ADME Processes in Pharmaceutical Sciences: Dosage, Design, and Pharmacotherapy [2 ed.] 3031504186, 9783031504181

Absorption, Distribution, Metabolism and Excretion (ADME) processes and their relationship with the design of dosage for

110 74 17MB

English Pages 501 [489] Year 2024

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface
Contents
Part I: The Basics of ADME Processes
Chapter 1: Introduction: Biopharmaceutics and Pharmacokinetics
1.1 Presentation
1.2 Some Practical Definitions: The Notion of Bioavailability
1.3 The Relationship Between Biopharmaceutics and Pharmacokinetics and the Pharmacological Response
1.4 The Free Drug Hypothesis
References
Further Reading
Chapter 2: Drug Release
2.1 Introduction
2.2 Zero-Order Drug Release
2.3 First-Order Drug Release
2.4 Intermediate Kinetics: Peppas Model
2.5 Drug Dissolution: The Noyes-Whitney Model and Its Derivatives
References
Further Reading
Chapter 3: Drug Absorption
3.1 Introduction
3.2 Factors Affecting Drug Absorption
3.3 The Fluid Mosaic Model and Beyond
3.4 How Can Drugs Be Absorbed?
3.5 Free (or Simple) Diffusion
3.5.1 Sink Condition and Absorption
3.5.2 The Impact of pH on Drug Absorption
3.6 Facilitated Diffusion
3.7 Active Transport
References
Further Reading
Chapter 4: Drug Distribution
4.1 Introduction
4.2 Factors Impacting on [The Extent and Kinetics of] Drug Distribution
4.3 Different Pharmacokinetic Models Describe Drug Distribution in Different Ways
4.4 Real and Apparent Volumes of Distribution
4.5 The Source of Disagreement: Discrepancies Between Apparent and Real Volumes
4.5.1 Binding to Plasma Proteins and Blood Cells
4.5.1.1 Binding Partners
4.5.2 Binding to Extravascular Tissue Elements
4.6 Pharmacokinetic Impact of Binding to Plasma Proteins and Extravascular Tissues
4.7 Some Therapeutic Considerations Related to Displacement from Plasma Protein and Tissue Binding Sites
References
Further Reading
Chapter 5: Drug Metabolism
5.1 Introduction: Definition of Xenobiotics
5.2 Metabolism: Biotransformation Reactions
5.3 Functionalization (Phase I) Reactions
5.3.1 Cytochrome P450
5.4 Synthetic (Phase II) Reactions
5.5 Phase III
5.6 First-Pass Effect
5.7 Factors Affecting Drug Metabolism: Intrinsic and Extrinsic Sources of Variability
5.7.1 Internal Factors
5.7.1.1 Genetic Differences
5.7.1.2 Sex Differences
5.7.1.3 Age
5.7.1.4 Variation Induced by Physiological and Disease States
5.7.2 External Factors
References
Further Reading
Chapter 6: Drug Excretion
6.1 Introduction
6.2 Systemic or Total Body Clearance
6.3 Organ Clearance
6.4 Renal Excretion
6.4.1 Glomerular Filtration
6.4.2 Tubular Secretion
6.4.3 Tubular Reabsorption
6.4.4 Renal Clearance
6.5 Intestinal (Hepatobiliary) Excretion
6.5.1 Enterohepatic Recycling
References
Further Reading
Chapter 7: Routes of Drug Administration
7.1 Introduction
7.2 Parenteral Routes of Administration
7.2.1 Intravenous (IV)
7.2.1.1 Pharmacokinetic characteristics of IV administration
7.2.1.2 Physicochemical and Technological Considerations for IV Drug Administration
7.2.2 Intramuscular (IM)
7.2.2.1 The Intramuscular Absorption Process
7.2.2.2 Technological Factors That Modify the IM Absorption of Drugs
7.2.3 Subcutaneous (SC)
7.2.3.1 The Subcutaneous Absorption Process
7.2.3.2 Formulations for Subcutaneous Administration
7.2.4 Specialized Parental Routes
7.3 Enteral Routes of Administration
7.3.1 Oral Administration of Drugs
7.3.1.1 Gastrointestinal Tract: Anatomical and Physiological Considerations
7.3.1.2 Drugs Absorption in the Gastrointestinal Tract
7.3.1.3 Primary Factors Influencing oral Drug Absorption
7.3.1.4 Secondary Factors Influencing GI Drug Absorption
7.4 Other Routes of Drug Administration
7.4.1 Buccal and Sublingual Routes of Administration
7.4.1.1 Pharmaceutical Forms for Sublingual and Buccal Administration
7.4.2 Rectal Administration of Drugs
7.4.2.1 Absorption of Drugs by the Rectal Route
7.4.3 Pulmonary Administration of Inhaled Drugs
7.4.3.1 Pulmonary Absorption of Drugs
7.4.3.2 Factors Influencing Pulmonary Drug Deposition
7.4.4 Nasal Route of Drug Administration
7.4.4.1 Physiological Factors Influencing the Nasal Absorption of Drugs
7.4.4.2 Technological Factors That Influence the Nasal Absorption of Drugs
References
Chapter 8: Compartmental Pharmacokinetic Models
8.1 Introduction
8.2 One-Compartment Pharmacokinetic Model
8.2.1 Intravenous Administration: Bolus
8.2.2 Intravenous Administration: Infusion at Constant Rate
8.2.3 Extravascular Administration of Immediate Release Dosage Forms
8.2.4 Multiple-Dosing Regimens
8.3 Two-Compartment Pharmacokinetic Model
8.3.1 Intravenous Administration: Bolus
8.3.2 Intravenous Administration: Infusion at Constant Rate
8.3.3 Extravascular Administration
8.4 Volumes of Distribution
References
Further Reading
Chapter 9: Regulatory Requirements and Applications of Physiologically Based Pharmacokinetic Models
9.1 Introduction
9.2 Current PBPK Regulatory Guidelines
9.3 Specific Concepts Related to PBPK Modelling in the Regulatory Field
9.3.1 Model Development
9.3.2 Model Evaluation
9.3.3 Model Verification
9.3.4 Model Validation
9.3.5 Model Refinement
9.3.6 Parameter Estimation Vs Optimization
9.3.7 Local and Global Sensitivity Analysis
9.3.8 Management of Inter-Individual Variability
9.3.9 Model Application
9.3.10 PBPK Modelling Tools
9.4 Relevance of PBPK Models for Regulatory Purposes
9.5 Applications of PBPK Models
9.5.1 DDI Prediction
9.5.2 Paediatrics
9.5.3 Organ Impairment
9.5.4 Food Effect
References
Chapter 10: Drug-Drug and Food-Drug Interactions of Pharmacokinetic Nature
10.1 Introduction
10.2 Plasma Protein Binding
10.2.1 Cyclosporine-Statins: A Case Where the Bound Moiety Is Relevant
10.3 Membrane Transport
10.3.1 Valproic Acid as a Substrate for Drug Transportation
10.4 Metabolism
10.4.1 Carbamazepine-Phenytoin: Mixed Effect of Interaction at the Enzyme and Transporter Levels
10.4.2 Rifampicin: Another Example of Enzymes and Transporters Induction
10.4.3 Valproic Acid-Lamotrigine
10.4.4 Azole Antifungals-Cyclosporine
10.5 Environmental pH
10.6 Food Effects
10.6.1 Food-Drug Interactions Based on Induced Changes in GIT Physiology
10.6.2 Food-Drug Interactions Based on Specific Mechanisms
10.6.3 Food-Effects on Bioequivalence and Regarding the Sex of Individuals
10.6.4 Food-Effects on Drug Disposition
References
Part II: Specialized Topics
Chapter 11: Nanocarriers: Delivery Routes
11.1 Introduction
11.2 Delivery Routes
11.2.1 Oral Route
11.2.2 Nasal Route
11.2.3 Parenteral Route
11.2.4 Ocular Route
11.2.5 Cutaneous Route
11.3 Conclusion
References
Chapter 12: Advanced Techniques for Quality Assessment of Nanocarriers
12.1 Introduction
12.2 Supramolecular Structure, Morphology, and Size Measurements
12.2.1 Supramolecular Structure
12.2.2 Morphology
12.2.3 Size Measurement
12.3 Quantitative Analysis to Characterize Nanocarriers
12.4 Conclusions
References
Chapter 13: Nanomedicines Obtained by 3D Printing
13.1 Introduction
13.2 Nanotechnology in Drug Delivery
13.2.1 Nanocrystals
13.2.2 Organic Nanocarriers
13.2.2.1 Lipid-Based Nanoparticles
13.2.2.2 Polymeric Nanoparticles
13.2.3 Inorganic Nanocarriers
13.2.3.1 Mesoporous Silica nanoparticles (MSN)
13.2.3.2 Other Inorganic Nanoparticles
13.3 3DP of Pharmaceuticals
13.3.1 Semisolid Extrusion (SSE)
13.3.2 Fused Deposition Modelling (FDM)
13.3.3 Inkjet Printing (IP)
13.3.4 Stereolithography (SLA)
13.4 Association Between 3DP and Nanotechnology
13.5 Other Applications
13.6 Quality Assessment and Future Perspectives
13.7 Conclusion
References
Chapter 14: Absorption, Distribution, Metabolism and Excretion of Biopharmaceutical Drug Products
14.1 Introduction
14.2 Absorption of Protein Drugs
14.3 Distribution of Protein Drugs
14.4 Metabolism and Excretion of Protein Drugs
14.5 Physicochemical Property Modification to Alter Drug Pharmacokinetics: Application to Subcutaneous Insulin
14.6 ADME Properties of Heparin and Low-Molecular Weight Heparins
14.7 ADME Properties of Gene-Based Biopharmaceuticals
14.8 Conclusions
References
Chapter 15: In Vitro and In Silico ADME Prediction
15.1 Introduction
15.2 Absorption
15.2.1 In Vitro Methods
15.2.1.1 Solubility
15.2.1.2 In Vitro Release and Dissolution Testing
15.2.1.3 Permeability
15.2.1.4 Active Transport
15.2.1.5 Gut Wall Metabolism
15.2.2 In Silico Methods
15.2.2.1 Solubility
15.2.2.2 Drug Release and Dissolution
15.2.2.3 Permeability
15.2.2.4 Active Transport
15.2.2.5 Gut Wall Metabolism
15.2.2.6 Dynamic Transit Models
15.3 Distribution
15.3.1 In Vitro Methods
15.3.2 In Silico Methods
15.4 Metabolism
15.4.1 In Vitro Methods
15.4.2 In Silico Methods
15.5 Excretion
15.5.1 In Vitro Methods
15.5.2 In Silico Methods
15.6 Physiologically-Based Pharmacokinetic Models
15.7 Conclusion
References
Chapter 16: Relationship Between Pharmacokinetics and Pharmacogenomics, and Its Impact on Drug Choice and Dose Regimens
16.1 Introduction
16.1.1 Drug Metabolism: Pharmacokinetics and Pharmacodynamics
16.1.1.1 The ADME Processes
16.1.2 Efficacy and Safety of Drugs
16.1.3 Drug Variability, Pharmacogenetics, Pharmacogenomics and the Role of Epigenetics
16.2 How Can We Explain Response Variability Due to Genetics
16.2.1 Type of Variations and Their Impact on Protein Function: Polymorphisms
16.2.2 Drug Metabolizing Enzymes
16.2.2.1 Phase I Enzymes
16.2.2.2 Phase II Enzymes
16.2.3 Drug Transporters
16.2.4 Other Polymorphic Targets
16.3 Relationship Between Pharmacogenetics and ADME Processes
16.3.1 Absorption
16.3.2 Distribution
16.3.3 Metabolism (Biotransformation)
16.3.4 Excretion
16.3.5 Drug Toxicity (Safety)
16.4 Dosing Modifications and Drug Selection
16.5 Integrative Pharmacogenomic Clinical Cases
16.6 Conclusions
References
Chapter 17: The Relationship Between Pharmacogenomics and Pharmacokinetics and Its Impact on Drug Choice and Dosing Regimens in Pediatrics
17.1 Introduction
17.2 Challenges in Pediatric Pharmacotherapy
17.2.1 Lack of Evidence in Pediatric Pharmacotherapy
17.2.2 Off-Label Use of Medications and Potential Risks Associated With It
17.2.3 Lack of Appropriate Formulations
17.3 Role of Pharmacogenomics and Pharmacogenetics in Pediatric Dosing
17.3.1 Role of Ontogeny in Pediatric Pharmacogenomics
17.3.2 Pharmacogenomic Testing in Pediatric Practice
17.4 Impact of Genetic Polymorphisms on the Pharmacokinetics of Drugs in Pediatrics
17.4.1 Codeine
17.4.2 Celecoxib
17.4.3 Azathioprine, 6-Mercaptopurine, and 6-Thioguanine
17.4.4 Methotrexate
17.4.5 Tacrolimus
17.4.6 Voriconazole
17.5 Conclusions and Future Directions
References
Chapter 18: Bioavailability and Bioequivalence
18.1 Introduction: Why Bioavailability and Bioequivalence Studies?
18.2 Bioavailability/Bioequivalence Studies Through a Drug Product Lifecycle
18.2.1 Investigational New Drug Period and New Drug Development
18.2.2 Generic Drug Development
18.2.3 The Scale-Up and Post-approval Changes (SUPAC)
18.3 Bioavailability and Bioequivalence Studies
18.3.1 Absolute and Relative Bioavailability Studies
18.3.2 Mass Balance Studies
18.3.3 Bioequivalence Studies
18.3.3.1 General Pharmacokinetic Study Design
18.3.3.2 Reference and Test Product
18.3.3.3 Subjects
18.3.3.4 Sample Collection and Sampling Times
18.3.3.5 Subjects with Pre-dose Plasma Concentrations
18.3.3.6 Data Deletion Because of Vomiting
18.3.3.7 Evaluation and Methods to Document BA/BE: Use of Plasma Drug Concentration
18.3.3.8 Evaluation and Methods to Document BA/BE: Use of Urinary Drug Excretion
18.3.4 Statistical Analysis and Acceptance Limits to Establish Bioequivalence
18.4 Bioequivalence Studies
18.4.1 The Basis for Determining Bioequivalence
18.4.2 Methods for Determining Bioequivalence
18.4.2.1 Pharmacokinetic (PK) Studies
18.4.2.2 In Vitro Studies: Biopharmaceutics Classification System (BCS) and BCS-Based Biowaiver
Biopharmaceutics Classification System
18.4.2.3 Other In Vitro Studies
Dissolution/Drug-Release Testing
18.4.2.4 Pharmacodynamic Studies
18.4.3 Statistical Approaches for Bioequivalence
18.4.4 Supra-Bioavailability
18.5 Conclusion
References
Chapter 19: Drug Transporters
19.1 Introduction
19.2 ATP-Binding Cassette (ABC) Transporters
19.3 Solute Carriers (SLC)
19.4 Cooperation Is Key
References
Further Reading
Chapter 20: Blood Flow Distribution and Membrane Transporters as Determinant Factors of Tissue Drug Concentration
20.1 Basic Principles of Drug Exchange Between Blood and Tissues
20.1.1 Tissue Uptake of Solutes
20.1.2 Return of Solutes Into the Bloodstream
20.2 Role of Membrane Transporters in the Mass Transfer of Solutes
20.2.1 Conveyor Concept
20.2.2 Influx and Efflux Transporters
20.2.3 Circadian Rhythm of Efflux Transporter Activity
20.3 Physiologically-Based Pharmacokinetic Modeling of Drug Disposition
20.3.1 Mass Transfer from Arteries to Veins
20.3.2 Mass Transfer from Blood to the Outside of the Body
20.3.3 Proposal for Updating the PBPK Model’s Structure
References
Index
Recommend Papers

