim, A. E. Foster, & R. A. Drezek, T cells enhance gold nanoparticle delivery to tumours in vivo. Nanoscale Research Letters, 6 (2011) 283. https://doi.org/10.1186/1556-276X-6-283.

278

36.

37.

38.

39.

40.

41.

42.

Chapter 9

R. Bardhan, S. Lal, A. Joshi, & N. J. Halas, Theranostic nanoshells: From probe design to imaging and treatment of cancer. Accounts of Chemical Research, 44 (2011) 936–946. https://doi.org/10.1021/ar200023x. C. Soica, I. Pinzaru, C. Trandafirescu, F. Andrica, C. Danciu, M. Mioc, D. Coricovac, C. Sitaru, & C. Dehelean, Silver-, gold-, and iron-based metallic nanoparticles. Des. Nanostructures Theranostics Appl. (Elsevier, 2018), pp. 161–242. https://doi.org/10.1016/B9780-12-813669-0.00005-1. S. C. Boca, M. Potara, A.-M. Gabudean, A. Juhem, P. L. Baldeck, & S. Astilean, Chitosan-coated triangular silver nanoparticles as a novel class of biocompatible, highly effective photothermal transducers for in vitro cancer cell therapy. Cancer Letters, 311 (2011) 131–140. https://doi.org/10.1016/j.canlet.2011.06.022. C. Tse, M. J. Zohdy, J. Y. Ye, M. O’Donnell, W. Lesniak, & L. Balogh, Enhanced optical breakdown in KB cells labeled with folatetargeted silver-dendrimer composite nanodevices. 1DQRPHGLFLQH Nanotechnology, Biology and Medicine, 7 (2011) 97–106. https://doi.org/10.1016/j.nano.2010.09.003. Y.-F. Huang, K. Sefah, S. Bamrungsap, H.-T. Chang, & W. Tan, Selective photothermal therapy for mixed cancer cells using aptamer-conjugated nanorods. Langmuir, 24 (2008) 11860–11865. https://doi.org/10.1021/la801969c. G. Liao, L. Wang, & W. Yu, Application of novel targeted molecular imaging probes in the early diagnosis of upper urinary tract epithelial carcinoma. Oncology Letters, (2018). https://doi.org/10.3892/ol.2018.9430. M. A. Fernández-Baldo, F. G. Ortega, S. V. Pereira, F. A. Bertolino, M. J. Serrano, J. A. Lorente, J. Raba, & G. A. Messina, Nanostructured platform integrated into a microfluidic immunosensor coupled to laserinduced fluorescence for the epithelial cancer biomarker determination. Microchemical Journal, 128 (2016) 18–25. https://doi.org/10.1016/j.microc.2016.03.012.

CHAPTER 10 ADVANCES IN NANOMATERIALS FOR ENHANCED PHOTODYNAMIC THERAPY MONOSHA PRIYADARSHINI,1,3 DHANASHREE MURUGAN,2,3 LOGANATHAN RANGASAMY,3 N. ARUNAI NAMBI RAJ*3 1

School of Advanced Sciences, Vellore Institute of Technology, Vellore 632014, India 2 School of Biosciences and Technology, Vellore Institute of Technology, Vellore 632014, India 3 Centre for Biomaterials, Cellular and Molecular Theranostics, Vellore Institute of Technology, Vellore 632014, India *Corresponding author: Dr N. Arunai Nambi Raj Professor of Physics Centre for Biomaterials, Cellular and Molecular Theranostics Vellore Institute of Technology Vellore 632014, Tamil Nadu, India Email: [email protected] Phone: +919443627064

Abstract Knowledge and best practices in cancer research, which has delved into the history of medicines and the cutting-edge innovation with a gazillion of ideas and advanced techniques to achieve a cold goal, are constantly evolving. Many cancers were not identified because they are related to the political multiplication of cells, always activated by carcinogens. However, there has always been a global quest to disentangle them from living beings. Photodynamic therapy is one of the modern, non-invasive and novel methods used for curing malignant tumours by injecting non-toxic chemical

280

Chapter 10

compounds into the targeted tissues that morph into cancer cells by leaving the normal cells alone. When the light of the appropriate wavelength, depending on the target area, is projected, energy transfers cascade and yield to form reactive oxygen, forcing the cells to undergo apoptosis and necrotic cell death. The photosensitisers, lights and tissue oxygenation determine the photodynamic therapy efficacies for a particular patient. This chapter will discuss the important aspects of the history, action mechanism, principle and clinical approaches of photodynamic therapy, such as the types, inclusions, delivery strategies of photosensitisers and principles of light mechanisms. Additionally, the glory and refinement of nanomedicines allowed the integration of the agents responsible for imaging to help deliver light modules for the past short time. The chapter will highlight the accelerative usage of nanomedicines in photodynamic therapy in the current art state. It precisely narrates the in vitro and in vivo research conducted effectively and offers an outlook on the potentials and challenges of photodynamic therapy concerning successful conversion into enormous clinical applications.

Abbreviations 5-FU ALA FRET HA HAL IARC MAL MTD mTHPP ORMOSIL PGA PHPP PLA PLGA SLP TPC UCNPs

5-fluorouracil Aminolaevulinic acid Fluorescence resonance energy transfer Hyaluronic acid Hexaminolevulinate International Agency for Research on Cancer Methyl-5- aminolevulinate Maximum tolerated dose 5,10,15,20-tetrakis(m-hydroxyphenyl) porphyrin Organically modified silica Polyglycolide 2,7,12,18-tetramethyl-3,8-di-(1-propoxyethyl)- 13,17bis-(3-hydroxypropyl) porphyrin Polylactide Polylactic glycolic acid Synthetic long peptides 5-(4-carboxyphenyl)-10,15,20-triphenylchlorin Up conversion nanoparticles

Advances in Nanomaterials for Enhanced Photodynamic Therapy

281

1. Introduction Cancer affects more lives than we realise. Given the prevalence of cancer in our culture, it is reasonable to wonder how many of us truly understand cancer, how it develops and how to treat and cure it successfully[1]. Cancer: Step Outside the Box is just a healthy cell that turns out to be unwholesome because of wild and uncharacterised cell growth when there is a defect at the DNA level. This is due to pathogenic coverage, toxins, or some other malignant aspect causing them to go awry. Angiogenesis is initiated by cancer cells’ anaerobic nature, which causes them to form their new blood vessels, which constantly drain energy and deprive healthy portions of the body of oxygen and glucose to survive[2]. According to Global Cancer Statistics 2020, the International Agency for Research on Cancer (IARC) reported an estimated 2.3 million new cases of cancer (11.7%) in the following order: lung (11.4%), colorectal (10.0%), prostate (7.3%), and stomach (5.6%) cancers. The essential foundations and methods behind the transition of normal healthy cells into dangerous cancer cells have been investigated during the preceding years. The survey obscured a vision of developing innovative and effective ways to combat this dangerous virus[3]. Currently, cumulative treatment approaches, such as surgery, immunotherapy, chemotherapy, radiation, and hormone therapy, are used, depending on the cancer stage. Almost every therapy has austere side effects on healthy tissues. It is costly for any affected individual to afford the therapies. Photodynamic therapy is an energetic multidisciplinary field that includes photobiology, photochemistry, and photomedicine for light production, filtering, and measurement technology. These all have light as a common theme. Photodynamic therapy is increasingly viewed as an inevitable modality for all cancer-related defects[4].

1.1 History Professor Hermann von Tappeiner, a pioneer of photobiology and head of the Pharmacological Institute at the Ludwig Maximilian University of Munich, coined ‘photodynamic action’ in 1904. In 1903–1905, the von Tappeiner group made its first attempts to use PDT to treat tumours and other skin diseases, such as lupus of the skin and condylomata of the female genitalia, by using various dyes, such as eosin, fluorescein, sodium dichloroanthracene disulphonate and ‘Grubler’s Magdalene red’. These dyes were mainly applied topically but also tried through intratumoral injections, and approving grades were testified. The advent of ionising radiation in cancer therapy probably led to its rigorous applications; hence,

Chapter 10

282

there was no long-term follow-up and PDT was soon completed and applied. Oscar Raab, a student of von Tappeiner, researched the harmful effects of acridine orange on Paramecium caudatum cells in one experiment. When incubated with acridine orange at a certain concentration, the cell lasted 1.5 hours and survived for roughly 15 hours under similar conditions in another experiment. A severe rainstorm was seen in one of the trials, prompting the alert student to ask if the light had a part, leading to the discovery of photodynamic activity. Oxygen was revealed to be essential for the photodynamic effect after a considerable amount of research on photosensitisation by the same researchers. Here, the acridine orange is the photosensitiser and the thunderstorm is the light source, which altogether leads to cell death [5]. In 1905, Jesionek and von Tappeiner successfully treated various dermatologic skin conditions, such as non-melanoma skin cancers, by using a photosensitiser of 5% eosin with an artificial light source. Further, few researchers postulated that the eosin could be incorporated into the cells like acridine orange [6]. When appropriate light is treated in the presence of oxygen, it leads to the cytotoxic reaction of the cell. These revolutionary experiments and subjects for human treatment paved a model for future curiosity and scientific awareness about photodynamic therapy [7].

1.2 Photodynamic therapy Photodynamic therapy treats patients with non-oncological (dermatology, gynaecology and urology) and oncological diagnoses with the promising outcome with minimal injuries. Photodynamic therapy is a form of light therapy involving light and a photosensitising chemical substance used in conjunction with molecular oxygen to elicit cell death. The three main processes through which PDT causes tumour demolition are as follows: 1) direct destruction of tumour cells, 2) damage to the tumour vasculature and construction of thrombus, followed by tumour infarction, and 3) stimulation of an immune response in contrast to the cancer cells [8–10]. 1.2.1 Photosensitisers Macrocyclic compounds (porphyrins, chlorins, bacteriochlorins, phthalocyanines, phthalocyanines, corroles, etc.) have the exceptional property of causing oxidative damage to cells. These compounds are evolving as the top diagnostic and healing agents for cancer. In 1950, the idea that hematoporphyrin is concentrated in tumour tissues accelerated research on novel porphyrin-based photosensitisers and their use in treating different cancers. The photosensitisers used in photodynamic therapy are

Advances in Nanomaterials for Enhanced Photodynamic Therapy

283

chemical compounds proficiently transformed to an excited state after light absorption. Most existing photosensitisers are aromatic and hydrophobic, and few are hydrophilic, which is not high enough for clinical use [11-12]. Concerning all these problems, the nanocarriers were incorporated into photosensitizers. The so-called ‘nanophotosensitisers’ can boost the solubility of photosensitisers, selectively treat the target tumour, and ensure delivery on the photosensitisers’ intracellular route. Several trials have been conducted to enhance photodynamic therapy efficacy. They have overcome the borders of nanophotosensitisers that display the reduced efficiency of reactive oxygen species generation because of the high local concentration of photosensitising molecules [13]. 1.2.2 Light Light, a significant driver in the growth of novel light sources, has remained the important driver of inexpensive, reliable diode lasers and light-emitting diodes, used as modus operandi for photodynamic therapy like broadband light from broadband light incandescent or arc lamps. It is also used as monochromatic light from tunable argon-pumped dye lasers or diode lasers. Lasers offer a narrow spectrum and high yield power in the vital spectral range. In contrast, the optical fibre permits the light, interstitially or endoscopically, to be delivered to any site in the human body. LEDs are non-coherent light sources that create wavelength bands wider than lasers. These light sources are usually around 25 nm and are scarcely used in photodynamic therapy to the light penetration depth. An ideal photosensitiser is triggered by light absorption at a 700–800 nm wavelength and offers light astuteness having depths of 5–6 mm[14–17].

2. Mechanism of action of PDT Photodynamic therapy is a dynamic process that is grounded on the photooxidation of living matter, linking three essential constituents: the photosensitiser, the radiation of light (with appropriate wavelength for maximum absorption of the substance) and oxygen. Photodynamic therapy uses non-ionising radiation and can be regulated repeatedly without causing long-standing complications. Photodynamic therapy kills cancer cells straight through apoptotic and non-apoptotic (necrosis, autophagy) trials and indirectly injures the tumour vasculature that supplies nutrients and oxygen to cancer cells[18-19]. The mechanism starts with absorption. The photon energy must correspond to the energy difference between the ground state of the absorbing molecule and one of its excited states (PS1 or PS2), resulting in the internal conversion of a molecule between electronic states

Chapter 10

284

having similar electronic spins. Consequently, the emission of excitation energy from a molecule in the form of light and the energy of the emitted photon correspond to the energy difference between the emitting and final states of the fluorescing mole. They both have the same electronic spin. There will be an intersystem movement of molecules between electronic states having different electronic spins, yielding a quantum of phosphorescence. Hence, two mechanisms mainly demonstrate the photodynamic effect (Figure 10-1).

2.1 Type I The photosensitiser, from its ambiences, transmits energy in its excited triplet state (T1) to the macromolecules. An electron is transmitted between the photosensitiser (T1 state) and the cancerous tissue (substrate), creating free radicals and anionic radicals in the photosensitiser and substrate. Electrons interact with oxygen molecules while remaining in their energetic elementary state. The initiated reactions lead to oxidative stress, creating reactive oxygen species and destroying the cancer cells[8], [19-20]. ܵ + ܴ‫ ܪ‬՜ ܵ‫ ܪ‬Ą + ܴĄ

ଷ ‫כ‬

ି

ܵ + ܴ‫ ܪ‬՜ ܵ Ą + ܴ‫ ܪ‬Ą

ଷ ‫כ‬



Where ܴ‫ ܪ‬Ą is the target substrate molecule, ܵ‫ ܪ‬Ą is the radical form of the ି photosensitiser, ܵ Ą is the radical anion form of photosensitiser, ܴĄ is the ା

radical derived from the target substrate molecule, and ܴ‫ܪ‬Ą is the cationic form of the target substrate molecule.

2.2 Type II The photosensitiser is present in its excited triplet form. It transfers energy straight to the oxygen molecule in the basic triplet state because they have the same spins. Then, the ground state molecular oxygen is elevated to the first excited singlet state: so-called singlet oxygen. Singlet oxygen acts rapidly with electrophile (B) molecules. It transfers energy to oxygen, bringing the photosensitiser to its fundamental state. This cycle repeats until a certain time [21]. ܵ + ଷܱଶ ՜ ଷܵ ‫ כ‬+ ଵܱଶ‫כ‬ ଵ ‫כ‬ ܱଶ + ‫ ܤ‬՜ ‫ܤ‬ை௑

ଷ ‫כ‬

Advances in Nanomaterials for Enhanced Photodynamic Therapy

285

Where ଷܱଶ is the triplet excited state of oxygen, ଵܱଶ‫ כ‬is the singlet oxygen molecule, and ‫ܤ‬ை௑ is the oxidised target substrate.

2.3 Cellular level Photodynamic therapy exerts a sturdy effect on cell division. Different sensitisers have different quantum yields of the photoinactivation of cells. Consequently, photodynamic therapy has a truncated mutagenic potential. The lifespan of the chief active photoproduct, ଵܱଶ is squat in cells, less than 0.05 μs and can draw out less than 0.02 μm from the production site. Hence, the mechanism of action is related to the intracellular localisation of the sensitisers in the nucleus and mitochondria. In contrast, most sensitisers are present outside the nucleus[6, 14].

2.4 Tissue level Canti et al. were the first to notice and demonstrate the immunological effects of PDT, acting through the ଵܱଶ pathway in vivo and in vitro. The low oxygen concentration in the region of many tumours may reduce the efficiency of PDT; therefore, PDT consumes oxygen itself by inducing reactions at influential low rates to increase their efficacy, leading to apoptosis (cell death mechanism)[6, 14].

2.5 Nanoparticles Nanotechnology is advantageous in PDT for three core reasons: 1. It reduces toxic effects on normal cells by aiming potential boosts photosensitisers concentration at the desired site. 2. It improves the solubility of hydrophobic photosensitisers; and 3. It uses zero-order release kinetics to ensure a constant rate of photosensitisers delivery at preferred sites.

286

Chapter 10

Photosensitisers are also conjugated in or restrained to nanoparticle stages via covalent or non-covalent interactions to yield a high surface-to-volume ratio. It is essential that oxygen species, which are the actual therapeutic agents, can diffuse in and out of the nanoparticle environment to use therapeutic efficacy by photosensitisation. Konan et al. classified the photosensitiser delivery process based on the presence or absence of a surface targeting molecule. First, concerning natural distribution patterns, liposomes, oil dispersions, biodegradable polymeric particles and hydrophilic polymer integrated with photosensitisers were termed ‘passive targeting system’ [21]. Second, strategies used to deliver the photosensitiser to the targeted tumour tissue using receptors or antigens were termed ‘active targeting system’ [12, 20].

3. Photosensitisers used in nanomedicine The nanocarriers aim to decrease the time to eliminate drugs and provide a defence mechanism from enzymatic or environmental degradation agents. The nanovesicles successfully fabricated with better affinity and firmness are recommended for targeted cancer cells. They also have magnetic reactivity or tissue-regeneration potential. The main disadvantages of photosensitisers are the short half-life inside the tissue and partial retention in the normal tissue. They are hard to produce and not stable. Hence, nanoparticles form the skeleton of photosensitising chemicals, transporting them more selectively to the tumour site with minimal toxicity and little harm to normal tissues with numerous roles to mitigate the disadvantages. Two groups reported the first study on nanophotosensitisers. Brasseur et al. used the photosensitiser ‘hematoporphyrin’ linked to an organic nanoparticle ‘polyalkylcyanoacrylate’ in 1991. The conjugation confirmed a simple discharge of photosensitisers from the formulation [22]. The objective is to investigate the use of nanoparticles coupled with photosensitisers. This will improve the productivity of photodynamic therapy conducted in vivo and in vitro against cancer [23, 24].

Advances in Nanomaterials for Enhanced Photodynamic Therapy

287

Figure 10-1: The Jablonski diagram explains Photodynamic Therapy’s photophysical and photochemical mechanisms. A: Light; B: Ground state singlet photosensitiser; C: Electronic transition; D: Excited state of the singlet photosensitiser; E: Triplet state photosensitiser; F: Intersystem crossing; G: Internal conversion; H: Fluorescence; I: Phosphorescence; J: Ground state triplet oxygen; K: Excited state of singlet oxygen.

3.1 Organic nanoparticle Organic nanoparticles comprise organic compounds (lipids, proteins, polysaccharides, or polymers) that have low toxicity and adaptability in transporting and promoting inactive phagocytic concentration or the measured release of miscellaneous range photosensitiser drugs. Passive biodegradable organic nanoparticles comprise the following elements: liposomes, oil dispersions, dendrimers, polymeric particles, and polymeric micelles. Moreover, passive non-biodegradable organic nanoparticles include ceramic-based, such as silica, alumina and titania [25].

Chapter 10

288

Table 10- 1: List of FDA-approved photosensitisers [25] Excitation wavelength

Approved

Porfimer sodium/Photofrin®

630 nm

Worldwide, withdrawn in EU for commercial reasons

5-ALA/Ameluz®/Levulan®

635 nm

Worldwide

Metvix®/Metvixia®

570–670 nm

Worldwide

temoporfin/mTHPC/Foscan®

652 nm

Europe

talaporfin/NPe6/Laserphyrin®

664 nm

Japan

verteporfin/Visudyne®

690 nm

Worldwide

Synthetic hypericin/SGX301

570–650 nm

Redaporfin®/LUZ11

749 nm

Photosensitisers

Orphan status in the EU Orphan status in the EU

Indication High-grade dysplasia in Barret’s Oesophagus. Obstructive oesophageal or lung cancer Mild to moderate actinic keratosis Nonhyperkeratotic actinic keratosis and basal cell carcinoma Advanced Head and neck cancer Early centrally located lung cancer Age-related macular degeneration Cutaneous T-cell lymphoma Biliary tract cancer

3.1.1 Liposomes Liposomes are among the most well-organised drug vehicles because of their modest archetypal structures, governable sizes and suitable preparation procedure for transmitting hydrophobic molecules in the aqueous medium. The central factor for using liposomal carriers is size management during aggregating with the photosensitiser. A higher amount of liposome accumulation in the tumour cells increases the availability of photosensitiser [25, 26].

Advances in Nanomaterials for Enhanced Photodynamic Therapy

289

3.1.2 Polymer nanoparticles Hydrophobic photosensitiser drugs are encapsulated at the core of the polymer nanoparticles because of their high stability, uniform particle size distribution and antifouling coating. This contributes to their inert targeting delivery via the enhanced permeability and retention (EPR) effect, increasing the blood circulation time. Polymers such as polyglycolide (PGA), polylactide (PLA), and their copolymer poly (D, L-lactide-coglycolide) (PLGA) have been predominantly used because of their versatility, physical sturdiness, biocompatibility, high drug-loading competence and controlled drug release. Non-biodegradable polymers are better than biodegradable polymeric carrier systems due to numerous advantages, such as facile synthesis, effortlessness of functionalisation, and structural integrity sturdiness[27–29]. 3.1.3 Silica nanoparticles Silica nanoparticles are chemically inert, resistant to pH changes, physically stable, and transparent to light. They can be attached to photosensitisers in physiologic conditions in the monomeric form to prevent self-aggregation. They can also be combined and accordingly affect the effectiveness of photoinactivation. The particle size needs to be carefully controlled as they are non-biodegradable. Organically modified silica (ORMOSIL) nanoparticles (ca. 30 nm) are monodispersed, spherical, steady in aqueous media and have fine pores of 0.50–1.00 nm in diameter to restrict drug release. They are also permeable toward molecular oxygen to generate singlet oxygen[27–30].

3.2 Inorganic nanoparticles Inorganic nanoparticles comprise an internal inorganic core of either a metal or a metallic oxide and an external organic shell that stabilises it in biological surroundings while preserving photosynthetic drugs. Adding biomolecules and inorganic nanoparticles were modified to eliminate the cytotoxic effects in the healthy cells caused by photosensitiser drugs. This further enhanced the photodynamic therapy targeting tumour cells as the drug actively absorbs specific receptors. Examples of energetic photosensitiser carrier nanoparticles include quantum dots, metal-based, up-converted, magnetic and carbon-based nanoparticles [31].

290

Chapter 10

3.2.1 Quantum dots Quantum dots made of semiconductors have several benefits for nanoparticles in photosensitisers, such as a wide absorbance cross-section and size-tunable optical characteristics, photostability, and metabolic resistance. The fluorescence resonance energy transfer (FRET) mechanism is processed. Further, quantum dots are used to sensitise the photodynamic therapy agent or follow the triplet energy transfer process to interact openly with molecular oxygen to yield singlet oxygen for inducing tumour cell death. Quantum dots exhibit a broad absorption spectrum, which provides flexibility to use adaptable excitation wavelengths to trigger the photosensitiser [13, 26]. 3.2.2 Up converter The up converter has near-infrared optical absorption coefficients. The photosensitiser can easily respond to certain wavelengths and terminate the cell [32]. 3.2.3 Metal The exact diameters of the particles must be obtained to minimise aggregation and diminish the biological efficiency of metal nanoparticles. Surface plasma resonance is localised when gold nanoparticles are exposed to light. This leads to an enhanced electromagnetic field and increased efficacy of photodynamic therapy when there is a conjugation between photosensitisers and gold nanoparticles[13, 26]. 3.2.4 Magnetic nanoparticles Magnetic nanoparticles are made from iron oxide or other superparamagnetic compounds containing the following properties: superparamagnetism, high field irreversibility, high saturation field, capacity to steer a magnetic field to particular places in the body and ability to create heat. Magnetite is the most promisingly used magnetic nanoparticle due to its biocompatibility [33]. 3.2.5 Carbon Good internalisation through endocytosis, fast elimination, no significant cytotoxicity and easy chemical design are properties required by a good nanoparticle for localising the photosensitisers at targeted cells. The most commonly used carbon nanomaterials are fullerene, carbon nanotube and

Advances in Nanomaterials for Enhanced Photodynamic Therapy

291

graphene. Fullerene (‫଺ܥ‬଴ ) has abundant pi bond electrons that act as efficient photosensitisers to produce singlet oxygen in photodynamic therapy. Carbon nanotubes can also act as photosensitisers or carriers for exogenous photosensitisers[26,29,30].

3.3 In vitro studies Photosensitisers and nanovesicles have been used to investigate the nanoparticle-mediated photodynamic therapy of cancer. Photosensitisers, light wavelength, and the molecular response of cells are three inseparable components of photodynamic therapy. We will describe several in vitro nanomedicine studies supported under varied photosensitisers. 3.3.1 Porphyrin Porphyrins are macrocyclic molecules with numerous physicochemical properties, such as anion binding, the steadiness of metals with uncommon oxidation states, electron transfer, and the development of irregular supramolecular assemblies that allow their use in photodynamic therapy. In contrast to their negative charge, they can produce singlet oxygen. For example, Chen et al. used human serum albumin nanoparticles for the distribution of 5,10,15,20-tetrakis(m-hydroxyphenyl) porphyrin (mTHPP), a porphyrin derivative and pheophorbides (chlorin derivatives), to leukaemia cells, leading to apoptosis that caused 50% of cell death. Porphyrin products such as photofrin and protoporphyrin IX have been expressed in quite a few different nanoparticle systems, such as metal oxide, chitosan, polymeric, silica and gold nanoparticles have planned the EPR mechanism to be operated to passively target solid tumours [13, 28, 30]. 3.3.2 Chlorin Chlorophyll derivatives, such as chlorins, discovered in the Spirulina species are commonly used in photodynamic treatment. These derivatives are chemically reduced porphyrins having higher-quality absorption capabilities that chelated metal ions may further alter. Chlorins are activated by light with a 650–700 nm wavelength, depending on the structure. The most prevalent drug is chlorin e6, studied in various nanovehicles, such as chitosan, silica, human serum albumin, hyaluronic acid, iron oxide, and innumerable polymeric nanoparticles. For example, Benachour et al. conjugated neuropilin-I–targeting peptide and silica nanoparticles loaded with 5-(4-carboxyphenyl)-10,15,20-triphenylchlorin (TPC). They did this to cause tumour angiogenic vessels to attain close to 100% of cell death.

292

Chapter 10

Recent research on the nanovehicle formulation of chlorins includes a sole platform called ‘upconversion nanoparticles’, in which wavelengths shorter than the excitation wavelength can emit light when there is sequential absorption of multiple photons. In turn, the photosensitiser will be activated within the same vehicle [13, 28, 30]. 3.3.3 Phthalocyanine The following properties, thermal stability, intense colour, high inertness, the ability to organise metal ions inside their core, and high absorption in the range of 670–700 nm, enable phthalocyanines to function as photosensitisers. For example, Ricci-Junior et al. studied the delivery of zinc (II) phthalocyanine to murine lymphoma cells via non-targeted PLGA nanoparticles to produce 60% of cell death. Qiao used similar lanthanidedoped upconversion nanoparticles but different drug concentrations, durations and irradiation power to produce up to 80% of cell death and fruitful imaging [13, 28, 30, 31].

3.4 In vivo studies Most in vitro research focused on nanoparticle formulation grounded in photodynamic therapy because many of the in vivo challenges concerning steady levels of drug delivery, site-selective delivery of light, and sufficient oxygen in tumours remain unresolved. 3.4.1 Porphyrin Porfimer sodium (Photofrin), a porphyrin derivative, is the first-generation photosensitiser studied in several murine models with varied nanoparticles. For example, Sun et al. used magnetic chitosan nanoparticles linked to 2,7,12,18-tetramethyl-3,8-di-(1-propoxyethyl)-13,17-bis-(3-hydroxypropyl) porphyrin (PHPP) to treat mice injected with human colon cancer xenografts. The magnetically induced targeting resulted in a major reduction in tumour size compared to non-targeted tumours [13]. 3.4.2 Chlorin Human serum albumin (HSA), glycol chitosan (HGC), iron oxide, hyaluronic acid (HA) and upconversion nanoparticles were weighed down for in vivo PDT experiments. Huang et al. established a solitary stand involving Ce6-conjugated iron oxide nanoparticles for magnetically directed drug delivery and subsequent photodynamic therapy. The results showed that the tumour size remained relatively stationary over 28 days following

Advances in Nanomaterials for Enhanced Photodynamic Therapy

293

nanoparticle administration and photodynamic therapy. Upconversion nanoparticles convert lower-level excitation photons to higher energy NIR photons for enhanced photodynamic therapy. Park et al. used Ce6-merged upconversion nanoparticles (UCNPs) to treat human glioma xenografts in mice, leading to tumour deterioration compared to controls [13, 32]. 3.4.3 Phthalocyanine Phthalocyanine photosensitisers have been investigated using numerous nanoparticles to obtain the desired result. For example, Nishiyama et al. weighed down phthalocyanine dendrimers inside polymeric micelles and treated human lung cancer xenografts in mice. Tumours that were treated with nanoparticles grew ten times faster than those that were not treated with nanoparticles after 30 days [13, 30].

4. Role of nanomedicine in PDT Nanomedicine is an emerging field in cancer diagnosis and therapeutics. The integration of nanotechnology with photodynamic therapy is one of the prospects of nanomedicine. Several in vitro and preclinical studies are underway in nanomedicine-based PDT. Karges et al. synthesised Ru(II) polypyridine complex that was linked to polymer (1,2-distearoyl-snglycero-3-phosphoethanolamine-N-[folate(polyethylene glycol)-2000] [ammonium salt]) (DSPE-PEG2000-folate). The polymer had lipophilic and hydrophilic portions that assisted in nanoparticle encapsulation. The Ru(II) polypyridine complex acted as a photosensitiser that absorbs laser light at 595 nm. Folate is used as a targeting ligand for the folate receptor, which is overexpressed in various tumour cells. The IC50 of the nanoparticle without irradiation in multicellular spheroids was about > 100 ȝM. In contrast, upon irradiation at 480 nm (10 min, 3.1 J/cm2) and 595 nm (60 min, 11.3 J/cm2), the IC50 values reduced drastically to 8.16 ± 0.87 ȝM and 9.62 ± 0.93 ȝ0 respectively. This result indicates the potential of nanoparticles to penetrate the multicellular spheroids and kill the cancer cells. This study shows the potential of incorporating nanoparticles in PDT [34]. However, in vivo studies using the same compound might highlight how this compound reacts to an animal system. Cheng et al. synthesised PEGylated AuNP– phthalocyanine 4 conjugate, where the AuNP enhances the drug’s delivery, and phthalocyanine 4 is a photosensitiser. Pc-4 requires 24–48 hours of injection as a control to reach the tumour site. However, when conjugated with AuNP, Pc-4 reaches the tumour site within two hours. This result proves that nanoparticles can improve PDT’s efficacy [35].