ADME Processes in Pharmaceutical Sciences: Dosage, Design, and Pharmacotherapy [2 ed.]
 3031504186, 9783031504181

  • 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

Alan Talevi · Pablo A. Quiroga   Editors

ADME Processes in Pharmaceutical Sciences Dosage, Design, and Pharmacotherapy Second Edition

ADME Processes in Pharmaceutical Sciences

Alan Talevi • Pablo A. Quiroga Editors

ADME Processes in Pharmaceutical Sciences Dosage, Design, and Pharmacotherapy Second Edition

Editors Alan Talevi Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP) La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET) CCT La Plata, La Plata, Argentina

Pablo A. Quiroga Quality Control of Pharmaceutical Products Pharmaceutical Toxicology, Faculty of Extract Sciences University of La Plata (UNLP) La Plata, Argentina Laboratorios Bagó S.A. La Plata, Argentina

ISBN 978-3-031-50418-1    ISBN 978-3-031-50419-8 (eBook) https://doi.org/10.1007/978-3-031-50419-8 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2018, 2024 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland Paper in this product is recyclable.

Preface

We present this second edition of the ADME Processes in Pharmaceutical Sciences and acknowledge the reception of the 2018 edition in the academic community. The structure of this volume is quite similar to that of the preceding one: the first part of the book is dedicated to the basic aspects of biopharmacy and pharmacokinetics, whereas the second part is dedicated to burgeoning topics in the field, including pharmaceutical nanocarriers; pharmacogenomics; bioequivalence studies; in vitro and in silico ADME models; the particularities of the ADME processes for biological medical products/biopharmaceuticals; drug transporters; and the relationship between cardiac output, pharmacokinetics, and circadian rhythms. Most chapters of the first edition have been kept (revised, updated, and expanded), and six new chapters have been incorporated to expand the scope of the volume. The book is specially conceived for pharmacy, medical, and professional nursery students as well as for students and professionals from other fields circumstantially working in relation to pharmaceuticals, including chemists, physicists, bioengineers, biochemists, molecular biologists, nurses, and professionals from the field of materials science. Accordingly, it abounds in didactical features, including highlights, recommendations, case studies, and questions. This volume congregates the knowledge and views of experts from the USA, Europe, and Latin America. We are thankful to the authors and the editorial team at Springer, and we hope that this would be a pleasant and renewed experience for the reader. La Plata, Argentina 

Alan Talevi Pablo A. Quiroga

v

Contents

Part I The Basics of ADME Processes 1

 Introduction: Biopharmaceutics and Pharmacokinetics���������������������    3 Alan Talevi and Pablo A. Quiroga

2

Drug Release��������������������������������������������������������������������������������������������   13 Alan Talevi and María E. Ruiz

3

Drug Absorption��������������������������������������������������������������������������������������   25 Alan Talevi and Carolina Leticia Bellera

4

Drug Distribution������������������������������������������������������������������������������������   55 Alan Talevi and Carolina Leticia Bellera

5

Drug Metabolism ������������������������������������������������������������������������������������   81 Alan Talevi and Carolina Leticia Bellera

6

Drug Excretion ����������������������������������������������������������������������������������������  111 Alan Talevi and Carolina Leticia Bellera

7

Routes of Drug Administration��������������������������������������������������������������  129 María E. Ruiz and Sebastián Scioli-Montoto

8

Compartmental Pharmacokinetic Models��������������������������������������������  173 Alan Talevi and Carolina Leticia Bellera

9

Regulatory Requirements and Applications of Physiologically Based Pharmacokinetic Models������������������������������  193 Marina Cuquerella-Gilabert, Matilde Merino-Sanjuán, Alfredo García-­Arieta, Victor Mangas-Sanjuán, and Javier Reig-López

10 Drug-Drug  and Food-Drug Interactions of Pharmacokinetic Nature ��������������������������������������������������������������������  221 Pietro Fagiolino, Marta Vázquez, Manuel Ibarra, Cecilia Maldonado, and Rosa Eiraldi

vii

viii

Contents

Part II Specialized Topics 11 Nanocarriers: Delivery Routes ��������������������������������������������������������������  253 Andrey Silva Morawski, José Adão Carvalho Nascimento Júnior, Mairim Russo Serafini, and Luiza Abrahão Frank 12 Advanced  Techniques for Quality Assessment of Nanocarriers����������  271 Jordano Cichelero Facchini, Anamaria Mendonça Santos, Eduarda Cristina Jacobus Ferreira, Mairim Russo Serafini, and Luiza Abrahão Frank 13 Nanomedicines  Obtained by 3D Printing����������������������������������������������  285 Nadine Lysyk Funk, Júlia Leão, Juliana dos Santos, João Vitor Raupp de Oliveira, Diego Fontana de Andrade, and Ruy Carlos Ruver Beck 14 Absorption,  Distribution, Metabolism and Excretion of Biopharmaceutical Drug Products����������������������������������������������������  309 Molly Graveno and Robert E. Stratford Jr. 15 In  Vitro and In Silico ADME Prediction������������������������������������������������  337 Angela Effinger, Caitriona M. O′Driscoll, Mark McAllister, and Nikoletta Fotaki 16 Relationship  Between Pharmacokinetics and Pharmacogenomics, and Its Impact on Drug Choice and Dose Regimens����������������������������  367 Matías F. Martínez and Luis A. Quiñones 17 The  Relationship Between Pharmacogenomics and Pharmacokinetics and Its Impact on Drug Choice and Dosing Regimens in Pediatrics��������������������������������������������������������  401 Jessica K. Roberts, Leslie Escobar, and Catherine M. Sherwin 18 Bioavailability and Bioequivalence��������������������������������������������������������  423 Z. Gulsen Oner and James E. Polli 19 Drug Transporters ����������������������������������������������������������������������������������  443 Alan Talevi and Carolina Leticia Bellera 20 Blood  Flow Distribution and Membrane Transporters as Determinant Factors of Tissue Drug Concentration������������������������  459 Pietro Fagiolino, Alan Talevi, Marta Vázquez, and Manuel Ibarra Index������������������������������������������������������������������������������������������������������������������  489