294

Chapter 10

5. Current clinical status of PDT Researchers worldwide are developing a photosensitiser that absorbs a consistent laser intensity, targets tumour cells precisely and is quickly excreted from the body to minimise side effects in patients [36]. However, not all molecules reach through the pipeline of clinical trials. Some of the studies which succeeded in completing the clinical trials have been mentioned (Table 10-2). However, nanoparticle-based PDT has not reached the phase of clinical trials. Most of the studies are still in the preclinical phase. Thus, some of the established photosensitisers that have completed clinical trials are detailed below. These data would provide information on photosensitisers' current status and provide insight into how nanotechnology can be linked with these photosensitisers to synthesise an improved novel PDT drug. Shafirstein et al. studied the efficacy of 5-aminolevulinic acid (ALA), a photosensitiser, in 29 patients with oral leukoplakia. Phase I of the trial determined the maximum tolerated dose (MTD) for ALA and radiant exposure duration for photodynamic daylight. In contrast, the second phase determined the lesion treatment efficacy and percentage regression. Phase I results demonstrated that the MTD was 8 J/cm2 for 1.5 hours at a pulse time of 1.5 ms. The efficacy of the therapy was evaluated after 90 days by percentage lesion regression. In total, 41% of the patients showed > 75% regression, and 53% showed > 50% regression. Although immunohistochemical studies showed that 73% of patients had an increased expression of p53 and 58% of patients had a decreased expression of Ki-67 [37]. p53 is a tumour-suppressor protein that disrupts tumour proliferation, enhances tumour cell apoptosis and protects normal cells from converting into malignant transformation [38]. Ki-67 is considered as a cell proliferation marker [39]. Szeimies and colleagues created the BF-200 gel nanoemulsion with ALA in a different study. The primary objective of this clinical trial was to compare the Aktilite® CL128 and PhotoDyn® 750 illuminations sources. The case study results demonstrated that the new formulation was effective and had no adverse effects. PhotoDyn® 750 (99%) showed a higher lesion clearance rate than Aktilite® CL128 (96%)[40]. After two years, the same group performed a comparison study with the BF-200 gel and another newly synthesised formulation with methyl-5-aminolevulinate (MAL). The study results depicted that BF-200 gel exhibited superior results to the MAL formulation. The complete clearance rate for BF-200 gel was 78.2%, whereas that for MAL formulation was 64.2% [41]. Table 10-2 shows detailed information on photosensitisers whose clinical trials were completed.

Stage of the trial

Phase II (Double masking)

Phase I (Openlabel)

Phase I and Phase II

Dosage

x HAL 5% with illumination x HAL 1% with illumination x HAL 0.2% with illumination x Placebo ointment without illumination

Three months postsurgery, a dose of pordimer sodium 2 mg/kg

Aminolevulinic acid orally at 0,1,2 hours before undergoing LED treatment

PDT Therapy

Hexaminolev ulinate (HAL)

Porfimer Sodium

Aminolevulin ic Acid (ALA)

Sr No

1.

2.

3.

405 nm

-

-

Laser intensity

30

10

262

No of pati ents

Oral cancer

Metastatic lung cancer

Cervical Intraepithelial Neoplasia

Type of cancer

Table 10-2: Comprehensive data of completed clinical trials in PDT

Results

[44]

[45]

x Patients (4-6%) had minor adverse reactions like discomfort during light treatment, mild headache, hypotensive post-treatment

[42], [43]

References

295

x After 13-18 days of treatment, tumour reduction and tumour necrosis were observed

x 95% improvement was observed when compared to the placebo group x Self-limiting local adverse reactions were in some women, including discharge, discomfort, and spotting

Advances in Nanomaterials for Enhanced Photodynamic Therapy

30

Phase I and II (Openlabel)

2mg/kg, 4 mg/kg or 6 mg/kg of IV administration

WST-11

6.

753 nm

45

2 mg/kg of Porfimer Sodium intravenously

Porfimer Sodium

180

5.

633 nm

Phase II (Openlabel)

Phase II

Aminolevulin ic Acid

4.

[47], [48]

[49], [50]

x 70% of stage I cancer had a complete response x 50% of T2 stage cancer had a complete response x 38% of T3/4 patients had a complete response

x 73.3% had a negative biopsy in the treated lobe x Minimal effects were observed on urinary and sexual function posttreatment

Head and Neck Cancer Precancerous/No n-malignant Condition

Localised prostate cancer

[46]

x No adverse reactions were observed

Nonmelanomatous Skin cancer

Chapter 10

Application of 20% ALA on superficial and nodular epidermally-derived lesions

296

5_ALA cream 20% topical administration

Topical applicationBF-200 ALA gel (78 mg/g)

Aminolevulin ic Acid

BF-200 ALA (5aminolevulini c acid) and methyl aminolevulin ate

BF-200 ALA gel

8.

9.

10.

I- 1-2 fields of approx 20 cm² BF-200 ALA gel II- Placebo gel III- BF-200 ALA gel with BF-RhodoLED

Methyl aminolevulinate (160 mg/g)

Intravenous dose of 2 mg/kg.

Porfimer Sodium

7.

Phase III (Triplemasking)

Phase III (Doublemasking)

635 nm

87

281

28

Phase I (Openlabel) Red light

10

Phase I

Actinic Keratosis

Non-aggressive basal cell carcinoma

Basal cell carcinoma

Metastatic lung cancer

x Response rate of BF200 ALA gel was measured as good 35.2%, good 24.1%, satisfactory 24.1%, nonsatisfactory 11.1% and impaired 5.6%

[53]

[52]

[51]

x 9/28 patients had a complete response x 50% of patients had a partial response x 5/28 patients had no response x Results of BF-200 ALA were better than methyl aminolevulinate

[44]

297

-

Advances in Nanomaterials for Enhanced Photodynamic Therapy

Phase I (Singleblinding)

Phase III (Openlabel)

I- ALA topical application, 1-hour incubation, 16 minutes 40 seconds (16:40) BLU-U exposure, application of sunscreen II- ALA topical application, 15 minutes incubation, 16:40 BLU-U exposure, application of sunscreen, 45 minutes daylight exposure III- ALA topical application, 15 minutes incubation, application of sunscreen, 1 hour daylight exposure

2-4 g of Methylamino levulinate and incubated for 90 minutes prior to red light

ALA

Methylamino levulinate

11.

12.

298

Red light

BLU-U blue light

20

24

Chapter 10

Actinic Keratosis

Actinic Keratosis

The total average number of lesions at day 0 was 197. After 12 weeks of treatment, 143 lesions showed complete response.

x PDT with BLU-U blue light and sunlight exposure group had 62.5% complete response while 25% had partial response x PDT with only BLU-U blue light had 37.5% complete response while 25% had partial response

[55]

[54]

Phase I and II (Doublemasking)

Phase III (Doublemasking)

I- 0.2% Hexyl amino levulinate mixed with Unguentum M cream II- 16% methyl amino levulinate

I- 20% ALA applied to upper extremities for 3 hours prior to 10 J/cm2 blue light II- Levulan Kerastick (vehicle solution) applied to upper extremities 3 hours prior to 10 J/cm2 blue light III- Only 10 J/cm2 blue light

Comparing 0.2% hexyl amino levulinate cream (HAL) and 16% methyl amino levulinate, MAL

Comparing ALA vs Levulan Kerastick

13.

14.

15.

Phase I (Openlabel)

Tazorac 0.1% gel followed by therapy with 20%aminolevulinic acid

Levulan (Tazorac) and ALA

Blue light

Daylight (2 hours)

Blue light

269

14

10

Actinic Keratosis

Actinic Keratosis

Actinic Keratosis

x 69.1% of the lesions were completely recovered from 20% of the ALA group. x 29.9% of the lesions were completely recovered from the Levulan Kerastick group.

x 73.4% of lesions showed complete response during HAL treatment x 77.8% of lesions showed complete response during MAL treatment

x No signs of mortality or severe adverse effects were observed

Advances in Nanomaterials for Enhanced Photodynamic Therapy

[59]

[57], [58]

[56]

299

Comparing ALA vs Levulan Kerastick

ALA- PDT therapy after cryotherapy

Metvix (comparing natural light vs conventional light)

16.

17.

18.

300

I- 20% ALA, broad area, 1-hour incubation II- 20% ALA, broad area, 2 hours incubation III- 20% ALA, broad area, 3 hours incubation IV- 20% ALA spot, 2 hours incubation V- Levulan Kerastick vehicle with 1-, 2-, 3hours incubations VI- Only blue light I- Cryotherapy followed by 3 amino levulinic acid II- Vehicle control 23 placebo topical application III- Cryotherapy followed by 2 aminolevulinic acid I- Metvix 160mg/g and natural daylight PDT II- Metvix 160mg/g and conventional PDT Phase III

Phase II (Doublemasking)

Phase II (Doublemasking)

100

166

BLU-U blue light

Natural light and Aktilite lamp

235

Blue light

Chapter 10

Mild actinic keratosis

Actinic Keratosis

Actinic Keratosis

[62]

[61]

x Patients showed high satisfaction in group I (50.9%), group II (44.1%) and group III (29.1%), respectively

x Group I had 89.2% lesions that made a complete recovery, while group II had a 92.8% rate.

[60]

x The first three study groups (23%, 24% and 25%) did not have a significant difference, inferring that the incubation time does not have much variation. In contrast, VI had 20% clearance after 12 weeks. x A 5% clearance rate was observed for the vehicle control.

5aminolevulini c Acid (5ALA)

Metvix

BF-200 ALA and MAL cream

BF-200 ALA

19.

20.

21.

22.

Phase III

Phase III (Doublemasking)

I- Placebo gel II- BF-200 ALA cream (78 mg/g)

Phase IIIb

Phase I and II

I- Placebo gel II- BF-200 ALA cream (78 mg/g) III- MAL cream (160 mg/g)

I-Metvix NDLnatural light-PDT II-Metvix conventional PDT III-Metvix placebo

5-Aminolevulinic Acid (Levulan Kerastick) (345 mg/1.5ml) + PDL585, ScleroPLUS laser

616

122

Aktilite CL128 and PhotoDy n 750

131

29

Broad spectru m light source

Natural daylight and conventi onal light

PDL585, ScleroP LUS laser (585 nm)

Actinic Keratosis

Actinic Keratosis

Actinic Keratosis

Leukoplakia and erythroplakia

x 41% of patients had > 75% regression (significant response) and 53% had > 25% regression (partial response) x The overall response rate of 94% at 90 days x Percent changes from baseline for metvixnatural daylight group was 68.4%, while conventional light was 71.5% x After 12 weeks of treatment, group I (3.9%), group II (48.4%) and group III (37.0%) showed complete response to the respective treatment x Complete clearance rate for group II (64%) and placebo group (11%) x Lesion complete clearance rate for group II (81%) and placebo (22%)

Advances in Nanomaterials for Enhanced Photodynamic Therapy

[40], [66]

[65]

[64]

[37], [63]

301

Metvix

ALA

23.

24.

302

I- ALA application on one extremity occluded II- ALA application on non-occluded III- Vehicle control on one extremity occluded II- Vehicle control on non-occluded

I-Metvix (160mg/g) NDL-natural lightPDT II-Metvix (160mg/g) blue-PDT

Phase II

Phase III (Openlabel)

Blue light

Natural daylight and blue light

70

26

Chapter 10

Actinic Keratosis

Actinic Keratosis

x Percentage of cleared lesions for each group was 88.7%, 70%, 16.7%, and 5.6%, respectively

x Percentage decrease in lesions after 1 month of treatment in group I were 89.6% and 94.6%

[68]

[67]

Advances in Nanomaterials for Enhanced Photodynamic Therapy

303

This information will enhance knowledge of the current photosensitisers and limitations of the photosensitisers to improve and synthesise a better, newer alternative photosensitiser.

6. Conclusion and future prospects Photodynamic therapy is one of the least invasive therapeutics against cancer. Thus, patients will be urged to adopt PDT to lessen the morbidity associated with other cancer therapies. However, the efficacy of PDT-based therapy is less than surgeries or chemotherapy in killing bulk tumour cells. Because the laser intensity of PDT may not be sufficient to kill large tumour masses because the photosensitisers can absorb the light intensity that reaches up to 5–10 mm in length [36]. This is a severe limitation of the currently available photosensitisers. Hence, PDT is usually used for superficial tumours, such as actinic keratosis. Thus, PDT is mainly used with chemotherapies, surgeries, immunotherapies, cancer vaccines, and the like. For instance, the nanocarriers of phenylboronic pinacol ester conjugated dextran with the emulsion of chemotherapeutic drugs of doxorubicin and photosensitiser chlorine [68]. Further, photosensitiser bremachlorin and long synthetic peptides (SLP) have epitopes of tumour antigens. Photosensitisers have also been used in combination with fluorescent dyes for in vivo imaging and to visualise the therapeutic effect of PDT. The future direction of PDT is synthesising new photosensitisers that are more effective in absorbing the laser intensity. Another possible future direction in PDT is synthesising the targeted delivery of photosensitisers at the site using small molecule drugs, antibodies or aptamers conjugated to the photosensitisers. Targeted therapeutics have also proven superior to conventional therapeutics because they can reduce the bystander effect and target tumour cells.

References [1] [2] [3]

S. Kwiatkowski et al., “Photodynamic therapy – mechanisms, photosensitizers and combinations,” Biomedicine & Pharmacotherapy, vol. 106, pp. 1098–1107, 2018, doi: 10.1016/j.biopha.2018.07.049. Naoyo Nishida et al., “Angiogenesis in cancer,” Vascular Health and Risk Management, vol. 2(3) pp. 213–219. 2006, doi: 10.2147/vhrm.2006.2.3.213 M. A. Calin et al., “Photodynamic therapy in oncology,” Journal of Optoelectronics and Advanced Materials, vol. 8, no. 3, pp. 1173– 1179, 2006, doi: 10.1634/theoncologist.11-9-1034.

304

[4] [5] [6] [7] [8] [9]

[10] [11] [12]

[13]

[14] [15]

[16]

Chapter 10

T. C. Zhu et al., “The role of photodynamic therapy (PDT) physics: PDT physics,” Medical Physics., vol. 35, no. 7Part1, pp. 3127–3136, Jun. 2008, doi: 10.1118/1.2937440. M.D. Daniell et al., “A history of PDT,” Aust. N.Z. J. Surg., vol.61, pp. 340-348, 1991, doi: 10.1111/j.1445-2197.1991.tb00230. D. P. Hader et al., “Comprehensive Series In Photochemical and Photobiological Sciences,” Surface Water Photochemistry, 2015, pp. P005-P006, doi: 10.1039/9781782622154-FP005 M.H. Gold, “A Historical Look at Photodynamic Therapy,” Aesthet Dermatol. Basel, Karger, 2016, vol 3, pp 1-7, doi: 10.1159/000439327 T. J. Dougherty, “Photodynamic therapy—new approaches,” Seminars in Surgical Oncology, vol. 5, no. 1, pp. 6–16, 1989, doi: 10.1002/ssu.2980050104. D. Andreadis et al., “Utility of photodynamic therapy for the management of oral potentially malignant disorders and oral cancer,” Translational Research in Oral Oncology, vol. 1, p. 2057178X1666916, Jan. 2016, doi: 10.1177/2057178X16669161. B. G. Maiya, “Photodynamic Therapy (PDT): 2. Old and new photosensitizers,” Reson, vol. 5, no. 6, pp. 15–29, Jun. 2000, doi: 10.1007/BF02833852. H. Abrahamse et al., “New photosensitizers for photodynamic therapy,” Biochemical Journal, vol. 473, pp. 347-364, doi: 10.1042/BJ20150942 C. K. Lim et al., “Nanophotosensitizers toward advanced photodynamic therapy of Cancer,” Cancer Letters, vol. 334, no. 2. Elsevier Ireland Ltd, pp. 176–187, Jul. 2013. doi: 10.1016/j.canlet.2012.09.012. A. Master et al., “Photodynamic nanomedicine in the treatment of solid tumors: Perspectives and challenges,” Journal of Controlled Release, vol. 168, no. 1. pp. 88–102, 2013. doi: 10.1016/j.jconrel.2013.02.020. U. Chilakamarthi et al., “Photodynamic Therapy: Past, Present and Future,” Chemical Record, vol. 17, no. 8. John Wiley and Sons Inc., pp. 775–802, 2017, doi: 10.1002/tcr.201600121. C. Mimikos et al., “Current state and future of photodynamic therapy for the treatment of head and neck squamous cell carcinoma,” World j. otorhinolaryngol.-head neck surg., vol. 2, no. 2, pp. 126–129, 2016, doi: 10.1016/j.wjorl.2016.05.011. V. K. Pustovalov et al., “Thermo-optical analysis and selection of the properties of absorbing nanoparticles for laser applications in

Advances in Nanomaterials for Enhanced Photodynamic Therapy

[17]

[18]

[19] [20] [21] [22] [23] [24] [25] [26]

[27]

305

cancer nanotechnology,” Cancer Nanotechnology, vol. 1, no. 1–6, pp. 35–46, Dec. 2010, doi: 10.1007/s12645-010-0005-1. M. Olek et al., “Photodynamic therapy for the treatment of oral squamous carcinoma—Clinical implications resulting from in vitro research,” Photodiagnosis and Photodynamic Therapy, vol. 27, pp. 255–267, 2019, doi: 10.1016/j.pdpdt.2019.06.012. J. M. Nauta et al., “Photodynamic therapy of oral cancer A review of basic mechanisms and clinical applications,” Eur J Oral Sci, vol. 104, no. 2, pp. 69–81, 1996, doi: 10.1111/j.1600-0722.1996.tb00049.x. G. M. F. Calixto et al., “Nanotechnology-based drug delivery systems for photodynamic therapy of cancer: A review,” Molecules, vol. 21, no. 3. MDPI AG, 2016. doi: 10.3390/molecules21030342. H. Abrahamse et al., “Nanoparticles for Advanced Photodynamic Therapy of Cancer,” Photomedicine and Laser Surgery. 35 (2017) 581–588. doi:10.1089/pho.2017.4308 W.-T. Li, “Nanotechology-based strategies to enhance the efficacy of photodynamic therapy for cancers,” Curr Drug Metab, vol. 10, no. 8, pp. 851–860, 2009, doi: 10.2174/138920009790274559. Channay Naidoo et al., “Photodynamic Therapy for Metastatic Melanoma Treat,” Technology in Cancer Research & Treatment, vol. 17, pp. 1-15, 2018, doi: 10.1177/1533033818791795 M. B. Oliveira et al., “Nanomaterials for Biomedical Applications,” Biotechnol. Journal, vol. 15, no. 12, p. 2000574, 2020, doi: 10.1002/biot.202000574. G.V. Roblero-Bartolón et al., “Use of nanoparticles (NP) in photodynamic therapy (PDT) against cancer,” Gaceta Medica de Mexico., vol. 151, pp. 85–98, 2015 D. van Straten et al., “Oncologic photodynamic therapy: Basic principles, current clinical status and future directions,” Cancers, vol. 9, no. 2. MDPI AG, p. 19, 2017, doi: 10.3390/cancers9020019. E. J. Hong et al., “Targeted and effective photodynamic therapy for cancer using functionalized nanomaterials,” Acta Pharmaceutica Sinica B, vol. 6, no. 4, pp. 297–307, 2016. doi: 10.1016/j.apsb.2016.01.007. M. Q. Mesquita et al., “An insight on the role of photosensitizer nanocarriers for photodynamic therapy,” Anais da Academia Brasileira de Ciencias, vol. 90, no. 1, pp. 1101–1130, 2018, doi: 10.1590/0001-3765201720170800.

306

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35]

[36] [37]

Chapter 10

K. Sztandera et al., “Nanocarriers in photodynamic therapy—in vitro and in vivo studies,” Wiley Interdisciplinary 5HYLHZV Nanomedicine and Nanobiotechnology, vol. 12, no. 3, 2020. doi: 10.1002/wnan.1599. H. Benachour et al., “Multifunctional peptide-conjugated hybrid silica nanoparticles for photodynamic therapy and MRI,” Theranostics, vol. 2, no. 9, pp. 889–904, 2012, doi: 10.7150/thno.4754. S. Wang et al., “Novel methods to incorporate photosensitizers into nanocarriers for cancer treatment by photodynamic therapy,” Lasers in Surgery and Medicine, vol. 43, no. 7, pp. 686–695, 2011, doi: 10.1002/lsm.21113. X. Chen et al., “Charge-reversal nanoparticles: novel targeted drug delivery carriers Acta Pharmaceutica Sinica B, 6 (2016) 268 Progress and perspectives on targeting nanoparticles for brain drug delivery,” 2016. D. B. Pemmaraju et al., “Photothermal therapy assisted bioactive nanoprobes for effective cancer theranostics,” 2019 WK E-Health and Bioengineering Conference, EHB 2019, pp. 6–9, 2019, doi: 10.1109/EHB47216.2019.8969989. S. Fathi Karkan et al., “Magnetic nanoparticles in cancer diagnosis and treatment: a review,” Artificial Cells, Nanomedicine, and Biotechnology, vol. 45, no. 1, pp. 1–5, Jan. 2017, doi: 10.3109/21691401.2016.1153483. J. Karges et al., “Polymeric Encapsulation of a Ru(II)-Based Photosensitizer for Folate-Targeted Photodynamic Therapy of Drug Resistant Cancers,” Journal of Medicinal Chemistry, vol. 64, p. 4622, 2021, doi: 10.1021/acs.jmedchem.0c02006. Y. Cheng et al., “Highly efficient drug delivery with gold nanoparticle vectors for in vivo photodynamic therapy of cancer,” Journal of the American Chemical Society, vol. 130, no. 32, pp. 10643–10647, 2008, doi: 10.1021/ja801631c. L. Benov, “Photodynamic therapy: Current status and future directions,” in Medical Principles and Practice, vol. 24, no. 1, pp. 14–28, 2015, doi: 10.1159/000362416. G. Shafirstein et al., “Using 5-aminolevulinic acid and pulsed dye laser for photodynamic treatment of oral leukoplakia,” Archives of Otolaryngology - Head and Neck Surgery, vol. 137, no. 11, pp. 1117–1123, 2011, doi: 10.1001/archoto.2011.178.

Advances in Nanomaterials for Enhanced Photodynamic Therapy

[38] [39] [40]

[41]

[42] [43]

[44] [45] [46] [47]

[48] [49]

307

F. Mantovani et al., “Mutant p53 as a guardian of the cancer cell,” Cell Death and Differentiation, vol. 26, no. 2. Nature Publishing Group, pp. 199–212, 2019. doi: 10.1038/s41418-018-0246-9. M. Sobecki et al., “The cell proliferation antigen Ki-67 organises heterochromatin,” eLife, vol. 5, no. MARCH2016, 2016, doi: 10.7554/eLife.13722. R. M. Szeimies et al., “Photodynamic therapy with BF-200 ALA for the treatment of actinic keratosis: Results of a prospective, randomized, double-blind, placebo-controlled phase III study,” British Journal of Dermatology, vol. 163, no. 2, pp. 386–394, 2010, doi: 10.1111/j.1365-2133.2010.09873.x. T. Dirschka et al., “Photodynamic therapy with BF-200 ALA for the treatment of actinic keratosis: Results of a multicentre, randomized, observer-blind phase III study in comparison with a registered methyl-5-aminolaevulinate cream and placebo,” British Journal of Dermatology, vol. 166, no. 1, pp. 137–146, 2012, doi: 10.1111/j.1365-2133.2011.10613.x. NCT01256424, “Dose-finding Study of Hexaminolevulinate (HAL) Photodynamic Therapy (PDT) to Treat Cervical Neoplasia - Full Text View - ClinicalTrials.gov.” P. Hillemanns et al., “A randomized study of hexaminolevulinate photodynamic therapy in patients with cervical intraepithelial neoplasia 1/2,” American Journal of Obstetrics and Gynecology, vol. 212, no. 4, p. 465.e1-465.e7, 2015, doi: 10.1016/j.ajog.2014.10.1107. NCT03344861, “Safety of PDT-Photofrin® Prior to Lung Surgery Full Text View - ClinicalTrials.gov.” NCT03638622, “Low-cost Enabling Technology for Image-guided Photodynamic Therapy (PDT) of Oral Cancer Cancer. - Full Text View - ClinicalTrials.gov.” NCT00002975, “Photodynamic Therapy in Treating Patients With Skin Cancer - Full Text View - ClinicalTrials.gov.” F. Civantos, “Photodynamic therapy for head and neck lesions in the subtropics,” JNCCN Journal of the National Comprehensive Cancer Network, vol. 10, no. SUPP.2. Harborside Press, p. S-65, 2012. doi: 10.6004/jnccn.2012.0179. NCT00453336, “Photodynamic Therapy With Porfimer Sodium in Treating Patients With Precancerous Lesions, Cancer, or Other Disease of the Aerodigestive Tract - Tabular View - ClinicalTrials.gov.” S. S. Taneja et al., “Final Results of a Phase I/II Multicenter Trial of WST11 Vascular Targeted Photodynamic Therapy for HemiAblation of the Prostate in Men with Unilateral Low Risk Prostate

308

[50] [51] [52] [53] [54] [55] [56] [57] [58]

[59] [60] [61] [62]

Chapter 10

Cancer Performed in the United States,” Journal of Urology, vol. 196, no. 4, pp. 1096–1104, 2016, doi: 10.1016/j.juro.2016.05.113. NCT00946881, “Safety and Tolerability Study Using WST11 In Patients With Localized Prostate Cancer - Full Text View ClinicalTrials.gov.” NCT00985829, “Evaluation of Efficacy of Photodynamic Therapy in Basal Cell Carcinoma - Full Text View - ClinicalTrials.gov.” NCT02144077, “Safety and Efficacy Study for the Treatment of Non-Aggressive Basal Cell Carcinoma With Photodynamic Therapy - Tabular View - ClinicalTrials.gov.” NCT01966120, “Safety and Efficacy Study for the Field-directed Treatment of Actinic Keratosis (AK) With Photodynamic Therapy (PDT) - Tabular View - ClinicalTrials.gov.” NCT03322293, “Daylight Photodynamic Therapy for Actinic Keratosis - Study Results - ClinicalTrials.gov.” NCT00926952, “Short Incubation Methylaminolevulinate Photodynamic Therapy Without Occlusion - Tabular View ClinicalTrials.gov.” NCT01053000, “Photodynamic Therapy With Levulan® +/- Topical Retinoid Pre-Treatment In The Treatment Of Actinic Keratoses Study Results - ClinicalTrials.gov.” NCT02149342, “Daylight-mediated Photodynamic Therapy of Actinic Keratoses:Comparing 0.2%HAL With 16%MAL - Study Results - ClinicalTrials.gov.” N. Neittaanmäki-Perttu et al., “Hexyl-5-aminolaevulinate 0·2% vs. methyl-5-aminolaevulinate 16% daylight photodynamic therapy for treatment of actinic keratoses: Results of a randomized doubleblinded pilot trial,” British Journal of Dermatology, vol. 174, no. 2. Blackwell Publishing Ltd, pp. 427–429, 2016. doi: 10.1111/bjd.13924. NCT02137785, “Safety and Efficacy Study of Photodynamic Therapy With Levulan Kerastick + Blue Light for Actinic Keratoses on the Upper Extremities - Tabular View - ClinicalTrials.gov.” NCT01475955, “Short-incubation Levulan Photodynamic Therapy Versus Vehicle for Face/Scalp Actinic Keratosis (AK) - Tabular View - ClinicalTrials.gov.” NCT02239679, “Controlled Study of the Occurrence of Actinic Keratosis on the Face After Cryotherapy + Aminolevulinic Acid (ALA) Photodynamic Therapy - Study Results - ClinicalTrials.gov.” NCT01475071, “Intra-individual Comparison of Efficacy and Safety of Metvix® Natural Daylight Photodynamic Therapy Versus

Advances in Nanomaterials for Enhanced Photodynamic Therapy

[63] [64] [65] [66] [67] [68]

309

Conventional Metvix® Photodynamic Therapy in Subject With Mild Actinic Keratoses - Study Results - ClinicalTrials.gov.” NCT00571974, “Treatment of Oral Premalignant Lesions With 5ALA PDT - Tabular View - ClinicalTrials.gov.” NCT01821391, “Phase 3b Study of Metvix NDL-PDT Versus Metvix c-PDT in Subjects With Actinic Keratoses - Tabular View ClinicalTrials.gov.” NCT02799069, “Evaluation of Safety and Efficacy of BF-200 ALA for the Treatment of Actinic Keratosis With Photodynamic Therapy - Study Results - ClinicalTrials.gov.” NCT02799082, “Evaluation of Efficacy and Safety of BF-200 ALA Used With Photodynamic Therapy in Patients With Actinic Keratosis. - Study Results - ClinicalTrials.gov.” NCT02373371, “Actinic Keratoses Treatment With Metvix® in Combination With Light - Study Results - ClinicalTrials.gov.” NCT01458587, “Levulan PDT Versus Vehicle for Extremity Actinic Keratoses (AK) - Study Results - ClinicalTrials.gov.”