Part I

The Basics of ADME Processes

Chapter 1

Introduction: Biopharmaceutics and Pharmacokinetics Alan Talevi and Pablo A. Quiroga

1.1 Presentation The terms Biopharmaceutics and Pharmacokinetics describe relatively young pharmaceutical sciences, consolidating in the second half of the twentieth century (Skelly 2010; Wagner 1981). Both have attracted increasing interest in recent decades, as the medical and pharmaceutical communities have  recognized their effective and potential contributions to design rational dosing recommendations and exploit our therapeutic arsenal in a better manner (Hochhaus et al. 2000). The significance of these areas of knowledge has been further enhanced by growing awareness of the direct relationship between the drug levels reached in different body compartments and the corresponding pharmacological effects, and by the opportunity to predict the in  vivo performance of a drug product (typically reflected by plasma drug concentrations or rate and amount of drug absorbed) from in vitro performance (Emami 2006). In fact, retrospective analyses of the causes of failure of drug candidates at clinical trials in leading pharmaceutical companies have revealed that in about one third of the cases, lack of efficacy is correlated with inadequate drug exposure (Morgan et  al. 2012; Cook et  al. 2014). Furthermore, last A. Talevi (*) Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina e-mail: [email protected] P. A. Quiroga Quality Control of Pharmaceutical Products, Pharmaceutical Toxicology, Faculty of Extract Sciences, University of La Plata (UNLP), La Plata, Argentina Laboratorios Bagó S.A., La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_1

3

4

A. Talevi and P. A. Quiroga

generation pharmaceutical carriers (e.g., pharmaceutical nanocarriers) have recently been introduced which, for the first time in history, could directly impact on the fate of a drug within the body, providing specific (or targeted) distribution and modifying elimination kinetics (Talevi and Castro 2017). These innovative drug products defy some well-established biopharmaceutical and pharmacokinetic concepts, which should be adapted or expanded to encompass novel pharmaceutical technologies. The propagation of complex drug delivery systems explains why today, more than ever, it is necessary for a wider audience to gain at least basic biopharmaceutical and pharmacokinetic knowledge, in order to understand the behavior of such systems, assist in their design, and optimize their use. What do we mean by Biopharmaceutics and Pharmacokinetics? Definition Biopharmaceutics Barbour and Lipper (2008) noted that, by introducing the prefix bio to the word pharmaceutics, it follows that Biopharmaceutics studies the interdependence of biological aspects of the living organism (i.e., the patient) and the physical/chemical principles that govern the preparation and behavior of the medicinal agent and/or the drug product (Barbour and Lipper 2008). The discipline examines the interrelationship of the physical/chemical properties of the drug, the dosage form in which the drug is delivered, and the route of administration, on the rate and extent of systemic drug absorption (Shargel et al. 2012). Pharmacokinetics Pharmacokinetics is a branch of pharmacology that studies, both mathematically and descriptively, how the body affects a drug after administration through absorption, distribution, metabolism, and excretion. It has been observed, in a colloquial manner, that pharmacokinetics deals with “what a biosystem does to a compound’, that is, with everything that happens to drug molecules within the body aside from pharmacodynamic events (Testa and Krämer 2006). In his seminal work, Torsten Teorell, who is often regarded as the father of modern pharmacokinetics, observed that at that time, clinicians were often interested in some aspects of drug action (e.g., the specific mechanism of action of a drug), whereas little attention was paid to kinetics, i.e., the time relation of drug action. Accordingly, he decided to derive “general mathematical relations from which it is possible, at least for practical purposes, to describe the kinetics of distribution of substances in the body” (Paalzow 1995). This type of mathematical equations lies at the core of pharmacokinetics.

1  Introduction: Biopharmaceutics and Pharmacokinetics

5

Some authors have assimilated both Biopharmaceutics and Pharmacokinetics into a single discipline. For conventional drug delivery systems, it is possible to consider both disciplines separately (with the focus of Biopharmaceutics on drug release from dosage forms, and the focus of Pharmacokinetics on those processes that take place from absorption onwards), although the frontiers between both disciplines may become blurry when considering last-generation drug delivery systems, for which drug release could occur during or after drug distribution. Conceived either as separate disciplines or as one, Biopharmaceutics and Pharmacokinetics jointly study what we succinctly call LADME processes, an acronym that refers to the processes of Liberation (i.e., drug release from the dosage form), Absorption, Distribution, Metabolism and Excretion. Briefly, absorption refers to drug movement from the absorption site (which often coincides with the site of administration) to systemic circulation, distribution represents the (reversible) transfer of drug molecules to the extravascular compartments, metabolism represents drug elimination due to biotransformation of drug molecules through enzyme-catalyzed chemical reactions and excretion denotes the physical removal of the drug from the body. The ADME processes are studied in separate chapters of this volume. Metabolism and excretion can be jointly regarded as elimination, whereas distribution, metabolism, and excretion are sometimes collectively referred to as a drug disposition. It is worth differentiating the epistemic object of Biopharmaceutics from biopharmaceuticals. Biopharmaceuticals (also known as biological medical products, biodrugs, or biologicals) are pharmaceutical drug products manufactured in, extracted from, or semi-synthesized from biological sources, including vaccines, blood, blood components, gene therapies, tissues, recombinant therapeutic proteins and living cells used in cell therapy. They can be composed of sugars, proteins, nucleic acids, complex combinations of these substances, or living cells or tissues. Biologicals, like any other drug product, may be studied by Biopharmaceutics and Pharmacokinetics (in fact, a whole chapter of this volume will focus on the kinetics of this type of medication). The procedures used to obtain and characterize biologicals fall within the realms of Biotechnology and Molecular Biology.

1.2 Some Practical Definitions: The Notion of Bioavailability For the best understanding of this and subsequent chapters by those readers who are not familiarized with pharmacology and pharmaceutical glossary, we will next provide some general definitions of terms which would be frequently used throughout this volume.

6

A. Talevi and P. A. Quiroga

Definition Active Pharmaceutical Ingredient (API) A substance used in a pharmaceutical product to furnish pharmacological activity, or to have a direct effect in the diagnosis, cure, mitigation, treatment, or prevention of disease (Code of Federal Regulations Title 21, U.S.  Food and Drug Administration). In other words, API refers to the biologically active component of a drug product, that is, the component directly responsible for a diagnostic or the pharmacological response. It is also commonly referred as drug or drug substance. Some drug products contain more than one API. Drug products Through appropriate manufacturing processes, APIs are combined with pharmaceutically inactive pharmaceutical ingredients called excipients, which comprise pharmaceutical vehicles, carriers, or drug delivery systems. It is closely related to the dosage form concept (physical form in which a drug product is produced and dispensed, e.g., tablets, capsules, syrups, etc.). Excipients may contribute to different aspects of the drug products, such as enhancing physical/chemical or microbial stability, improving organoleptic properties, bulking up solid formulations that contain potent active ingredients in small amounts, and enhancing drug dissolution or absorption. The pharmaceutical carrier does not possess intrinsic pharmacological activity, although its composition and manufacturing process impact drug release and absorption, and thus, drug pharmacokinetics. Some last-­ generation pharmaceutical carriers may also influence the distribution and elimination of drugs. Biophase Drug effect site. Physical region (environment) in which the pharmacological target is located. Systemic treatment or systemic medication A pharmacological treatment in which the therapeutic agent (the drug) reaches its site of action through the bloodstream. We will  consider that a drug molecule has reached systemic circulation once it has left the left ventricle of the heart at least once, that is, once it has reached the aorta. Topical medication Medication that is locally applied to a particular place on or in the body where it is intended to elicit its action. Many topical medications are directly applied onto or into the skin. Topical medications may also be inhalational or applied onto the surface of tissues other than the skin, such as eye drops applied onto the conjunctiva, or ear drops placed in the ear.

1  Introduction: Biopharmaceutics and Pharmacokinetics

Therapeutic window Also known as therapeutic range, it refers to the range of drug concentrations in a bodily fluid (usually plasma) that provides safe and effective therapy. The lower limit of the therapeutic range is called the minimum effective concentration (MEC), whereas the upper limit is called the minimum toxic concentration (MTC). Below the lowest concentration in the window, it is likely that the drug will fail to exert its intended effect. If the drug concentration rises above the therapeutic window, detrimental intensification of the drug’s intended and unintended (off-target) actions will occur. Tabulated limits of the therapeutic window often come from population-based studies. However, to some extent, each individual has a unique therapeutic window for each drug, since there is inter-individual (and also intra-­individual!) variability in drug sensitivity. Drugs with narrow therapeutic windows present small differences between therapeutic and toxic doses. For these drugs, as well as for drugs with marked pharmacokinetic variability, therapeutic drug monitoring (clinical practice of measuring specific drugs at designated intervals) is often performed to guarantee that the drug achieves the desired concentrations in patient fluids (Kang and Lee 2009). Right after initiating drug therapy with such therapeutic agents, the physician may find it useful to measure the plasma drug concentration, tailor the dosage to the individual, and “personalize” the therapeutic intervention. Drug target Or pharmacological target. A biomolecule with which the drug specifically interacts to elicit a therapeutic response. Most drug targets are proteins, although some drugs target other biomolecules (e.g., DNA and RNA). Bioavailability Bioavailability refers to the extent and rate at which a drug reaches its site of action. As the determination of drug levels at the site of action is not feasible in some cases (for instance, the reader may imagine how invasive it would be to measure drug levels in the central nervous system), bioavailability assessment at the site of action is often replaced by measuring systemic bioavailability, that is, the extent and rate at which the drug reaches systemic circulation. This approach is more convenient because drug levels are more frequently quantified in serum or plasma. The plasma drug levels are directly related to those at the site of action. This does not mean that drug levels across different organs will be uniform and identical to those in the plasma it means that the higher the drug levels in the plasma, the higher the drug levels in extravascular tissues. It is important to emphasize that, although we will focus on drug bioavailability, measuring the drug bioavailability of other types of compounds (e.g., toxins, drug metabolites) could occasionally be of interest. Bioavailability assessment is typically performed using drug plasma concentration versus time profiles.