CHAPTER 11 NANOPARTICLE-BASED DRUG DELIVERY FOR BONE DISORDERS SHEBA RANI NAKKA DAVID,1* SANJOY KUMAR DAS2 AND SOUMALYA CHAKRABORTY3 1

School of Pharmacy, University of Wyoming, Laramie, Wyoming, USA Institute of Pharmacy Jalpaiguri, Jalpaiguri, West Bengal, India 3 National Institute of Pharmaceutical Education and Research (NIPER), S.A.S. Nagar, Punjab, India 2

*Corresponding author: Dr Sheba Rani David Assistant Professor School of Pharmacy University of Wyoming 1000 E. University Avenue Laramie, Wyoming, 82071 United States of America Email: [email protected] Phone: +1- 307-766-6482

Abstract Bones are one of the body’s primary organs and play a vital role in everyday bodily functions. Therefore, bone disorders can cause severe life-long complications and significantly affect an individual’s mortality and morbidity rates. However, determining effective treatments for these bone disorders is still a challenge. Hence, one possible approach is to use nanoparticles as vehicles to deliver the drugs more effectively and to their targeted site of action.

Nanoparticle-Based Drug Delivery for Bone Disorders

311

Abbreviations ALN ALN-NDs ALN-PEG2k-ALN ALP APA BMP2 CPNPs DNA DOX HA MSNs NDs PLGA siRNA SV TNT ZOL

Alendronate Alendronate-conjugated nanodiamonds ALN-conjugated poly (Ethylene Glycol) 2000- ALN Alkaline phosphatase ALN-PEG2k-ALN Bone morphogenetic protein 2 Calcium phosphate nanoparticles Deoxyribonucleic acid Doxorubicin Hydroxyapatite Mesoporous silica nanoparticles Nanodiamonds Poly(Lactic-Co-Glycolic Acid) Small interfering ribonucleic acid Simvastatin Titanium nanotube Zoledronic acid

1. Introduction Bones are calcified tissues consisting of viable cells embedded in a mineralised organic matrix. The normal bone matrix usually comprises 50– 70% of inorganic material, around 20–40% of the organic matrix, 5–10% of water and less than 3% of the fat tissue. The bones’ critical functions are protecting vital organs, providing mechanical support for body movement, and regulating mineral homeostasis. These functions depend on healthy bones maintained by the constant remodelling of the bone when the mature bone is removed and replaced with the new bone. This remodelling process is mediated by the bone-resorbing cells, osteoclasts, bone-forming cells and osteoblasts [1]. Effective treatments for common bone disorders, such as bone metastases, osteoarthritis, osteoporosis and osteomyelitis, are not available yet. Therefore, targeted treatment delivery using nanoparticles may help increase the efficacy of current treatments. Osteosarcoma, Chondrosarcoma and Ewing’s sarcoma are the three kinds of bone cancer that can affect any bone. Although bone metastases are under cancer metastases, in this case, the primary tumour spreads and invades the bones. The bone is one of the most common sites affected by metastatic

312

Chapter 11

cancer, meaning that treating bone metastases can significantly improve the morbidity of the cancers. Osteoarthritis, also known as degenerative joint disease, is the most common form of arthritis. It occurs when the cartilage between the joints wears down, which can cause swelling of the joints, pain and stiffness. Nanoparticles can be used for osteoarthritis drugs concerning local drug delivery systems because nanoparticles can potentially increase the drug retention time in the body fluids. Osteoporosis is when the bone weakens and becomes brittle and fragile because of the low bone mass and degeneration of bone tissue. Using nanodiamonds, gold nanoparticles, and calcium phosphate nanoparticles assists in osteoporosis treatment. Indeed, these nanoparticles can stimulate new bone growth and mineralisation of the bone and promote bone cell activity. Finally, osteomyelitis is an infection of the bone and bone marrow. The bone becomes infected because of local bacteria exposure (during surgery or open fracture), or the infection has spread throughout the body, including the bone and bone marrow. Osteomyelitis can be classified into three types: acute, subacute and chronic. Antibiotics, chemotherapeutics and gene therapy, such as plasmid deoxyribonucleic acid (DNA) or small interfering ribonucleic acid (siRNA), are commonly used in the treatment of bone problems (siRNA). Some limitations of the current conventional treatments for bone disorders are low targeting efficiency, adverse effects on other organs and tissues, short plasma half-life, and poor bioavailability. In medicine, nanoparticles can be used as vehicles to deliver the therapeutic agents to their site of action to achieve a productive and safe treatment. Additionally, some nanoparticles, such as nanodiamonds, gold nanoparticles and calcium phosphate nanoparticles, have unique characteristics that can stimulate new bone growth by promoting bone cell activity. One class of therapeutic agents of bone metastases is bisphosphonates because they have antitumor effects and can be used to treat bone metastases [2–3]. Bisphosphonates have been well established in treating other bone diseases because of their specific affinity to bone tissues. Moreover, bisphosphonates can strengthen the bone and help treat and prevent osteoporosis [4]. With their specific affinity to bone tissues, bisphosphonates and nanoparticles can be combined to increase the efficacy of bisphosphonates and deliver them more effectively.

Nanoparticle-Based Drug Delivery for Bone Disorders

313

2. Bone-targeting drugs and targeted drug delivery for bone The concept of targeted drug delivery to the bone for the treatment of bone disorders can be a potential strategy to address the issue of treatment efficacy because the body’s visceral organs will mostly absorb non-targeted drugs. Bone-targeting drugs will allow the drugs to be consistently delivered in the therapeutic range, thus ensuring the high efficacy of the drugs.

3. Theory for nanoparticle-based drug delivery Nanoparticles offer unique characteristics for potential targeted drug delivery for bone disorders, especially in improving drugs’ efficacy. Some advantages of using nanoparticles in the drug delivery systems are as follows: 1. they protect the drugs from an early degradation by the body fluids and increase the drug’s retention time in the body; 2. they can carry the drug to its specific target while keeping the drug’s concentrations in the therapeutic range to deliver maximum effects; 3. they can carry more drug molecules and increase the stability of hydrophobic drugs due to the large surface area of nanoparticles; and 4. specific delivery can be achieved with loading targeting molecules by modifying the nanoparticles’ surfaces [5].

4. Types of nanoparticles Nanoparticles used for drug delivery for treating bone disorders can be classified into two different categories: organic and inorganic nanoparticles. These broad classifications are based on their material of origin. Both organic and inorganic nanoparticles have a few different types of nanoparticles that are used for treating bone disorders. Nanoparticles are used for therapeutic applications and diagnostics by cell labelling. However, this chapter will discuss the therapeutic uses of nanoparticles.

4.1 Organic nanoparticles Organic nanoparticles, such as polymeric nanoparticles, have a great potential to be used in drug delivery. Moreover, organic-based nanoparticles are environmentally friendly due to their inherent nature. A few examples

314

Chapter 11

of organic nanoparticles used for the drug delivery of bone treatment are poly (L-lactide-co-glycolide), chitosan and lipid nanoparticles [5].

4.2 Poly(L-lactide-co-glycolide) nanoparticles Poly (L-lactide-co-glycolide) (PLGA) nanoparticles can be used in bone drug delivery. They show low toxicity, tuneable drug release kinetics and ease of functionalisation with target molecules. Additionally, PLGA nanoparticles can deliver hormones and proteins, specifically for bones and parathyroid hormones. The delivery is done by protecting the hormone from enzymatic degradation and improving the diffusion of the hormone into circulation. Alendronate, an anti-resorptive medication, is another medicine in which PLGA nanoparticles exhibit a therapeutic effect. The PLGA nanoparticles can reduce the systemic adverse effects of the drug by regulated delivery in small doses. Wang et al. developed simvastatin-loaded PLGA nanoparticles with tetracycline (TC)-based bone-targeting moieties to treat osteoporosis. An in vitro release study revealed an 80% cumulative release of simvastatin from PLGA nanoparticles in PBS at 72 hours. An in vitro cell evaluation indicated that the nanoparticles had an excellent cellular uptake capacity and showed great biocompatibility with MC3T3E1 cells, thereby reducing the cytotoxic effects of SIM. An enhanced bone cell targeting efficiency was observed by incorporating tetracycline moiety compared to PLGA nanoparticles alone. In vivo study showed an enhanced curative effect of simvastatin from PLGA nanoparticles with tetracycline targeting moiety compared to the nanoparticle without targeting moiety or pure simvastatin [6]. Pignatello et al. developed a novel biomaterial by conjugating PLGA and alendronate (an amino-bisphosphonate) to fabricate bone-targeted nanoparticles. Due to bisphosphonate residue, a greater extent of absorption in hydroxyapatite was observed for alendronate-conjugated PLGA compared to the PLGA alone. A favourable cytotoxic profile was further confirmed from the lack of hemolysis or absence of cytotoxic effects on endothelial cells or trabecular osteoblasts [7].

4.3 Chitosan nanoparticles Chitosan nanoparticles are derived from the de-acetylated form of chitin and are a naturally occurring polysaccharide [8]. Chitosan nanoparticles are used as gene carriers because they effectively protect the gene from nuclease degradation. The positively charged chitosan nanoparticles can emit strong

Nanoparticle-Based Drug Delivery for Bone Disorders

315

electrostatic interactions with negatively charged nucleotides [9]. However, further modification of the chitosan nanoparticles needs to be done, as chitosan nanoparticles show low cell internalisation efficiency.

4.4 Lipid nanoparticles Lipid nanoparticles comprise various lipid-based nanocarriers, such as solid lipid nanoparticles, lipid nanocapsule carriers, lipid drug conjugate carriers and nanostructured lipid carriers. Lipid nanoparticles can be used for bone drug delivery because of their high biocompatibility and kinetic and physical stability. Lipid nanoparticles can deliver nucleic acids, such as siRNA, for gene therapy. Their small size allows them to deliver several nucleic acids into the cells by endocytosis. siRNA is used to decrease the synthesis of sclerostin, a protein secreted by osteocytes. This protein inhibits osteoblast differentiation. Liang et al. developed CH6 aptamer-functionalised lipid nanoparticles encapsulating osteogenic pleckstrin homology domain-containing family O member 1 (Plekho1) siRNA (CH6-LNPs-siRNA). CH6 facilitated in vitro osteoblast-selective uptake of Plekho1 siRNA, mainly via macropinocytosis, and boosted in vivo osteoblast-specific Plekho1 gene silencing. These actions promoted bone formation, improved bone microarchitecture, increased bone mass and enhanced mechanical properties in osteopenic and healthy rodents. The results present a promising opportunity for aptamer-functionalised lipid nanoparticles as a carrier for siRNA-based bone anabolic strategy [10]. Wang et al. prepared a conjugate of distearoyl phosphoethanolamine polyethylene glycol with 2(3-mercaptopropylsulfanyl)-ethyl-1,1-bisphosphonic acid and grafted it into the doxorubicin-loaded liposomes and micelles for targeting bones. The affinity of the micellar and liposomal formulations to hydroxyapatite (HA) was assessed in vitro. The results indicated that all the thiolBP-incorporated nanocarriers had a stronger HA affinity than their counterparts without thiolBP. The thiolBP-decorated liposomes also displayed a strong binding for a collagen/HA composite scaffold in vitro. More importantly, thiolBPdecorated liposomes increased retention in the collagen/HA scaffolds after subcutaneous implantation in rats. The designed liposomes were able to entrap the bone morphogenetic protein-2 in a bioactive form, indicating that the proposed nanocarriers could deliver bioactive factors locally in mineralised scaffolds for bone tissue engineering [11]

316

Chapter 11

4.5 Inorganic nanoparticles The human bone comprises inorganic materials (50–70%). Hydroxyapatite is one major inorganic substance, such as calcium phosphate, calcium carbonate, calcium fluoride, calcium hydroxide, and citrate. Inorganic nanoparticles have similar mechanical structures to bone tissues; hence, bone drug delivery using inorganic nanoparticles can improve the delivery of the drugs, thus enhancing the drug efficacy. Moreover, inorganic nanoparticles are considered safe and biodegradable with the possibility of undergoing surface modification, making them a better promising field than viral vectors for targeted nanodelivery. The other desirable features are reduced systemic toxicity, sustained therapeutic efficacy, lesser side effects and easy spatial configuration in biomedical implants. These features contrast pure proteins, which give inorganic nanoparticles the leverage to be a promising future strategy. Some inorganic nanoparticles used for bone drug delivery are mesoporous silica nanoparticles, titanium nanotubes, gold nanoparticles, nanodiamonds and calcium phosphate nanoparticles [5].

4.6 Mesoporous silica nanoparticles Mesoporous silica nanoparticles are a mesoporous form of silica. They can deliver drugs to the bone because they have a large surface that allows efficient drug loading [12]. It is particularly useful in treating osteoporosis because mesoporous silica nanoparticles can simultaneously act as a drug vehicle and bone bioactive agent [13]. Mesoporous silica nanoparticles can either enhance osteoblast activity, promote new bones, or decrease bone resorption by reducing osteoclast activity. Martinez-Carmona et al. created a novel multifunctional nanodevice based on doxorubicin (DOX)-loaded mesoporous silica nanoparticles (MSNs) for bone cancer treatment. The MSN-based nanodevice exhibited a significantly higher internalisation degree in human osteosarcoma cells. Moreover, DOX loading at lower concentrations led to almost 100% osteosarcoma cell death than healthy bone cells, significantly preserving their viability. Besides this result, this nanodevice exhibits a cytotoxicity effect on tumour cells eightfold higher than that caused by the free drug [14]. Sun et al. fabricated MSNs anchored by zoledronic acid (ZOL) to target bone sites and delivered the antitumour drug DOX spatiotemporally controlled manner. MSNs-ZOL had better bone-targeting ability compared with non-targeted MSNs. Zol-anchored DOX-loaded MSNs showed a pH-

Nanoparticle-Based Drug Delivery for Bone Disorders

317

sensitive DOX release behaviour. The cellular trafficking study indicated that Zol-anchored DOX-loaded MSNs entered cells through the ATPdependent pathway. Then, they localised in lysosomes to achieve effective intracellular DOX release. The antitumour results indicated that Zolanchored DOX-loaded MSNs exhibited the best cytotoxicity against A549 cells. They significantly decreased cell migration in vitro [15].

4.7 Titanium nanotubes Titanium metals are commonly used in bone implants in the form of pins, plates and screws based on their resistance to corrosion, mechanical strength and biocompatibility [16]. The use of titanium nanotubes can potentially further increase the integration of the bone implant to the surrounding tissue and, hence, enhance osseointegration [17]. Additionally, the nanostructure of titanium nanotubes can potentially improve new bone formation by increasing cell adhesion, proliferation and osteoblast differentiation. Alendronate, an anti-resorptive drug, has stimulated osteoblast proliferation and differentiation and inhibited osteoclast cells. However, there is an issue with the delivery of this drug as it can cause toxic effects at high doses. Thus, a solution has been proposed to deliver alendronate using titanium nanotubes, promoting osseointegration and stimulating the formation of new bone cells. This solution can be used as a potential treatment for osteoporosis. Wei et al. applied bone morphogenetic protein 2 (BMP2)/macrophagederived exosomes to improve the bio-functionality of titanium nanotube implants to favour osteogenesis. Incorporating BMP2/macrophage-derived exosomes dramatically improved the expression of early osteoblastic differentiation markers, alkaline phosphatase (ALP) and BMP2. It indicates the pro-osteogenic role of the titanium nanotubes incorporated with BMP2/macrophage-derived exosomes. The titanium nanotubes functionalised with BMP2/macrophage-derived exosomes activated autophagy during osteogenic differentiation [18]. Liu et al. developed a new dual-controlled, local, bone-targeting delivery system by loading the tetracycline-grafted simvastatin (SV)-loaded polymeric micelles in titanium nanotube (TNT). Ti surfaces with TNT–bone-targeting micelles could promote cytoskeletal spreading, early adhesion, alkaline phosphatase activity and extracellular osteocalcin concentrations of rat osteoblasts, with concomitant-enhanced protein expression of BMP2 [19].

318

Chapter 11

4.8 Gold nanoparticles Gold nanoparticles may be used in bone drug delivery because they can enhance osteoblast activity and decrease osteoclast differentiation by inhibiting the osteoclastogenesis promoter, RANK ligand, and reducing the levels of reactive oxygen species. Additionally, gold nanoparticles could also induce osteogenic differentiation by activating the mechanosensitive p38 mitogen-activated protein kinase pathway. This is done by applying mechanical stress on the membranes of mesenchymal stem cells.

4.9 Nanodiamonds Nanodiamonds can be used in bone drug delivery because of their exemplary biocompatibility, ease of surface functionalisation enabling protein or polymer linkages, and high surface area. Similarly, like gold nanoparticles, nanodiamonds can upregulate osteoblast proliferation and differentiation and downregulate osteoclast activity [20]. Ryu et al. designed nanodiamonds (NDs) conjugated with alendronate (ALN) for bone-targeted delivery as ALN-NDs, which exhibited a high affinity to hydroxyapatite (HA, the mineral component of bone) because of the presence of ALN. ALN-NDs exhibited a preferential cellular uptake by MC3T3-E1 osteoblast-like cells compared to NIH3T3 and HepG2 cells, suggesting their cellular specificity. The in vivo study revealed that ALNNDs effectively accumulated in bone tissues after intravenous injection. These results confirm the superior properties of ALN-NDs with advantages of high HA affinity, specific uptake for MC3T3-E1 cells, positive synergistic effect for ALP activity and in vivo bone-targeting ability [21].

4.10 Calcium phosphate nanoparticles Calcium phosphate nanoparticles have a similar structure to the inorganic materials of bone minerals. It also has high biocompatibility and biodegradability. This result shows that calcium phosphate nanoparticles can be used for bone drug delivery [22]. Like chitosan nanoparticles, calcium phosphate nanoparticles can deliver nucleic acids. They can help stabilise the electrostatic interactions because they have positive calcium ions to neutralise the negatively charged nucleotides. Chu et al. fabricated double ALN-conjugated poly (ethylene glycol) 2000 (ALN-PEG2k-ALN), which were termed as APA that were further formulated with modified calcium phosphate nanoparticles (CPNPs). These were called APA-CPNPs, which have an ALN targeting moiety and hydrophilic poly (ethylene glycol)

Nanoparticle-Based Drug Delivery for Bone Disorders

319

arms tiled on the surface for bone-targeted drug delivery. The superior bonebinding ability of APA-CPNPs was verified via the ex vivo imaging of bone fragments. An in vitro release experiment demonstrated that APA-CPNPs can release drugs faster in an acid environment than in a neutral environment. Cell viability experiments indicated that blank APA-3 CPNPs possessed excellent biocompatibility with normal cells [23].

5. Strategies for drug delivery Nanoparticles can potentially enhance the current delivery systems of the drugs. Some drug delivery methods to the bone tissue are oral delivery and transdermal delivery. Figure 5.1 presents a few different nanodrug delivery strategies used in rheumatoid arthritis.

6. Intravenous delivery Although oral drug delivery would be convenient and improve patient compliance, some drugs are degraded by enzymes, rendering them therapeutically ineffective. Therefore, enzymatically destroyed medications must be administered straight to the bloodstream for quicker and more efficient biological activity. Intravenous drug delivery avoids the first-pass metabolism and enhances the bioavailability of the active therapeutic ingredient. Meka et al. formulated dexamethasone (DEX) liposomes coated with a novel peptide ligand known as ART-2 to treat arthritis through targeted delivery. These were intravenously injected into rats with collageninduced arthritis (CIA) and were accumulated in the inflamed area. ART-2 coated dexamethasone liposomes demonstrated more significance than both free DEX and liposomes without ART-2 coating in inhibiting arthritis progression [24]. The effectiveness of targeting the inflamed joints was analysed by Heo et al. for methotrexate-loaded dextran sulphate nanoparticles administered intravenously to CIA mice tail. The results ascertained the better targeting of nanoparticles aggregated 12 times more in the inflamed joints than free methotrexate, significantly enhancing the therapeutic efficacy of methotrexate [25]. As mentioned under the nanodiamonds section of this chapter, Ryu et al. prepared ALN-NDs, which displayed a higher targeting efficiency with improved deposition in bone tissue through intravenous injection [21].

320

Chapter 11

Figure 5.1. Nanodrug delivery strategies used in rheumatoid arthritis. Few potential routes of administration are (A) intravenous drug delivery, (B) transdermal drug delivery, (C) oral drug delivery, and (D) intraarticular drug delivery. Created with BioRender.com

Nanoparticle-Based Drug Delivery for Bone Disorders

321

7. Transdermal delivery Nanoparticles can enhance the penetration of the drugs into the skin more effectively due to their physical stability, drug release potential and ability to penetrate the skin more efficiently. This is also the case for nanostructured lipid carriers [26]. One drug that can be possibly used with nanostructured lipid carriers is methotrexate, which can be used for rheumatoid arthritis. Methotrexate is used to reduce the pro-inflammatory cytokines. However, the intraarticular delivery of the drug yields low efficacy because there is a low concentration of the drug at the site of rheumatoid arthritis due to the fast excretion of the drug from the joints. Thus, nanostructured lipid carriers are used to address this issue by formulating a gel that incorporates nanostructured lipid carriers and methotrexate with a chemical enhancer to increase permeability [27]. Nonetheless, this application of nanostructured lipid carriers to methotrexate has been limited to animal studies. The formulated gel could reduce the pro-inflammatory cytokines and hence reduce inflammation.

8. Oral delivery Oral delivery can be the preferred delivery method because it is noninvasive and convenient to the patient. However, some challenges that can arise from the oral delivery are lower drug concentration before entering circulation due for enzymatic degradation in the gastrointestinal tract and poor permeation across the mucosal barrier due to the molecule size of the drugs. Nanoparticles can be utilised in oral delivery to address these issues because they can cross the intestinal epithelium. One example shows the use of ȕ-casein nanosized micelle loaded with celecoxib can reduce the toxicity issues of celecoxib and enhance its dispersibility. This approach can treat osteoarthritis and rheumatoid arthritis [28].

9. Intraarticular delivery Diarthrodial joints are an ideal location for intraarticular drug delivery for local delivery. These joints leverage several benefits to an array of arthropathies. The intraarticular route of administration has its challenge of fast egressing the administered ingredients from the joint space, which is prominent for particles lesser than 100 nm. Nevertheless, 100 nm–5-μmsized particles satisfactorily achieved better drug retention [29]. Despite this shortcoming, this delivery route has many benefits, such as minimal systemic exposure, lesser adverse effects, enhanced bioavailability, and

322

Chapter 11

substantially lesser drug cost. Thus, intraarticular delivery has become a profitable research avenue. Currently, hyaluronic acid formulations and corticosteroids are the major constituents of FDA-approved intraarticular preparations. Nonetheless, various approaches, such as gene therapy, cell therapy, liposomes and microparticles, are also being researched in ongoing preclinical trials. Li et al. developed nanoparticles from silica, a non-metallic inorganic ingredient extensively used in rheumatoid arthritis therapeutics. MSN synthesised as a core-cone structure was administered via intraarticular injection. Hyaluronan synthase type 2 (HAS2) was delivered using functionalised polyethyleneimine on the surface of the moiety. MSNs successfully enhanced the production of endogenous hyaluronates, substantially improving synovial joint cartilage protection. The results demonstrated that intraarticular administration was capable of better therapeutic efficacy with a longer duration and was convenient and efficient compared to HA injection [30]. Headland et al. fabricated a complex nanoparticle that combined the antiinflammatory membrane junction protein A1 coupled with the neutrophilderived microvesicles, abundantly present in synovial joint fluid. The introduction of the microvesicle complex through intraarticular injection demonstrated improved extracellular matrix accumulation and protection of the cartilage due to cartilage degradation reduction and enhanced chondrocyte functionality [31].

10. Nano toxicity Because of the enhanced absorption of the nanoparticles, the toxicity of nanoparticles may pose a challenge because nanoparticles can cause physical and chemical damage to healthy cells. Additionally, nanoparticles may also cause cell dysfunction as the nanoparticles can bind to the biological molecules in abnormal or unexpected amounts. Further, the random membrane insertion of the nanoparticles may induce cell death and physically block microcirculation.

11. Conclusion and future directions Although some of the nanoparticles can be used to deliver the drugs more effectively, the safety and toxicity issues of the nanoparticles themselves must be further studied. Further research on the use of nanoparticles in the

Nanoparticle-Based Drug Delivery for Bone Disorders

323

delivery of drugs for bone disorders should observe the efficacy over the risks of nanotoxicity because some of the nanoparticles are only studied in animal studies. Moreover, studies regarding the interactions between nanoparticles and normal biological cells and tissues should be conducted because nanoparticles can interfere with normal biological cells' functions. With the improvement in technology, the use of nanoparticles in local delivery and targeted specific delivery to bone tissues and cells may play a significant role in treating bone disorders. The current conventional treatment for bone disorders is expected to improve soon without nanoparticles.

References 1. 2. 3. 4. 5.

6.

7.

8.

X. Feng, Chemical and biochemical basis of cell-bone matrix interaction in health and disease. Current Chemical Biology, 3 (2009) 189–196. https://doi.org/10.2174/187231309788166398. R. Coleman, The use of bisphosphonates in cancer treatment. Annals of the New York Academy of Sciences, 1218 (2011) 3–14. https://doi.org/10.1111/j.1749-6632.2010.05766.x. P. Clézardin, I. Benzaïd, & P. I. Croucher, Bisphosphonates in preclinical bone oncology. Bone, 49 (2011) 66–70. https://doi.org/10.1016/j.bone.2010.11.017. R. E. Coleman & E. V. McCloskey, Bisphosphonates in oncology. Bone, 49 (2011) 71–76. https://doi.org/10.1016/j.bone.2011.02.003. H. Cheng, A. Chawla, Y. Yang, Y. Li, J. Zhang, H. L. Jang, & A. Khademhosseini, Development of nanomaterials for bone-targeted drug delivery. Drug Discovery Today, 22 (2017) 1336–1350. https://doi.org/10.1016/j.drudis.2017.04.021. H. Wang, J. Liu, S. Tao, G. Chai, J. Wang, F. Hu, & H. Yuan, Tetracycline-grafted PLGA nanoparticles as bone-targeting drug delivery system. International Journl of Nanomedicine, 10 (2015) 5671–5685. R. Pignatello, E. Cenni, D. Micieli, C. Fotia, S. Avnet, M. G. Sarpietro, F. Castelli, & N. Baldini, A novel biomaterial for osteotropic drug nanocarriers: Synthesis and biocompatibility evaluation of a PLGA – ALE conjugate. Nanomedicine, 4 (2009) 161–175. S. K. Shukla, A. K. Mishra, O. A. Arotiba, & B. B. Mamba, Chitosan-based nanomaterials: A state-of-the-art review. International Journal of Biological Macromolecules, 59 (2013) 46–58. https://doi.org/10.1016/j.ijbiomac.2013.04.043.

324

9. 10.

11.

12.

13.

14.

15.

16. 17.

Chapter 11

S. Mao, W. Sun, & T. Kissel, Chitosan-based formulations for delivery of DNA and siRNA. Advanced Drug Delivery Reviews, 62 (2010) 12–27. https://doi.org/10.1016/j.addr.2009.08.004. C. Liang, B. Guo, H. Wu, N. Shao, D. Li, J. Liu, L. Dang, C. Wang, H. Li, S. Li, W. K. Lau, Y. Cao, Z. Yang, C. Lu, X. He, X. Pan, B. Zhang, C. Lu, H. Zhang, K. Yue, A. Qian, P. Shang, J. Xu, L. Xiao, Z. Bian, W. Tan, Z. Liang, F. He, L. Zhang, G. Zhang, J. Diseases, B. Proteome, T. Science, & H. Kong, Aptamer-functionalized lipid nanoparticles targeting osteoblasts as a novel RNA interferencebased bone anabolic strategy. Nat Med, 21 (2017) 288–294. https://doi.org/10.1038/nm.3791.Aptamer-functionalized. G. Wang, N. Z. Mostafa, V. Incani, C. Kucharski, & H. Uludag, Bisphosphonate-decorated lipid nanoparticles designed as drug carriers for bone diseases. Journal of Biomedical Materials Research, 100A (2011) 684–693. https://doi.org/10.1002/jbm.a.34002. C. Argyo, V. Weiss, C. Bräuchle, & T. Bein, Multifunctional mesoporous silica nanoparticles as a Universal platform for drug delivery. Chemistry of Materials, 26 (2014) 435–451. https://doi.org/10.1021/cm402592t. A. El-Fiqi, T.-H. Kim, M. Kim, M. Eltohamy, J.-E. Won, E.-J. Lee, & H.-W. Kim, Capacity of mesoporous bioactive glass nanoparticles to deliver therapeutic molecules. Nanoscale, 4 (2012) 7475–88. https://doi.org/10.1039/c2nr31775c. M. Martínez-carmona, D. Lozano, M. Colilla, & M. Vallet-regí, Lectin-conjugated pH-responsive mesoporus silica nanoparticles for targeted bone cancer treatment. Acta Biomaterialia, 65 (2017) 393– 404. https://doi.org/10.1016/j.actbio.2017.11.007. W. Sun, Y. Han, Z. Li, K. Ge, & J. Zhang, Bone targeted mesoporous silica nanocarrier anchored by zoledronate for cancer bone metastasis. Langmuir, 32 (2016) 9237–9244. https://doi.org/10.1021/acs.langmuir.6b02228. H. Tschernitschek, L. Borchers, & W. Geurtsen, Nonalloyed titanium as a bioinert metal--a review. Quintessence International (Berlin, *HUPDQ\ࣟ 1985), 36 (n.d.) 523–30. R. A. Gittens, T. McLachlan, R. Olivares-Navarrete, Y. Cai, S. Berner, R. Tannenbaum, Z. Schwartz, K. H. Sandhage, & B. D. Boyan, The effects of combined micron-/submicron-scale surface roughness and nanoscale features on cell proliferation and differentiation. Biomaterials, 32 (2011) 3395–3403. https://doi.org/10.1016/j.biomaterials.2011.01.029.

Nanoparticle-Based Drug Delivery for Bone Disorders

18. 19.

20. 21.

22.

23.

24.

25.

26.