7

8

A. Talevi and P. A. Quiroga

Considering the previous definition of bioavailability, we may observe that bioavailability has a quantitative component (extent) and a kinetic component (rate). Figure 1.1 shows a hypothetical situation in which the same dose of a given drug has been administered through different routes of administration (intravenous, intramuscular, and oral). Even if, as in the example, the quantitative aspect of bioavailability is the same (as reflected in identical areas under the curve for the four plasma vs. time profiles), the kinetic aspect of bioavailability will significantly vary for different administration routes (as reflected by the time at which the drug reaches its maximum plasma concentration). The moment at which the drug levels surpass the MEC and the time period during which the drug levels remain within the therapeutic window will differ on each occasion.

Fig. 1.1  Plasma concentration-time profiles for different routes of administration: intravenous (IV), intramuscular (IM), oral. Assume that the area under the curve is the same for the three profiles, thus indicating that the total amount of drug that has been absorbed is the same in all cases. The kinetic aspect of bioavailability differs in all three cases. The IV administration implies immediate systemic bioavailability of the entire administered dose; in the other extreme, oral administration implies the slowest absorption and thus the lowest bioavailability

1  Introduction: Biopharmaceutics and Pharmacokinetics

9

1.3 The Relationship Between Biopharmaceutics and Pharmacokinetics and the Pharmacological Response Generally, a pharmacologic response occurs when drug molecules interact in a specific manner with copies of the drug target. Whereas it is common to speak of the drug target as a receptor, not all drug targets correspond to the receptor category. The usually specific recognition event between the drug and the drug target has been classically explained through the key and lock analogy (although now we know that: a) neither the drug nor the target are rigid entities, and b) considering a systems biology perspective, drug responses can hardly be explained by a single and punctual interaction event). In very general terms (more detailed and comprehensive information can be found in a pharmacology textbook), drugs often work by blocking the interaction of a drug target with an endogenous ligand or by inducing a conformational change in the drug target that results in a pharmacologic response. From previous comments, it follows that the magnitude of the pharmacological effect essentially depends on two factors. On the one hand, the number of drug molecules that, at a given moment, are interacting (bound) with the correspondent target copies. On the other hand, how favorable is (thermodynamically speaking) such an interaction. The higher the affinity of the drug for its target, the better the intrinsic potency of the drug. Note that a maximal response is expected if, at a given moment, all the available copies of the target are occupied by drug molecules (saturated system). In addition, living systems may resort to different strategies to terminate the action of the drug (e.g., inactivating the drug target). It is interesting to highlight that the encounter between a drug molecule and a target molecule is a probabilistic event depending on the collision probability between the partners of the drug-target complex. Thus, it depends on the number of drug molecules neighboring the molecular target, and the number of copies of the drug target neighboring the drug molecules. It is important to meditate on the statements in italics. If we think about this, it explains the relevance of Biopharmaceutics and Pharmacokinetics. The intensity of the pharmacological response not only depends on the intrinsic potency of the drug, but also on how many drug molecules occupy, at a given time point, drug target molecules. In turn, this directly depends on the number of drug molecules available to the target molecules at the site of action. Even if a drug has a high intrinsic potency, if it does not have access to the biophase in sufficient amount to occupy a pharmacologically relevant proportion of the drug target copies, there will not be pharmacological response.

10

A. Talevi and P. A. Quiroga

1.4 The Free Drug Hypothesis The free drug hypothesis (also called “free drug theory” or “free drug principle” by many authors) offers a conceptual framework to formalize the previous discussion and for the understanding, in subsequent chapters, of many issues related to drug distribution and drug elimination mechanisms. It has been widely applied in drug discovery and development programs to establish pharmacokinetic–pharmacodynamic relationships and determine therapeutically relevant doses that will be used in preclinical and clinical stages, and provides a rationale to monitor the free drug levels of certain drugs under certain clinical scenarios (Dasgupta 2002¸Dasgupta 2008). This hypothesis can be summarized by two prepositions (Smith et al. 2010): (a) The free drug concentrations on both sides of a biological barrier will be the same if the distribution pseudo-equilibrium (steady state) has been achieved. (b) The free drug is the molecular  species that exerts pharmacological activity (only the collision of the free drug molecules with the target is likely to contribute to binding). The drug concentration in the biophase determines the magnitude of the pharmacological response (at least below the saturating concentrations). Numerous studies have demonstrated that the average free drug concentration that is present in vivo at the mean efficacious dose is in good agreement with the in vitro potency (see, for instance: Troke et al. 1990 and Yamada et al. 2007). However, what does free drug mean? in physiological media, some drug molecules will bind reversibly to components in plasma (primarily proteins) and tissues, whereas others will be free (that is, unbound, interacting with no other thing that solvent molecules). Free drug molecules diffuse across biological barriers much more rapidly than their bound counterparts do. There are many exceptions to the free-drug hypothesis (Trainor 2007; Smith et al. 2010). For instance, if a drug has low passive permeability, diffusion into a cell or deep compartment is slow relative to the changes in plasma concentration, and equal concentrations in vascular and extravascular compartments will be hardly achieved. A major category of exceptions to this hypothesis involves drugs that are substrates of drug transporters. Exceptions to the second part of the hypothesis include drugs whose mode of action involves multiple mechanisms or the activation of target-mediated events, and drugs that bind to their targets with a very slow rate of dissociation from the complex, which will sustain receptor occupancy and exert pharmacological effects long after free drug levels have dropped off.

1  Introduction: Biopharmaceutics and Pharmacokinetics

11

References Barbour NP, Lipper RA (2008) Introduction to biopharmaceutics and its role in drug development. In: Krishna R, Yu L (eds) Biopharmaceutics applications in drug development. Springer, Boston, pp 1–25 Cook D, Brown D, Alexander R et al (2014) Lessons learned from the fate of AstraZeneca’s drug pipeline: a five-dimensional framework. Nat Rev Drug Discov 13:419–431 Dasgupta A (2002) Clinical utility of free drug monitoring. Clin Chem Lab Med 40:986–989 Dasgupta A (2008) Monitoring free drug concentration. In: Dasgupta A (ed) Handbook of drug monitoring methods. Humana Press, Totowa NJ, pp 41–65 Emami J (2006) In vitro – in vivo correlation: from theory to applications. J Pharm Sci 9:169–189 Hochhaus G, Barrett JS, Derendorf H (2000) Evolution of pharmacokinetics and pharmacokinetic/ dynamic correlations during the 20th century. J Clin Pharmacol 40:908–917 Kang JS, Lee MH (2009) Overview of therapeutic drug monitoring. Korean J Intern Med 24:1–10 Morgan Piet H, Van Der Graaf J, Arrowsmith DE et  al (2012) Can the flow of medicines be improved? Fundamental pharmacokinetic and pharmacological principles toward improving Phase II survival. Drug Discov Today 17:419-424 Paalzow LK (1995) Torsten Teorell, the father of pharmacokinetics. Upsala J Med Sci 100:41–46 Shargel L, Wu-Pong S, Yu ABC (2012) Applied Biopharmaceutics & Pharmacokinetics. MacGraw-­ Hill, New York Skelly JP (2010) A history of biopharmaceutics in the Food and Drug Administration 1968-1993. AAPS J 12:44–50 Smith DA, Di L, Kerns EH (2010) The effect of plasma protein binding on in vivo efficacy: misconceptions in drug discovery. Nat Rev Drug Discov 9:929–939 Talevi A, Castro GR (2017) Targeted therapies. Mini Rev Med Chem 17:186–187 Testa B, Krämer SD (2006) The biochemistry of drug metabolism--an introduction: part 1. Principles and overview. Chem Biodivers 3:1053–1101 Trainor GL (2007) The importance of plasma protein binding in drug discovery. Expert Opin Drug Discov 2:51–64 Troke PF, Andrews EJ, Pye GW, Richardson K (1990) Fluconazole and other azoles: translation of in vitro activity to in vivo and clinical efficacy. Rev Infect Dis 12:S276–S280 Wagner JG (1981) History of pharmacokinetics. Pharmacol Ther 12:537–562 Yamada S, Kato Y, Okura T et  al (2007) Prediction of alpha 1-adrenoceptor occupancy in the human prostate from plasma concentrations of silodosin, tamsulosin and terazosin to treat urinary obstruction in benign prostatic hyperplasia. Biol Pharm Bull 30:1237–1241

Further Reading This introductory chapter has discussed some pharmacological concepts in a very brief and general manner. The reader is directed to pharmacology textbooks for a deeper insight into pharmacology basis. Chapter 4 of the volume Pharmacology: Principles and Practice (by Hacker et al., Elsevier, 2009) would be an excellent place to find a broader explanation on ligand binding and tissue response

Chapter 2

Drug Release Alan Talevi and María E. Ruiz

2.1 Introduction Definition Drug release refers to the process by which drug molecules are transferred from their initial position in a drug delivery system to the outer surface, and in turn, as solutes, into the release medium (Fu and Kao 2010). “Drug release” is often used interchangeably with drug dissolution, although drug dissolution represents only one step in the more complex process of drug release (Siepmann and Siepmann 2013). Dissolution from drug delivery systems occurs only when the drug is initially in the solid state, as in a tablet or ointment, or when it deliberately precipitates at the site of administration, as in subcutaneous or intramuscular depot formulations, such as insulin glargine (Lindauer and Becker 2019). Drug release is a prerequisite of drug absorption. Drug molecules should be dissolved at the absorption site for their absorption to occur.

Drug release can be a complex phenomenon that depends on the delivery system and may comprise one or more of the following processes: diffusion, partitioning, swelling, osmosis, erosion, disintegration, and/or degradation of the delivery system (Fu and Kao 2010; Bruschi 2015). A. Talevi Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina M. E. Ruiz (*) Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina Quality Control of Pharmaceutical Products /Laboratory of Bioactive Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_2

13

14

A. Talevi and M. E. Ruiz

Diffusion involves the mass transfer of drug molecules associated with random molecular motions down a concentration gradient: the concentration difference is diminished over time by the spontaneous flux of matter. Partitioning refers to the mass transfer of a solute across immiscible or partially miscible phases of different polarities. For instance, portioning between an oil vehicle and aqueous tissue fluids governs the drug release kinetics in intramuscular long-acting lipophilic solutions, such as injectable formulations of estradiol valerate, haloperidol decanoate, and testosterone cypionate (Weng Larsen and Larsen 2009). During swelling, hydrophilic constituents of the release system (usually natural or synthetic polymers) swell upon interaction with water. As the solvent or penetrant enters the polymer, the polymer expands, resulting in greater spaces between the polymeric chains, which favors the dissolution and/or migration of the drug cargo. Drug release from hydrogel-based delivery systems is governed by this mechanism (Kim et al. 1992). To some extent, swellable systems behave similarly to osmotic-controlled systems, in which the delivery system is surrounded by a semipermeable outer membrane with one or more small laser-drilled holes. As water is absorbed through a semipermeable membrane, the resulting osmotic pressure pushes out the active pharmaceutical ingredient through the holes (Almoshari 2022). In both cases, the entry of water triggers the release of the drug. Erodible drug delivery systems also have common points with swellable systems, as they are mostly based on water-swellable polymers, but the release of the loaded drug is not only deferred through progressive swelling but also because of the chemical or biochemical erosion or degradation of the polymer (Maroni et al. 2016). In certain systems (e.g., immediate release tablets), the rate at which drug molecules dissolve from the solid dosage form may determine the release rate from the therapeutic system. Dissolution refers to the mixing of two phases and the formation of a new homogeneous phase (solution). In the context of pharmaceuticals, the term drug dissolution generally refers to the process in which a substance in the solid state forms a solution. Dissolution of the active ingredient (s) from tablets is often facilitated by disintegration, either due to swelling of polymeric substances (e.g., starch) or the use of effervescent excipients, such as in the case of orally disintegrating or orally dissolving tablets (Markl and Zeitler 2017; Ghourichay et al. 2021). Drug delivery systems are classified based on their release kinetics. In principle, one can differentiate between immediate- and modified-release delivery systems. Classical modified-release systems, in turn, include delayed-release (characterized by a lag time before release starts, e.g., enteric-coated formulations and targeted-­ release systems), and extended-release (designed to deliver the drug for prolonged periods). More sophisticated patterns of drug release are also possible, as observed in pulsatile-release delivery systems, in which the rapid and transient release of drugs within short periods is preceded and followed by predetermined off-release periods (Jain et al. 2011), or in bimodal release systems that alternate fast and slow release (Streubel et al. 2000). The kinetics of drug release may follow zero- or first-order kinetics, or intermediate behaviors.

2  Drug Release

15

2.2 Zero-Order Drug Release Zero-order delivery systems release the drug at a constant rate (i.e., the release rate is independent of the concentration or amount of drug that remains in the delivery system). For systemically absorbed drugs, if the drug is released at a constant rate during enough time, as drug accumulates in the body, the rate of absorption (which will be constant) will be equaled by the rate of drug elimination (which is most frequently a first-order process); thus, practically invariable drug concentrations will be observed in the biological media. If the obtained constant drug levels are within the therapeutic range, improved efficacy and safety will be observed. No sub- or supratherapeutic levels of the drug will be observed once the steady state has been reached, because no fluctuations will occur as long as the release rate holds, and the elimination rate remains unmodified. Mathematically, zero-order release is described by the following equation (Dash et al. 2010):

Mt  M0  K0  t

(2.1)

where Mt represents the accumulated mass of drug released over time t, M0 is the amount of drug released at t  =  0 (which is typically 0) and K0 is the zero-order release constant, which depends on the features of the delivery systems (i.e., composition, architecture, geometry) and is expressed in units of mass/time. It follows that, for zero-order delivery systems, a graph of Mt vs. time will be a straight line (Fig. 2.1). Among the drug delivery systems that provide zero-order release over a reasonably long time we may mention osmotic systems, actuated pumps, multilayered tablets, membrane-controlled dosage forms, and vaginal rings (Keraliya et al. 2012; Yadav et al. 2013; Laracuente et al. 2020; Boyd et al. 2020). Fig. 2.1 Graphical representation of zero-­ order kinetics. The slope of the straight line corresponds to the zero-order constant of drug release K0. It has been assumed that the accumulated mass of released drug at t = 0 is zero

16

A. Talevi and M. E. Ruiz

If we consider the mass balance for a drug released into systemic circulation at constant rate K0, assuming that the drug is eliminated from plasma following first-­ order kinetics:



dA  K0  k  A dt

(2.2)

where k is the first-order elimination rate constant, expressed in units of time−1, and A is the amount of drug in the vascular compartment. At early time points (immediately after drug release into the bloodstream began) the negative elimination term in the mass balance is small compared to the absorption term, since elimination follows first-order kinetics and the amount of drug in the body compartment (A) is small. Therefore, the positive term in the mass balance surpasses the negative term, and dA/dt is positive. Consequently, A (and the plasma concentration Cp) initially increase. Eventually, as the drug accumulates in the body, both terms become equal in magnitude and cancel each other out. dA/dt will become zero and no further change in the drug levels will be observed as long as the release rate K0 remains constant. In other words, a steady state with no fluctuation at all is achieved (Fig. 2.2). The situation is mathematically identical to the intravenous infusion that is discussed in more detail in Chap. 8. Integration of Eq. 2.2 leads to: K A  0  1  e  k t k





(2.3)

Example Intravaginal rings are small soft polymeric drug delivery devices designed to provide controlled drug release for intravaginal administration over extended periods. They can be easily inserted and removed by patients for treatment interruption. They are currently used for several therapeutic indications, including hormone replacement, contraception, and luteal supplementation in the context of assisted fertility programs. Some active pharmaceutical ingredients incorporated into these devices include estradiol, estradiol-acetate, etonogestrel, ethinylestradiol, progesterone, and segesterone acetate. Some advantages of these systems include sustained drug release, avoidance of first-­ pass metabolism, and a large drug-loading capacity. Zero-order release is achieved over a significant fraction of the useful life because the drug molecules follow a uniform diffusion pathway length within the torus. Some issues that have been reported include adverse reactions associated with initial burst re-lease in some devices (a problem that may be mitigated through membrane coating) and involuntary ring expulsions that compromise treatment adherence (Boyd et al. 2020).

17

2  Drug Release Fig. 2.2  Concentration vs. time profile for a zero-­ order drug release delivery system that delivers the drug in the vascular compartment at constant rate K0. Eventually, a steady state is reached

2.3 First-Order Drug Release First-order release kinetics imply that the instant rate of release of the drug from the delivery system will be proportional, for any value of t, to the amount of drug remaining in the device, that we will denote as Qt:



dQt  k1  Qt dt

(2.4)

where k1 is the first-order rate constant of drug release (which could be time-­ independent or vary as the length of the diffusion path or the surface area of the device change over time. If k1 is time-independent, integration of Eq. 2.4 between 0 and any given time t leads to the following expression:

ln Qt

ln Q0

k1 t



(2.5)

which after some minimal algebraic operations can be rewritten as a monoexponentially decaying function:

Qt  Q0  e  k1 t

(2.6)

Q0 is the initial amount of drug present in the delivery system. If the entire amount of drug present in the device at t = 0 is released at sufficiently long times (i.e., when t tends to ∞), then Q0 = M∞, and:

M  M t  M  e  k1 t

(2.7)

A. Talevi and M. E. Ruiz

18

Thus:



M t  M  1  e  k1 t



(2.8)

Equation 2.8 shows that the accumulated amount of drug released from the device asymptotically approaches M∞. As expected from first-order release kinetics, the rate of drug release will progressively decrease over time, as the reservoir of drug in the device decreases. It can be easily noted that a graph of Mt versus t will be very similar to Fig.  2.2, with the cumulative amount of released drug gradually approaching M∞. The preceding equations have been satisfactorily used to describe the behavior of diffusion-controlled systems that exhibit Fickian behavior. In particular, if the area across which the release occurs remains constant over time. For example, a non-­ rodable polymeric membrane may mediate drug diffusion between a core loaded with the drug and the release medium (Fig. 2.3). Assuming that the diffusion along the reservoir and the partition between the reservoir and the membrane are fast, diffusion through the membrane becomes the rate-limiting step, and first-order kinetics are observed under sink conditions (at short times after drug release starts, or if the volume of the release medium is sufficiently high or if the drug molecules are being constantly removed from the release medium). In contrast, if the reservoir is saturated (due to the presence of a reservoir of solid drug), the rate of drug flux

Fig. 2.3  Schematic representation of drug release from a reservoir system. The drug reservoir (right) is separated from the release medium (left) but a membrane that acts as the main diffusional barrier. (1) diffusion across the reservoir; (2) partitioning between the core and the membrane; (3) diffusion through the membrane; and (4) partitioning between the membrane and the release medium. The evolution of the amount of drug remaining in the reservoir over time is described by Eq. 2.6, whereas the evolution of the accumulated amount of release drug is governed by Eq. 2.8

2  Drug Release

19

would be constant, and zero-order release is observed until the reservoir is depleted (Bruschi 2015). In monolithic systems, the active ingredient is loaded directly into a matrix that simultaneously acts as storage medium and a mediator of diffusion. If the drug is loaded uniformly into the matrix, it will be released from the matrix under the control of diffusion. For instance, first-order kinetics have been used to satisfactorily describe the release of highly soluble drugs from inert, insoluble, non-swellable porous matrixes (Mulye and Turco 2008). When the drug is coarsely dispersed (i.e., not dissolved), it should be dissolved before being released. When the drug is dissolved, the diffusion takes place at the device surface and the release rate typically decreases over time as the surface is depleted of drug and the mean distance that the drug molecules need to travel before being released keeps increasing. This corresponds to a moving-boundary condition. Solutions to Fick’s equation have been obtained for a broad variety of initial conditions and boundary conditions. That is the case, for example, of Higuchi and the Baker-Lonsdale models (Higuchi 1961; Baker and Lonsdale 1975). The description of the wide diversity of drug release models is out of the scope of the present chapter.

2.4 Intermediate Kinetics: Peppas Model Peppas model of drug release (also known as Ritger-Peppas model, Korsmeyer-­ Peppas model or simply as Power law) was initially conceived to describe drug release from polymeric drug delivery systems through a simple exponential relationship between fractional drug release and the release time:



Mt  K  t n M

(2.9)

where K is sometimes termed the transport constant. The value of the exponent of release n (also called diffusional or transport exponent) provides clues on the possible mechanism(s) of drug release at play. When drug release is exclusively governed by diffusion (Fickian release), n will have a value of 0.5 for flat geometries. In the other extreme, n takes 1 for zero-order kinetics. Intermediate values of n (between 0.5 and 1 for flat geometries) occur when mixed mechanisms of release are present (e.g., diffusion plus polymer relaxation). The limit values of n will be modified for cylindrical or spherical systems (Ritger and Peppas 1987; Peppas and Sahlin 1989). More recently, the model has been successfully applied to describe the release kinetics of other systems, such as liposomes (Wu et al. 2019).