325

F. Wei, M. Li, R. Crawford, Y. Zhou, & Y. Xiao, Exosomeintegrated titanium dioxide nanotubes for targeted bone regeneration. Acta Biomaterialia, 86 (2019) 480–492. X. Liu, X. Li, L. Zhou, S. Li, J. Sun, Z. Wang, Y. Gao, Y. Jiang, H. Lu, Q. Wang, & J. Dai, Colloids and Surfaces %ௗ Biointerfaces Effects of simvastatin-loaded polymeric micelles on human osteoblast-like. Colloids and Surfaces % Biointerfaces, 102 (2013) 420–427. https://doi.org/10.1016/j.colsurfb.2012.06.037. V. N. Mochalin, O. Shenderova, D. Ho, & Y. Gogotsi, The properties and applications of nanodiamonds. Nature Nanotechnology, 7 (2012) 11–23. https://doi.org/10.1038/nnano.2011.209. T. Ryu, R. Kang, K. Jeong, D. Jun, J. Koh, D. Kim, S. Kyung, & S. Choi, Bone-targeted delivery of nanodiamond-based drug carriers conjugated with alendronate for potential osteoporosis treatment. Journal of Controlled Release, 232 (2016) 152–160. https://doi.org/10.1016/j.jconrel.2016.04.025. S. Bose & S. Tarafder, Calcium phosphate ceramic systems in growth factor and drug delivery for bone tissue engineering: A review. Acta Biomaterialia, 8 (2012) 1401–1421. https://doi.org/10.1016/j.actbio.2011.11.017. W. Chu, Y. Huang, C. Yang, Y. Liao, X. Zhang, M. Yan, S. Cui, & C. Zhao, Calcium phosphate nanoparticles functionalized with alendronate conjugated polyethylene glycol (PEG) for the treatment of bone metastasis. International Journal of Pharmaceutics, 516 (2017) 352–363. https://doi.org/10.1016/j.ijpharm.2016.11.051. R. R. Meka, S. H. Venkatesha, B. Acharya, & K. D. Moudgil, Peptide-targeted liposomal delivery of dexamethasone for arthritis therapy. Nanomedicine, 14 (2019) 1455–1469. https://doi.org/10.2217/nnm-2018-0501. R. Heo, D. G. You, W. Um, K. Y. Choi, S. Jeon, J.-S. Park, Y. Choi, S. Kwon, K. Kim, I. C. Kwon, D.-G. Jo, Y. M. Kang, & J. H. Park, Dextran sulfate nanoparticles as a theranostic nanomedicine for rheumatoid arthritis. Biomaterials, 131 (2017) 15–26. https://doi.org/10.1016/j.biomaterials.2017.03.044. L. M. Andrade, C. de Fátima Reis, L. Maione-Silva, J. L. V. Anjos, A. Alonso, R. C. Serpa, R. N. Marreto, E. M. Lima, & S. F. Taveira, Impact of lipid dynamic behavior on physical stability, in vitro release and skin permeation of genistein-loaded lipid nanoparticles. European Journal of Pharmaceutics and Biopharmaceutics, 88 (2014) 40–47. https://doi.org/10.1016/j.ejpb.2014.04.015.

326

27.

28.

29.

30.

31.

Chapter 11

N. K. Garg, B. Singh, R. K. Tyagi, G. Sharma, & O. P. Katare, Effective transdermal delivery of methotrexate through nanostructured lipid carriers in an experimentally induced arthritis model. Colloids and Surfaces % Biointerfaces, 147 (2016) 17–24. https://doi.org/10.1016/j.colsurfb.2016.07.046. M. Bachar, A. Mandelbaum, I. Portnaya, H. Perlstein, S. Even-Chen, Y. Barenholz, & D. Danino, Development and characterization of a novel drug nanocarrier for oral delivery, based on self-assembled ȕcasein micelles. Journal of Controlled Release, 160 (2012) 164–171. https://doi.org/10.1016/j.jconrel.2012.01.004. H. M. Burt, A. Tsallas, S. Gilchrist, & L. S. Liang, Intra-articular drug delivery systems: Overcoming the shortcomings of joint disease therapy. Expert Opinion on Drug Delivery, 6 (2009) 17–26. https://doi.org/10.1517/17425240802647259. M. Zheng, H. Jia, H. Wang, L. Liu, Z. He, Z. Zhang, W. Yang, L. Gao, X. Gao, & F. Gao, Application of nanomaterials in the treatment of rheumatoid arthritis. RSC Advances, 11 (2021) 7129– 7137. https://doi.org/10.1039/D1RA00328C. S. E. Headland, H. R. Jones, L. V. Norling, A. Kim, P. R. Souza, E. Corsiero, C. D. Gil, A. Nerviani, F. Dell’Accio, C. Pitzalis, S. M. Oliani, L. Y. Jan, & M. Perretti, Neutrophil-derived microvesicles enter cartilage and protect the joint in inflammatory arthritis. Science Translational Medicine, 7 (2015). https://doi.org/10.1126/scitranslmed.aac5608.

CHAPTER 12 BIOHEAT TRANSFER AND APPLICATIONS OF MEDICAL THERMOGRAPHY IN PRECLINICAL DIAGNOSIS AND CONTROL SATHISH KUMAR GURUPATHAM Kennesaw State University, Kennesaw, Georgia, GA 30144, USA *Corresponding author: Dr Sathish Kumar Gurupatham Associate Professor Kennesaw State University Kennesaw, Georgia, United States of America Email: [email protected] Phone: +1 - 470-578-3074

Abstract Bioheat transfer is the study of thermal energy’s movement in living systems. Heat transfer plays a crucial role in temperature-based biochemical processes. Hence, the knowledge of bioheat transfer becomes more significant in understanding various functions in living systems. Further, because the mass transport of blood through tissue induces a consequent thermal energy transfer, bioheat transfer methods are applicable for diagnostic and therapeutic applications involving either mass or heat transfer. It is a novel field with a lot of research potential. This chapter will discuss the importance of thermoregulation in the human body, modes of bioheat transfer, and role of medical thermography in the preclinical diagnosis of certain medical conditions.

328

Chapter 12

Abbreviations BMR THL HP S MR W E C K R RMR m DITI FDA

Basal Metabolic Rate Total Heat Loss Rate Metabolic Heat Production rate Rate of Storage of Heat Rate Rate of Metabolic Energy Transformation Rate of Work Rate of Evaporative Heat Transfer Rate of Convective Heat Transfer Rate of Conductive Heat Transfer Rate of Radiant Heat Transfer Resting Metabolic Rate Body Mass Digital Infrared Thermal Imaging Food and Drug Administration

1. Introduction Bioheat transfer has various applications in biology, medicine and engineering. Additionally, the experts in this field (the users of this treatise) come from a wide range of backgrounds and are not equally strong in all the scientific backgrounds involved, such as biology, medicine, physics, mathematics and engineering. Thus, the authors aim to develop the topics in an orderly manner that can be followed relatively easily. The authors also focus on maintaining a balance between the biological and physical sciences to show their interactions and limitations. Although bioheat transfer focuses on the transport of thermal energy in living systems, the knowledge of the field is significant because the biochemical processes are based on body temperature. Bioheat transfer is a blend of biology, medicine and engineering. It is always a challenge to address the concerns of readers who come from various fields, such as biology, medicine, mathematics, physics and engineering, with varied knowledge levels [1]. This process involves the heat transfer modes, such as conduction, convection, radiation, metabolism and evaporation. The knowledge of bioheat transfers is applied in evaluating hyperthermia in cancer treatment, laser surgery, cryosurgery, cryopreservation, and thermal comfort. This chapter discusses the brief introductory concepts and their significance in the applications of bioheat transfer.

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

329

2. Temperature regulation of human body The species, such as insects, fish and reptiles, adjust their body temperature almost to match their environments. They are called ‘poikilotherms’, whereas the mammals and birds that do not change their body temperature with their environment are called ‘homeotherms’. The enzymes in our human body need to maintain their shape to catalyse the biochemical reactions in the cells. The change in body temperature leads to changes in their geometries, preventing them from catalysing biological processes. According to the second law of thermodynamics, heat always flows from a high-temperature body to a lower one. In most placental mammals, the average body temperature ranges from 36–40ºC, whereas the average temperature of the environment is lower than that. However, some places record higher temperatures, unusually, during the summertime. This process creates a temperature gradient between the body and the environment, specifically toward the environment [2]. During metabolism, energy is provided for life functions via chemical reactions within living body cells. This process generates heat along with energy. Additionally, heat could also be generated in the body through physical activities. This process, referred to as providing energy for living functions, also generates heat. The excess heat created must be delivered to the environment at the same rate as generated because the body temperature must be maintained in homeotherms despite changing environmental conditions.

3. Heat exchange mechanism with the environment According to the second law of thermodynamics, conduction, convection, radiation, and evaporation are the four ways of heat exchange of the human body with its terrestrial exposure. Among all these modes of heat exchange, the rate of heat transfer is proportional to the area through which heat transfer occurs. The heat transfer rate by conduction, convection and radiation corresponds to the temperature gradient between the environment and the skin surface. In contrast, for the skin’s evaporation, the variation of water vapour pressure between the environment and surface is considered instead of temperature variations [3].

330

Chapter 12

3.1 Conduction Conductive heat transfer is the transfer of energy from the random motion of molecules in a solid or fluid at a higher temperature to molecules in a solid or fluid at a lower temperature (Figure 1). The rate at which the conductive heat transfer occurs is directly proportional to the exposed area of contact through which heat transfer occurs, the temperature difference between the skin and the surrounding, and the thermal conductivity of the contact material. Suppose the contribution from the conduction mode has to be a significant share in the heat exchange process of the body. In that case, the temperature gradient between the body’s surface temperature and the surrounding temperature needs to be large, and the contact material’s conductivity must be higher. The body’s surface contact also needs to be significant. The lying position is the best example of the conductive heat transfer than sitting or standing positions [4].

3.2 Convection The transfer of heat by convection, which involves the mass transfer of medium (air or water), is considered in two ways (Figure 1). In the first case, the medium surrounding the body undergoing convective heat transfer is considered calm without any current. The medium moves with some velocity that expedites the heat transfer rate in the second case. The first case is natural convection, whereas the second is forced convection. Blowing the air from our mouth to cool the soup before having it is an example of forced convection.

3.3 Natural convection Consider a person standing in a closed room, and the mean temperature of their skin is 33ºC, and the surrounding air temperature is 24ºC. By conduction, the cooler air in touch with their skin becomes warmer and moves up, and the next layer of cold air meets the skin and undergoes the same. This natural convection mechanism transfers heat from the human body [5].

3.4 Forced convection It is naturally understandable that the heat loss by convection for a given temperature between the air and body surface is increased by increasing the medium velocity. This outcome can be observed when a fresh breeze on

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

331

sunny days can make you feel cold. The laws of fluid dynamics govern the process of convection. They include some physical properties of the body’s medium, size, and shape. For instance, at the air temperature of 25°C, the temperature of the person’s skin was 33°C. By natural convection in still air, they lost 24 W/m2. If this person is exposed to a breeze with a velocity of 5 m/s, then the heat loss would be around 150 W/m2 by forced convection [6].

Figure 12-1. Heat exchange of human body with Surroundings through various modes. Created with BioRender.com

3.5 Radiation Absorbing and emitting electromagnetic radiation occurs in all gases, liquids and solids. The emitted energy from a plane surface of a full radiator or black body is proportional to the absolute surface temperature’s fourth power (Figure 12-1): Thermal radiation (W/m2) = ıT4

(1)

where the Stefan-Boltzmann constant (ı) is 5.67 X 10í8 (W/m2·K4). The sun could be considered a black body that emits thermal radiation at 74,000 kW/m2. However, not more than 1360 W/m2 of its thermal radiation arrives at the earth’s outer atmosphere, known as the solar constant, due to

Chapter 12

332

the sun’s distance from the earth. The human body interacts with its surroundings by absorbing and emitting radiation [7].

3.6 Solar radiation It is important to know that all solar radiation at the earth’s outer atmosphere does not reach the ground. However, this amount is sufficient to impact the earth’s living beings greatly. Gases scatter part of solar radiation, which is on the way to the atmosphere, causing the solar radiation to go to the ground or back to space. When they arrive at the surface, the direct and second fraction solar radiation combine diffuse and direct short-wave solar radiation. One of the ways to save large animals in regions where solar radiation is intense is to provide furs’ insulation [8].

4. Evaporation Evaporation is when the water changes its phase from liquid to vapour by absorbing heat. Depending on the temperature, the amount of heat (latent) is nearly 2.4 kJ per gram of water to be evaporated. 7KH EDVDO KHDW SURGXFWLRQRIDQDGXOWKXPDQ is 40 W, which could be removed by the evaporation rate of 2 g per minute. The difference in water vapour pressure and surrounding air pressure, wind velocity, size, and shape of the body affect the rate of evaporative heat loss. A total of 149 ml of water [9] per hour should be evaporated to dissipate the heat generated by basal metabolic rate (BMR) in a human by the weight of 80 kg, which is approximately 0.3% of the total water content in the human body. The gain of radiant heat per unit of body mass is more significant when the surface area is larger. This would be another drain on storing the body water in the evaporation environment, which is a significant challenge in dry regions.

4.1 Heat balance equation The body temperature remains constant if the heat loss and heat production are equal. Here, the body’s temperature is defined as the mean temperature of the entire body. In contrast, the body’s core temperature is another concept [10]. Therefore, the heat balance equation for the constant mean temperature of the body in the simplest form can be written as follows: THL = rate of total heat loss

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

333

HP = THL HP = rate of metabolic heat production Dimensions are watts (W), preferably in relation to a unit area of body surface (W·mí2) or unit body mass (W·kgí1). S = MR í (± W) í (± E) í (± C) í (± K) í (± R) S = rate of storage of heat (positive = increase in body heat content, negative = decrease in body heat content) MR = rate of metabolic energy transformation (always positive in a living animal. During rest MR = HP during positive work MR = HP + W) W = rate of work (positive = external work accomplished, negative = mechanical work absorbed by the body) E = rate of evaporative heat transfer (positive = evaporative heat loss, negative = evaporative heat gain) C = rate of convective heat transfer (positive = transfer to the environment, negative = transfer into the body K = rate of conductive heat transfer (positive = transfer to the environment, negative = transfer into the body) R = rate of radiant heat transfer (positive = transfer to the environment, negative = heat absorption by the body)

4.2 Metabolic rate and body mass BMR is the number of calories that the human burns as the body performs a basic (basal) life-sustaining function. Resting metabolic rate (RMR) is defined as the number of calories burned while staying in bed all day [11]. BMR defines the basal metabolism rate, which makes up about 60–70% of the spent calories. This includes the energy that the human body uses to maintain the basic functions of living and breathing, including: x x x x

Our heartbeat Cell production Respiration The maintenance of body temperature

Chapter 12

334

x Circulation x Nutrient processing The unique metabolism rate, or BMR, is influenced by several factors: age, weight, height, gender, environmental temperature, dieting, and exercise habits. The BMR per unit of body mass is related to the body mass inversely in many species, from elephants to mice. Formally, this expression can be written as a related equation for BMR and body mass of whole animals: where: BMR = k·m3/4 BMR = basal metabolic rate = basal rate of heat production m = body mass k = 3.4, if m in kg and BMR in Watts Increasing two-thirds power of the body mass with similar shapes results in enhanced surface area. Hence, the surface area per unit of body mass is largest in the small bodies. Increasing the mass decreases the surface area [2,3][12,13]. Further, the rate of heat exchange per unit of body mass is higher because the size of the surface area is proportional to the heat exchange within the environment if all other parameters are equal.