20

A. Talevi and M. E. Ruiz

2.5 Drug Dissolution: The Noyes-Whitney Model and Its Derivatives Most theoretical models describing the dissolution of active pharmaceutical ingredients are based on diffusion equations. The simplest model to describe the dissolution of a chemical compound in a stirred dissolution medium is the Noyes-Whitney model. This model proposes that dissolution involves two steps: (a) the detachment of molecules from the surface of the solid, forming solvated molecules at the solid-­ liquid interface, and (b) the subsequent transport of these solvated molecules from the solid-liquid interface into the bulk solvent under agitation. The Noyes-Whitney model assumes that the first step (detachment) occurs relatively quickly, which implies that the saturation concentration Cs is rapidly achieved at the solid-liquid interface. It follows that the kinetics of drug dissolution are governed by the diffusion of molecules from the interface into the bulk solvent (what we call transport-controlled dissolution). In words by Noyes and Whitney themselves, “the rate at which a solid substance dissolves in its own solution is proportional to the difference between the concentration of that solution and the concentration of the saturated solution, the phenomenon may be considered as a simple process of diffusion.” (Noyes and Whitney 1897) Scenarios where detachment of molecules from the solid surface become the rate-limiting step of drug dissolution (which is termed interface-controlled dissolution) are not of general interest in the field of pharmaceutics, because the agitation intensity required to obtain interface control is too large to be of practical relevance in pharmaceutically relevant situations. If Mt is now the (accumulated) mass of the dissolved compound, then the Noyes-­ Whitney model can be mathematically expressed as:



dM t  w   C s  Ct  dt

(2.10)

where w is the dissolution rate constant, and Ct is the drug concentration in the well-­ stirred bulk solution (which increases over time and tends to equal Cs). Soon later, Nernst and Brunner relied on Fick’s first law to propose a physical meaning of the proportionality factor w. The Nernst-Brunner model considers an unstirred diffusion layer of width h that separates the solid surface from the bulk solution. Beyond this stagnant layer rapid mixing occurs and no concentration gradient is observed. According to them:



dM t D  S   C s  Ct   dt h

(2.10)

where S is the surface area of the solid that is exposed to the dissolution medium, and D is the diffusion coefficient in the dissolution medium.

2  Drug Release

21

Equations 2.9 and 2.10 are accurate for the description of the solution from a flat surface (a system in which only the diffusion in one direction, orthogonal to the solid-liquid interface, is mathematically relevant). They also represent a fairly good approximation for large spherical particles with radii significantly larger than h. Solids with other geometries require more complex mathematical treatments. Nevertheless, the Nernst-Brunner model is sufficient to visualize why increasing the surface area exposed to the solvent (typically, by reducing the particle size) is a common strategy for increasing the dissolution rate of poorly soluble compounds. On the other hand, although the dissolution rate and solubility are very different concepts, there is a direct relationship between them: Eqs. 2.9 and 2.10 shows that the greater the solubility of the drug in the dissolution medium, the faster the dissolution process. In fact, when Ct is small in comparison with Cs (under sink conditions), it can be neglected, and Eq. 2.10 may be approximated as follows: dM t D  S   Cs dt h



(2.11)

and the dissolution rate becomes directly proportional to the solubility of the dissolving drug (and constant if S and h are assumed to be constant). If the sink condition cannot be assumed, the dissolution rate will progressively fall as Ct grows and approaches the saturation condition. However, it is clear that, apart from very specific experimental conditions (flat systems), S cannot be regarded as constant: for non-flat geometries, the area exposed to the solvent progressively decays as dissolution progresses. Moreover, it has been reported that for relatively large particles (with diameters above 50 μm), the effective diffusion boundary layer is constant with respect to the particle size, but for smaller particles h is proportional to the particle diameter (Galli 2006). Hixson and Crowell (1931) recognized this limitation of the diffusion layer model. In their own words: “Except where the solid surface presented to the action of the liquid is plane, there is a constant change in its area, and thereby a slowing down of the rate of dissolution, for it is almost an axiom in chemical philosophy that the rate of chemical action is, ceteris paribus, directly proportional to the surface exposed to that action.” In their famous Cube root law, they accounted for the time dependence of the surface area of the solid during dissolution, arriving to a model with better generalization:

3

M 0  3 M t  zt

(2.12)

Where z is a constant. Hixon and Crowell model still assumes that dissolution occurs orthogonally to the dissolving surface, that the dissolving particles are monodispersed (of equal size) and have spherical geometry that does not change over time, and that the particles remain intact throughout the dissolution (i.e., they do not disintegrate or break into smaller particles).

22

A. Talevi and M. E. Ruiz

References Almoshari Y (2022) Osmotic pump drug delivery systems  – a comprehensive review. Pharmaceuticals (Basel) 15:1430 Baker RW, Lonsdale HK (1975) Controlled release: mechanisms and rates. In: Tanquary AC, Lacey RE (eds) Controlled release of biologically active agents. Plenum Press, New York Boyd P, Merkatz R, Variano B et al (2020) The ins and outs of drug-releasing vaginal rings: a literature review of expulsions and removals. Expert Opin Drug Deliv 17:1519–1540 Bruschi ML (2015) Main mechanisms to control the drug release. In: Strategies to modify the drug release from pharmaceutical systems. Woodhead Publishing, Sawston Dash S, Murthy PN, Nath L et al (2010) Kinetic modeling on drug release from controlled drug delivery systems. Acta Pol Pharm 67:217–223 Fu Y, Kao WJ (2010) Drug release kinetics and transport mechanisms of non-degradable and degradable polymeric delivery systems. Expert Opin Drug Deliv 7:429–444 Galli C (2006) Experimental determination of the diffusion boundary layer width of micron and submicron particles. Int J Pharm 313:114–122 Ghourichay MP, Kiaie SH, Nokhodchi A et  al (2021) Formulation and quality control of orally disintegrating tablets (ODTs): recent advances and perspectives. Biomed Res Int 2021:6618934–6618912 Higuchi T (1961) Rate of release of medicaments from ointment bases containing drugs in suspension. J Pharm Sci 50:874–875 Hixson AW, Crowell JH (1931) Dependence of reaction velocity upon surface and agitation I. Theoretical consideration. Ind Eng Chem 23:923–931 Jain D, Raturi R, Jain V, Bansal P, Singh R (2011) Recent technologies in pulsatile drug delivery systems. Biomatter 1(1):57–65 Keraliya RA, Patel C, Patel P et al (2012) Osmotic drug delivery system as a part of modified release dosage form. ISRN Pharm 2012:528079 Kim SW, Bae YH, Okano T (1992) Hydrogels: swelling, drug loading, and release. Pharm Res 9:283–290 Laracuente ML, Yu MH, McHugh KJ (2020) Zero-order drug delivery: state of the art and future prospects. J Control Release 327:834–856 Lindauer K, Becker R (2019) Insulin depot absorption modeling and pharmacokinetic simulation with insulin glargine 300 U/mL. Int J Clin Pharmacol Ther 57:1–10 Markl D, Zeitler JA (2017) A review of disintegration mechanisms and measurement techniques. Pharm Res 34:890–917 Maroni A, Zema L, Cerea M et  al (2016) Erodible drug delivery systems for time-controlled release into the gastrointestinal tract. J Drug Deliv Sci Technol 32:229–235 Mulye NV, Turco SJ (2008) A simple model based on first order kinetics to explain release of highly water soluble drugs from porous dicalcium phosphate dihydrate matrices. Drug Dev Ind Pharm 21:943–953 Noyes AA, Whitney WR (1897) The rate of solution of solid substances in their own solutions. J Am Chem Soc 19:930–934 Peppas NA, Sahlin JJ (1989) A simple equation for the description of solute release. III. Coupling of diffusion and relaxation. Int J Pharm 57:169–172 Ritger PL, Peppas NA (1987) A simple equation for description of solute release I. Fickian and non-Fickian release from non-swellable devices in the form of slabs, spheres, cylinders or discs. J Control Release 5:23–36 Siepmann J, Siepmann F (2013) Mathematical modeling of drug dissolution. Int J Pharm 453:12–24 Streubel A, Siepmann J, Peppas NA et al (2000) Bimodal drug release achieved with multilayer matrix tablets: transport mechanisms and device design. J Control Release 69:455–468

2  Drug Release

23

Weng Larsen S, Larsen C (2009) Critical factors influencing the in  vivo performance of long-­ acting lipophilic solutions--impact on in vitro release method design. AAPS J 11:762–770 Wu IY, Bala S, Škalko-Basnet N et al (2019) Interpreting non-linear drug diffusion data: utilizing Korsmeyer-Peppas model to study drug release from liposomes. Eur J Pharm Sci 138:105026 Yadav G, Bansal M, Thakur N et al (2013) Multilayer tablets and their drug release kinetic models for oral controlled drug delivery system. Middle-East J Sci Res 16:782–795

Further Reading This chapter is intended as a fast introduction to the topic of drug release. A deeper insight into the subject and a more exhaustive mathematical treatment can be found in the volume Understanding Drug Release and Absorption Mechanisms, by Grassi et al. (CRC Press, 2006). Another good option may be Strategies to Modify the Drug Release from Pharmaceutical Systems, edited by Bruschi and included in the reference list of the chapter

Chapter 3

Drug Absorption Alan Talevi and Carolina Leticia Bellera

Overview Drug absorption involves the mass transfer of a drug from the site of absorption to systemic circulation, and all the processes that the drug molecules undergo during such transfer. The extent and rate of drug absorption depend on several factors, including physicochemical features of the drug, the features of the pharmaceutical dosage form, and anatomical and physiological factors. In many cases, crossing epithelia (e.g., in the case of buccal, oral, or rectal administration) or endothelia (e.g., in the case of intramuscular administration) becomes the rate-limiting step of the entire absorption process. For this purpose, drug molecules might exploit the paracellular way or, more commonly, the transcellular way. In the latter, the critical step of drug absorption usually involves crossing cell membranes. Most drug molecules cross the cell membrane by simple diffusion, which can be appropriately modeled using Fick’s first law. Molecules resembling physiological compounds may also be absorbed by carrier-mediated transport which, to some extent, can be modeled by the Michaelis-Menten equation.

A. Talevi (*) · C. L. Bellera Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_3

25

26

A. Talevi and C. L. Bellera

3.1 Introduction Definition Absorption is a mass transfer process that involves the movement of unchanged drug molecules from the site of absorption (which often, but not always, coincides with the site of administration) to the bloodstream. We will focus on systemic absorption, that is, the movement of drug molecules into the general circulation. Absorption involves overcoming different biological barriers, but also surviving pre-systemic chemical or biochemical modifications (e.g., hydrolysis in the gastric medium or first-pass metabolism). Considering that a drug molecule has been absorbed as soon as it has left the left ventricle of the heart with no chemical modifications (i.e., no covalent bond formation or breakage), any pre-systemic phenomenon experienced by drug molecules can be regarded as constitutive of the absorption process.

Establishing the difference between systemic and topical administration of a drug is likely the first step in explaining the importance of drug absorption. The systemic administration of a medication uses the circulatory system to distribute drug molecules. In contrast, topical medications are applied to a particular location in the body to treat an ailment locally (typically, the skin or mucous membranes, although topical medications also involve inhalational medications, eye drops or ear drops, among others). When systemic administration is used, the whole body (and not only the site of action) is potentially exposed to drug molecules. Conversely, topical administration is intended to circumscribe drug exposure at the affected site of the body and its surroundings. Frequently, adverse reactions to medications are due to the interaction of the drug with elements of the body different from the drug molecular target (i.e., off-target interactions). Adverse reactions to topical medications are often local effects, such as irritation or localized allergic reactions. Systemic administration is more likely to result in more severe adverse reactions than topical administration. Nevertheless, topical administration is not always viable when the pharmacological target of the drug is located in difficult-to-reach organs. Special mention must be made to targeted therapies, where drug molecules are preferably delivered to the site of action through targeted delivery systems (e.g., targeted nanosystems, viral vectors). This concept will be discussed separately in the specific chapters on nanocarriers and drug delivery. Accordingly, this chapter discusses conventional medications only. It follows from the previous discussion that drug absorption is a crucial process when resorting to systemic administration, while it is not required (and often, unwanted) when using topical administration; for drugs applied directly in the

3  Drug Absorption

27

surroundings of their target, such as local anesthetics, absorption often terminates the therapeutic effect (Sultatos 2011).

3.2 Factors Affecting Drug Absorption It should be remembered that drug molecules are not administered in isolation but are incorporated in a pharmaceutical dosage form. However, as discussed in the previous chapter, to overcome biological barriers the drug must be in its free form; that is, it must have been released from the pharmaceutical carrier and in solution (Sultatos 2011). Only then will drug molecules be able to initiate their movement toward the systemic circulation. The previous applies for conventional dosage forms but might not necessarily be true for last-generation pharmaceutical carriers such as nanocarriers. In addition, some dosage forms may provide the drug already in solution (e.g., pharmaceutical solutions). Important When dealing with conventional dosage forms, drug release/drug dissolution are prerequisites for absorption. • Based on the previous concepts, absorption depends on: • Physicochemical properties of the drug, fundamentally: its aqueous solubility, its ionization constant(s), its diffusivity across and affinity for biological barriers that it shall encounter before arriving at general circulation (which, in turn, depends mostly on molecular size and lipophilicity) and its affinity to biological systems such as enzymes or transporters that it may come across during absorption. • Pharmaceutical dosage form and dosage form composition, which directly influences drug release. Dosage forms that ensure faster release will tend to produce enhanced bioavailability in kinetic terms (higher peak plasma concentration at shorter times after administration); this is particularly true if drug dissolution is the rate-limiting step for absorption to occur. For example, González-Canudas et  al. (2019) recently compared the bioavailability of a test dexketoprofen-­ trometamol oral solution with that of 25 mg dexketoprofen-trometamol tablets. Using a randomized, crossover design, they studied the bioavailability of both dosage forms administered in fasting state to a group of 27 healthy subjects. Both treatments were delivered with 250  mL of water at room temperature. It was found that the mean time to peak effect for the oral solution was of only 0.33 hours, compared to 0.50 hours in the case of the tablet. Moreover, the peak concentration of the drug was increased in about 26% for the solution (Fig. 3.1).

28

A. Talevi and C. L. Bellera

Fig. 3.1 Comparison of bioavailability of a dexketoprofen-trometamol oral solution and dexketoprofen-­trometamol tablets. Reproduced from González-Canudas et  al. (2019) under a Creative Common Attribution License

Occasionally, pharmaceutical carriers might include components capable of modulating drug uptake and/or first-pass metabolism. For instance, it has been reported that some common excipients can modify the expression and/or activity of drug transporters and cytochrome p450 enzymes (Tompkins et  al. 2010; Hodaei et al. 2014; Al-Mohizea et al. 2014). Permeation enhancers (also known as penetration enhancers or percutaneous absorption promoters), on the other hand, are chemical substances that increase drug flux through the skin barrier (Kováčik et al. 2020). A similar concept is being actively investigated to improve transport of poorly absorbed active pharmaceutical ingredients across the intestinal epithelium (Maher et al. 2019). In the case of tablets, particle size, amorphism and polymorphism, granulation process parameters and compression pressure are some among many process variables that can have a substantial impact on drug absorption and bioavailability, as they govern the dissolution rate (Sandri et al. 2014; Censi and Di Martino 2015; Kinoshita et al. 2017; Santos et al. 2010). Some specific drug delivery systems have been designed to modify the rate of delivery of the drug to the absorption site and the rate of drug release and/or dissolution, thus influencing the kinetics of absorption. For instance, gastroretentive

3  Drug Absorption

29

delivery systems have been studied to improve the oral absorption of compounds with narrow absorption windows (Davis 2005). Self-emulsifying drug delivery systems are an excellent choice for improved oral delivery of lipophilic drugs (Gursoy and Benita 2004). Modified-release formulations represent another example of the influence elicited by dosage forms on drug release and, consequently, on drug absorption and drug bioavailability. • Anatomical and physiological features of the site of absorption, including expression levels of drug transporters, pH, anatomical adaptations that may favor absorption, the characteristics of tight junctions between adjacent cells of the absorptive tissue and the blood perfusion at the site of absorption. Considerable physiological changes that impact on drug absorption can be observed during pathological processes as well as with age. For example, changes in gastric function in the elderly are likely to modify the kinetics of absorption. The incidence of achlorhydria is approximately 10–20% among elderly patients, compared to less than 1% in younger subjects, whereas hypochlorhydria may be present in approximately 20% of individuals over the age of 70 years (Gidal 2006). A study conducted in a group of healthy elderly subjects showed that 11% had a median fasting gastric pH greater than 5, and that the median time for postprandial return for pH values was significantly longer in elderly individuals (Russell et al. 1993). These changes may be clinically significant for drugs that require and acidic environment for dissolution, release, or activation, for which bioavailability may be reduced. On the contrary, for drug products that contain acid labile active ingredients, an increased gastric pH may result in enhanced drug absorption. The extent and rate of absorption of a drug administered through the oral route might vary if the medication has been administered before, along with or immediately after a meal in comparison with administration in fasting conditions. Transdermal absorption could be enhanced by mean of different physical stimuli, such as local heat or mechanical friction. Transdermal absorption depends on age and skin type (Singh and Morris 2011). It has been shown, for instance, that transdermal permeation of hydrophilic compounds may depend on skin hydration, which varies with age, and that dug permeation across neonatal skin is superior compared with healthy adult skin, because the skin takes some months after birth to mature and attain a similar barrier capability as those of an adult. Moreover, Afro-Caribbean skin tends to be less permeable because it has a greater number of keratinocyte layers within the stratum corneum and a lower stratum corneum water content compared with other races. These factors should be jointly considered when designing dosage forms or when studying a drug product from a biopharmaceutical viewpoint. The pharmacokinetic processes that a drug undergoes after administration are often described through the LADME scheme: Liberation, Absorption, Distribution, Metabolism and Excretion, which are considered in separate chapters in the first part of this book.

30

A. Talevi and C. L. Bellera

Definition The acronym LADME refers to Liberation, Absorption, Distribution, Metabolism and Excretion. These are pharmacokinetically relevant processes that a drug undergoes after entering the body. Although these processes generally follow such a sequence, they should not be regarded as discrete, non-­ overlapping events. Each drug molecule that enters the body has its own, distinctive trajectory. For instance, some drug molecules may be still at the absorption site while others have already been absorbed and started their distribution or elimination. In other words, one process continues to occur while the next one begins, and on certain occasions (e.g., sustained release delivery systems) all five processes may be occurring simultaneously to different drug molecules. Conventional drug delivery systems can only have a direct impact on the kinetics of drug release and absorption, and only an indirect impact on the remaining processes (by influencing the rate and extent of drug release and absorption). Last-­generation delivery systems, in contrast, may possibly modify all the LADME processes in a direct manner.

3.3 The Fluid Mosaic Model and Beyond Crossing the selective barrier posed by cell membranes is, in general, the most relevant mass-transfer process during absorption. To understand the role of biological membranes and tissue barriers in drug absorption, it is necessary to begin with a basic understanding of cell membranes. While we assume that the reader is familiar with the generalities of the cell membrane structure, we provide an overview to give some context for the study of the mechanisms of drug absorption. Note that the principles governing absorption are similar to those governing distribution and uptake by elimination organs. Accordingly, this and the subsequent sections will be of great value for the understanding of both Chaps. 4 and 5. The fluid mosaic model of the cell membrane was devised by Singer and Nicolson (1972) based on thermodynamic considerations and experimental evidence and proposes that phospholipids are the main constituent of cell membranes. This type of lipids (Fig. 3.2) generally consists of two hydrophobic fatty acids (“tails”) and a hydrophilic phosphate group (“head”). These fatty acids and phosphate group are esterified with a glycerol molecule. Phosphate groups can be modified with simple organic molecules such as choline. Phospholipids provide a mosaic structure predominantly arranged as a fluid bilayer (the matrix of the mosaic), with their phosphate groups in contact with the aqueous phase (either the extracellular medium or the cytosol), while the nonpolar