4.3 Metabolic heat production v. metabolic rate The rate of converting chemical energy to heat is metabolic heat production (HP). In contrast, the metabolic rate (MR) is expressed as a rate of chemical energy conversion to heat and additional mechanical work. It would be equal when no physical work is involved, such as in a resting condition. This increase in the MR will be maintained until a certain point, after which it is delivered to heat [3,4][13,14].

5. Sweating In humans, the sweat secreted from the sweat gland takes away the body heat in the form of latent heat while evaporating.

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

335

6. Panting Many species do not have sweat glands. Therefore, they rely on panting to enhance the evaporative heat loss. The flowing of air on inhalation over the respiratory tract’s wet surfaces is caused by the efficient exchange of heat and water. This is done by the saturation of exhaled air at internal body temperature. The respiratory evaporative heat loss rate depends on the ventilation rate and inhaled air’s water vapour pressure. Therefore, these two factors affect the maximum heat loss by panting [14].

7. Shivering Shivering is an involuntary tremor of skeletal muscle. It is considered here exclusively as a thermoregulatory effector mechanism for increasing HP. The rhythm is generated in the spinal cord, and its frequency depends on body mass. Generally, it is of the order of 10 Hz in adult humans [4,5][14,15].

8. External insulation: Fur The peculiar properties of some hair do not change the level of clothing insulation. However, it maintains trapped air around the body and transferring of heat is done by conduction primarily [5,6][15,16].

9. Internal insulation: Fat and the principle of core and shell A subcutaneous layer of blubber or fat is mainly responsible for permanent or seasonal cold environments in bare-skinned animals, such as terrestrial mammals and aquatic. The insulation value of fur is superior compared to fat. Hence, the thickness of the blubber needs to be more.

10. Temperature regulating mechanism The integrated temperature information is obtained from the inputs of multiple body sensors across the human body. Importantly, the hypothalamus in the brain is stimulated by electrical signals (Figure 12-2). According to the need of whether to remove or retain the heat, certain activities are initiated, as discussed below.

Chapter 12

336

For instance, naturally, the body will start losing heat to the surroundings due to the temperature gradient when the surrounding temperature goes down. The body initiates the following activities to generate the heat in the body to maintain its heat balance and temperature (Figure 12-3): 1. Blood vessel constriction 2. Brown fat burning 3. Skeleton muscle shivering Blood vessel constriction: The blood flow rate through the blood vessels is decreased by restricting the blood vessel itself. Brown fat-burning: Brown fat, also known as the brown adipose tissue, is a special type of body fat that is activated to burn energy and generate heat. This brown fat is activated only when the body is subjected to a cold environment. Skeleton muscle shivering: The heat is generated by the voluntary contractions from the skeleton muscles. In contrast, the body gains heat from the surrounding during the increased surrounding temperature. This excess heat needs to be transferred to the surroundings to maintain the body temperature. The body initiates the following activities: 1. Blood vessel dilation 2. Sweat gland activation

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

Figure 12-2. Temperature regulation mechanism. Created with BioRender.com

Figure 12-3. Temperature regulation activities. Created with BioRender.com

337

338

Chapter 12

Blood vessel dilation (vasodilation): The dilatation of blood vessels raises the blood flow rate, promoting heat transfer to the skin and, consequently, to the environment. Sweat gland activation: In humans, the maximum sweat rate is between 10 and 15 g per minute per square metre of the surface area. Hence, sweat secretion is a powerful mechanism for temperature regulation [6,7][16,17]. However, the evaporation of sweat occurs at large when the pressure of water vapour on the skin is greater than the pressure of surrounding air. This process is called saturation vapour pressure, increasing exponentially with temperature. At a given temperature, the air has a maximum water vapour pressure. At this point, the relative humidity is 100%.

11. Bioheat transfer mathematical model The knowledge of bioheat transfer phenomena has applications in the prevention, treatment, preservation and protection techniques for biological systems, such as heat or cold treatments to destroy tumours, improve patients’ outcomes after brain injury, and protect humans from extreme environmental conditions. Although blood flow effects on heat transfer in living tissue have been under investigation for more than a century, the mathematical modelling has remained standstill due to the complexity of thermal interaction between the vasculature and tissue. There are two approaches to understanding the effect of blood flow in a biological system, namely continuum models and vascular models [18]. In the continuum model, the effect of blood flow in the region of interest is averaged over a control volume, assuming there is no blood vessel present. Instead, a term is introduced in the equation to represent that effect. The continuum models are simple to use, whereas, in vascular models, the blood vessels are represented as tubes buried in tissue. Hence, only a few tubes are considered to neglect others.

11.1 Pennes bioheat transfer model It is a popular continuum model wherein the effect of blood flow in the tissue is modelled as a heat source or sink to the traditional heat conduction equation. The Pennes bioheat equation is expressed as follows:

(1)

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

339

where (qm) is the metabolic heat generation in the tissue, and the second term (qblood) on the right side of the equation considers the contribution of blood flow to the local tissue temperature distribution.

12. Thermography All objects at a temperature above absolute zero (0 K) emit thermal radiation in electromagnetic waves (or photons) because of the changes in the atoms or molecules’ electronic configurations. This means that every object we come across will radiate thermal energy because no object can attain 0 K practically. Thermal imaging cameras, also called infrared cameras, create thermal images by detecting the body’s thermal radiation. Every object has a unique thermal signature. It changes when moisture, heat, cold or wood destroying insects are introduced. The changes can be small or large. However, this incredible thermal image scanning technology can detect the thermal signatures, which is impossible through naked eyes [19]. A thermal camera has lenses, just like visible light cameras. However, in this case, the lens focuses on the waves from infrared energy present in all objects onto an infrared sensor array. Thousands of sensors on the array convert infrared energy into electrical signals, which create a video image. The infrared camera measures and displays a ‘thermal profile’ of objects concerning the temperature of surrounding objects. So, a person, warmer than the surrounding air, appears ‘white’, whereas the cooler surrounding air or buildings will appear in varying shades of grey. Based on the analysis of the colour spectrum, it is easy to understand the temperature profile of the bodies (Figure 12-4(a) and (b)). The human body maintains a constant temperature while overcoming the various forms of heat generation and heat loss that might happen suddenly and temporarily, as explained in the above section. Medical thermography (digital infrared thermal imaging [DITI]) is a popular preclinical diagnosis and control tool for treating homeostatic imbalances. Thermography uses the heat from the human body non-invasively to diagnose a host of health care conditions. Thermography is completely safe and uses no radiation. Medical thermography equipment usually has an IR camera and a standard PC or laptop computer. Medical thermography equipment usually has two parts: the IR camera and a standard PC or laptop computer. These systems

340

Chapter 12

have only a few controls and are relatively easy to use. The system measures temperatures ranging from 10°C to 55°C with an accuracy of 0.1°C.

Figure 12-4 (a) Thermal images of four avocados at different ripeness levels and their variation in temperatures

Figure 12-4(b) Thermal images of four peaches at different ripeness levels and their variation in temperatures

With the advent of technology, high-speed computers and accurate thermal imaging cameras capture the heat signature of the human body. Different medical conditions can be determined based on the abnormally hot and cold areas of the specific regions of the body. The Food and Drug Administration (FDA) for medical devices has approved the thermography procedure. This procedure is applied to the screening for breast cancer, extra-cranial vessel disease (head and neck vessels), neuro-musculoskeletal disorders and vascular disease of the lower extremities.

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

341

13. Some of the standard applications of thermography x x x x x

Breast pathologies Extra-cranial vessel disease Neuro-musculo-skeletal Vertebrae (nerve problems/arthritis) Lower extremity vessel disease

13.1 Breasts pathologies The early detection of breast cancer, benign tumours, mastitis and fibrocystic breast disease is facilitated by thermography [20]. Thermography captures the thermal signature of the breast, which needs to be investigated for the cancer development along with the other breast and compared to evaluate the difference. The abnormal thermal signature of the affected breast indicates the tumour’s metabolism. This method does not require the conventional and painful procedure of compressing the breast in question during the diagnosis. Importantly, there is no radiation involved.

13.2 Extra-cranial vessel disease Similarly, multiple conditions related to the blood flow through the neck vessels and head could be assessed with thermal imaging. Additionally, vascular disease leading to stroke could be detected early. Various types of headaches [21], such as migraine, facial nerve injury, and visualisation disorders, can also be detected by thermal images of the head and neck. Moreover, dental decay and cavities could also be detected without conventional X-rays.

13.3 Neuro-musculo-skeletal Thermography accurately diagnoses neck and back disorders [22] and has been in use since the late 1970s. The strained or torn muscles release chemicals, which increase the heat liberated and are seen in thermal images around the region. Similarly, back strains produce heat patterns that indicate spinal injuries. Thermography can demonstrate the permanency of spinal injuries leading to a person’s disability. It has been used in the trial courts for many years to prove injury and assist in rating permanent impairment.

342

Chapter 12

13.4 Lower extremity vessel disease Thermography can also detect painlessly deep vein thrombosis and other circulatory disorders of the lower extremities, which might cause the loss of a limb or stroke if unchecked early. Thermal imaging of feet helps to detect diabetes in the early stages In such cases, the thermal images of feet would show colder by 1–2ºC than the lower leg, even several years before. It gives more time to treat and prevent diabetes [23].

References 1. 2. 3.

4. 5.

6.

7. 8.

Y. C. and A. Ghajar, Heat and Mass 7UDQVIHU Fundamentals and Applications, 6th Editio (McGraw-Hill Education, 2020). C. L. Tan & Z. A. Knight, Regulation of Body Temperature by the Nervous System. Neuron, 98 (2018) 31–48. https://doi.org/10.1016/j.neuron.2018.02.022. T. Kurazumi, Y., Tsuchikawa, T., Kakutani, K., Yamato, Y., Torii, T., Matsubara, N. and Horikoshi, Mean skin temperature taking into account convective heat transfer DUHDVௗ Calculation method of seiza sitting, cross-legged sitting, sideway sitting, both-kness-erect sitting, leg-out sitting, lateral and supine positions. Journal of Environmental Engineering (Transactions of AIJ), 69 (2004) 19–26. https://doi.org/10.3130/aije.69.19_2. A. P. Gagge & R. R. Gonzalez, Mechanisms of Heat Exchange: Biophysics and Physiology. Compr. Physiol. (Wiley, 1996), pp. 45– 84. https://doi.org/10.1002/cphy.cp040104. R. Nielsen & B. Nielsen, Influence of skin temperature distribution on thermal sensation in a cool environment. European Journal of Applied Physiology and Occupational Physiology, 53 (1984) 225– 230. https://doi.org/10.1007/BF00776594. R. Nielsen & B. Nielsen, Measurement of mean skin temperature of clothed persons in cool environments. European Journal of Applied Physiology and Occupational Physiology, 53 (1984) 231–236. https://doi.org/10.1007/BF00776595. T. Horikoshi, A Review of Thermal Comfort Studies in Japan. Indoor Environment, 2 (1993) 325–330. https://doi.org/10.1177/1420326X9300200511. M. N. Mead, Benefits of Sunlight: A Bright Spot for Human Health. Environmental Health Perspectives, 116 (2008). https://doi.org/10.1289/ehp.116-a160.

Bioheat Transfer and Applications of Medical Thermography in Preclinical Diagnosis and Control

9. 10.

11.

12.

13. 14.

15.

16. 17. 18.

19.

343

M. S. Popson, M. Dimri, & J. Borger, Biochemistry, Heat and Calories (2022). Glossary of terms for thermal physiology. Second edition. Revised by The Commission for Thermal Physiology of the International Union of Physiological Sciences (IUPS Thermal Commission). Pflugers $UFKLYࣟ European journal of physiology, 410 (1987) 567– 87. K. Brück, Thermal Balance and the Regulation of Body Temperature. Hum. Physiol. (Berlin, Heidelberg: Springer Berlin Heidelberg, 1989), pp. 624–644. https://doi.org/10.1007/978-3642-73831-9_25. S. L. Heathcote, P. Hassmén, S. Zhou, L. Taylor, & C. J. Stevens, How Does a Delay Between Temperate Running Exercise and HotWater Immersion Alter the Acute Thermoregulatory Response and Heat-Load? Frontiers in Physiology, 10 (2019). https://doi.org/10.3389/fphys.2019.01381. C. R. Taylor, Relating mechanics and energetics during exercise. Advances in veterinary science and comparative medicine, 38A (1994) 181–215. Y. Kurazumi, K. Fukagawa, T. Sakoi, A. Naito, M. Imai, R. Hashimoto, E. Kondo, & T. Tsuchikawa, Clothing Thermal Insulation for Typical Seasonal Clothing of Infant with Infant Thermal Manikin. Engineering, 13 (2021) 372–387. https://doi.org/10.4236/eng.2021.137027. R. L. Marsh, D. J. Ellerby, J. A. Carr, H. T. Henry, & C. I. Buchanan, Partitioning the Energetics of Walking and Running: Swinging the Limbs Is Expensive. Science, 303 (2004) 80–83. https://doi.org/10.1126/science.1090704. J. De Witte & D. I. Sessler, Perioperative Shivering. Anesthesiology, 96 (2002) 467–484. https://doi.org/10.1097/00000542-20020200000036. H. T. Hammel, Thermal Properties of Fur. American Journal of Physiology-Legacy Content, 182 (1955) 369–376. https://doi.org/10.1152/ajplegacy.1955.182.2.369. D. R. Hodgson, L. J. McCutcheon, S. K. Byrd, W. S. Brown, W. M. Bayly, G. L. Brengelmann, & P. D. Gollnick, Dissipation of metabolic heat in the horse during exercise. Journal of Applied Physiology, 74 (1993) 1161–1170. https://doi.org/10.1152/jappl.1993.74.3.1161. C. K. Charny, Mathematical Models of Bioheat Transfer. (1992), pp. 19–155. https://doi.org/10.1016/S0065-2717(08)70344-7.

344

20.

21. 22.

23.

Chapter 12

E. Y. K. Ng & E. C. Kee, Advanced integrated technique in breast cancer thermography. Journal of Medical Engineering & Technology, 32 (2008) 103–114. https://doi.org/10.1080/03091900600562040. J. W. Lance & M. Anthony, Thermographic studies in vascular headache. Medical Journal of Australia, 1 (1971) 240–243. https://doi.org/10.5694/j.1326-5377.1971.tb87544.x. N. Zaproudina, Z. Ming, & O. O. P. Hänninen, Plantar Infrared Thermography Measurements and Low Back Pain Intensity. Journal of Manipulative and Physiological Therapeutics, 29 (2006) 219–223. https://doi.org/10.1016/j.jmpt.2006.01.003. J. J. van Netten, J. G. van Baal, C. Liu, F. van der Heijden, & S. A. Bus, Infrared Thermal Imaging for Automated Detection of Diabetic Foot Complications. Journal of Diabetes Science and Technology, 7 (2013) 1122–1129. https://doi.org/10.1177/193229681300700504.