3  Drug Absorption

CH3

OH

N H3C

P

CH3

O O

O CH3

31

O O O

Hydrophobic tails

O

CH3

Hydrophilic head

Fig. 3.2  General structure of a phospholipid, essential constituent of a cell membrane

fatty acid chains are sequestered away from contact with water and oriented to the inner space of the bilayer. Many integral proteins (or occasionally, glycoproteins or lipoproteins) can be found embedded in the membrane (buried to different degrees in its hydrophobic interior). Integral proteins are classified as transmembrane proteins if they extend from one side of the membrane to the other. Integral proteins are amphipathic, with highly polar groups protruding from the membrane into the aqueous phase and nonpolar groups embedded within the hydrophobic core of the bilayer. Additionally, peripheral proteins are also found, bound to the membrane by weak interactions and not strongly associated with membrane lipids. Membrane proteins present comparatively high amounts of α-helical domains, in contrast with soluble globular proteins, which in general exhibit a smaller fraction of α-helices in their native structure. The arrangement of the cell membranes (Fig. 3.3) acts as a bi-dimensional fluid that, in principle, allows free lateral diffusion of membrane components. Cholesterol plays a significant role in regulating membrane fluidity and conferring structural stability. Singer and Nicolson’s model has been expanded to include important elements of the membrane architecture, some of which had already been envisaged by Singer and Nicolson themselves (Alberts et  al. 2002). Among these elements, we may mention membrane asymmetry, the existence of lipid rafts (specialized areas in membranes where proteins and some lipids, primarily sphingolipids and cholesterol, are concentrated) and flip-flop movements. The cell membrane is permeable to small nonpolar molecules, whereas molecules with high polarity (such as charged molecules), high molecular weight and/or high conformational freedom will have difficulty crossing it. Cells have developed specialized transport mechanisms that utilize membrane proteins, including channels and carriers. Embedded in biological membranes are carriers that facilitate the passage of their substrates through the membranes (uptake transporters) and carriers that facilitate the inverse movement (efflux transporters). Channels and some types of carriers carry substrates down a concentration gradient, without direct or indirect ATP consumption. Other carriers require energy to transport their

32

A. Talevi and C. L. Bellera Extracellular fluid Glycoprotein Glycolipid

Protein channel (transport protein)

Peripherial protein

Globular protein Phospholipid

Integral protein (globular protein)

Cholesterol

Cytoplasma

Fig. 3.3  Fluid mosaic model

substrates. In some very specific cases, molecules interacting with cell surface receptors trigger the formation of endocytic vesicles, which move receptor-bound molecules into the cell (endocytosis) or through the cell (transcytosis). This form of transport is more frequent for large molecules (e.g., some therapeutic proteins) and advanced drug delivery systems (e.g., actively targeted pharmaceutical nanocarriers). The combined contribution of all the elements (lipid bilayer, channels, uptake and efflux carriers, endocytosis) determines the selective permeability of a given cell to a given molecule. Selective permeability is crucial for the regulation the inner cell microenvironment.

3.4 How Can Drugs Be Absorbed? In principle, any substrate being absorbed can potentially use the paracellular and/ or the transcellular transport (Fig. 3.4). Paracellular transport refers to the transfer of substances across an epithelium or endothelium through the intercellular space between adjacent cells. Paracellular transport is exclusively passive and which substances can use this route of transport depends on the tight junction characteristics of the epithelium or endothelium. For instance, the particularly tight junctions at the blood-brain barrier preclude the paracellular route. In contrast, other tissues present “leaky” tight junctions, favoring the paracellular route. When the paracellular route is available, small polar molecules are preferentially transported through it.

3  Drug Absorption

33

Fig. 3.4  Different mechanisms that may be involved in drug movement across epithelia or endothelia

Defnition The tight junctions are protein structures that form a selective barrier in the paracellular space between adjacent epithelial or endothelial cells, limiting movement of ions, solutes, and water. Tight junction strands represent an impermeable barrier, as a “wall”, but a series of permeable aqueous “pores” perforate such wall. Although the structure of tight junctions is complex and dynamic and varies across different epithelia and capillary beds (González-Mariscal et  al. 2003; Bazzoni 2006), evidence points to tight junction-associated claudins as the basis for the selective size, charge, and conductance properties of the paracellular pathway (Van Itallie and Anderson 2004). Tight junction permeability can be significantly affected by pathological stimuli, but it is also actively investigated as a possible way to enhance drug delivery of hydrophilic drugs (González-Mariscal et al. 2017).

Conversely, the transcellular route offers a wide spectrum of possibilities that extend the diversity of the substrates that may be taken up by the cells. These possibilities include simple diffusion (passive and without intervention of any

34

A. Talevi and C. L. Bellera

membrane protein), facilitated diffusion (down the concentration gradient) either through protein channels or carrier proteins, active transport (which may occur even up the concentration gradient, with energy consumption) and transcytosis. Important Passive diffusion is seemingly the most frequent transfer mechanism involved in drug transport through cell membranes. For diffusion, solutes should be sufficiently hydrophobic to partition into the lipid bilayer of the plasma membrane. Since most drugs are either weak acids or bases, the ability of a given drug to partition into the membrane is highly modified by pH conditions.

3.5 Free (or Simple) Diffusion In free diffusion, drug molecules use the transcellular route spontaneously, down the concentration gradient and without the involvement of any membrane protein. As crossing the cell membrane is usually the slowest step of the entire absorption process (thus governing its kinetics), we will focus on the concentration gradient through the cell membrane. The kinetics of free diffusion can be adequately described, in principle, using Fick’s first law. Fick’s first law can be considered as a specific case of Fick’s second law when applied in steady state conditions. For mathematical simplicity, let us imagine the absorption site as a well-defined compartment (which is usually not) where a certain amount Q1 of the dissolved drug is homogeneously dispersed (Fig. 3.5). We will then assume that a negligible movement of molecules occurs parallel to the membrane plane. Thus, the only diffusion process that we focus on is the movement of drug molecules from the site of absorption to the inside of the cells. In other words, we will study the drug flux along a single spatial coordinate (the direction orthogonal to the epithelium or endothelium surface, whatever applies) and this will be called x-direction. As previously said, the kinetics of simple diffusion can be satisfactorily modeled using Fick’s first law:



J  D •

dC dx

(3.1)

In this equation, J is the flux in the x-direction, i.e., the mass flow per unit area, expressed as mass per unit time per unit area (for example, mg/s ּ cm2); dC/dx is the concentration gradient, in concentration unit per distance unit (e.g., (mg/cm3)/cm) and D represents the diffusion coefficient of the solute within the membrane, which is expressed in units of area per time (e.g., cm2/s). D is sometimes referred to as the binary diffusion coefficient, reminding us that it depends on the nature of both the solute and diffusion medium (Friedman 2008a). The diffusivity of molecules in different media, including cell and artificial membranes, is known to be influenced by molecular size and molecular shape of the diffusing solute (Hoheisel 1988; Li et al.

3  Drug Absorption

35

C1 C2  S donor compartment

receiving compartment

Fig. 3.5  A simple model to describe drug absorption through free diffusion. The site of absorption corresponds to the left (donor) compartment (which contains drug to be absorbed), whereas the right compartment represents the receiving (acceptor) compartment (which contains absorbed drug)

2009; Leung et al. 2012; Yang and Hinner 2015; Chan et al. 2015). It has also been proposed that permeability could be entropically favored by low molecular flexibility (Veber et al. 2002; Rezai et al. 2006), as the conformational freedom of the solute would be reduced within the membrane compared to the aqueous medium. The consequent loss of entropy would oppose mass transfer across the diffusion barrier. Nevertheless, flexibility can sometimes contribute to absorption if it allows the formation of intramolecular hydrogen bonds, tuning the polar surface area, as has been proposed to explain the oral bioavailability of cyclosporine A (a cyclic peptide of high molecular weight) (Wang et al. 2018) and piracetam (a very polar small molecule) (Ribeiro et al. 2017). The negative sign in Eq. (3.1) accounts for the fact that the flux is positive if the concentration gradient is negative which indicates that the net flux occurs from higher to lower concentrations. As the drug is absorbed, the amount of drug at the site of absorption decreases. If we assume a linear change in concentration across the membrane (which occurs at steady-state, that is, if the flux is the same at every cross-section in the membrane and constant in time) (Friedman, 2008a), Eq. 3.1 can be integrated into:



J   D·

C 

(3.2)

where δ denotes the thickness of the membrane and ΔC is the difference between the solute concentrations in the lipid phase at the interface between the membrane and the surrounding aqueous environment. If Clipid,1 denotes the concentration of the drug in the side of the membrane in contact with the source or donor compartment, and Clipid,2 represents the drug concentration on the side of the membrane facing the receiving compartment, Eq. 3.2 can be rewritten as:

36



A. Talevi and C. L. Bellera

J   D·

C

lipid , 2

 Clipid ,1







(3.3)

However, solute concentrations at the lipid phase are relatively difficult to measure. Therefore, they can be replaced by the corresponding aqueous concentrations in the donor and receiving compartments (C1 and C2, respectively) considering that they are related by means of the partition coefficient K between the lipid phase of the membrane and the aqueous environment. Provided that the partition equilibrium has been reached:



J   D·K ·

C2  C1 

(3.4)

This explains why the logarithm of the relative concentration of a compound when it partitions between a two-phase system, usually water and octanol, is usually used as crude predictor of ability of a that compound to permeate through the cell membrane by free diffusion, as octanol is assumed to have a lipophilicity comparable to that of a cell membrane (Seddon et al. 2009). A lipophilic drug will display a high affinity to the membrane (high K) and this is associated with high rate of absorption; however, an extremely lipophilic drug will possibly tend to remain “trapped” in the membrane, a phenomenon which is termed membrane retention (Lee and Choi 2019; Soriano-Meseguer et al. 2020). Beside this issue, bear in mind that dissolution is a precondition for absorption and that drugs with low solubility will tend to present dissolution problems. Especially when designing new drugs intended for oral administration (which are usually administered using solid dosage forms) an adequate hydrophilic-lipophilic balance is pursued (Dahan et al., 2016). Noteworthy, D, K and δ can be grouped together into the permeability coefficient P:



P

D·K 

(3.5)

Noteworthy, the notations Papparent or Peffective are usually used when estimating the permeability coefficient experimentally (e.g., by measuring permeability across a cell monolayer such as Caco-2 cells) to underscore their vulnerability to bias due to experimental conditions (Brodin et al. 2009). The flux J can be expressed as:



J=

dQ dt • S

(3.6)

where dQ/dt is the instantaneous rate of diffusion corresponding to the mass of the solute that has diffused during the infinitesimal period dt, and S represents the absorptive surface area. By replacing with Eqs. 3.5 and 3.6 in Eq. 3.4 and writing (C2-C1) as ΔC, we obtain:

3  Drug Absorption

37

dQ   P • S • C dt



(3.7)

Note that the factor ΔC determines whether the preceding equation acquires a negative or positive sign (the remaining factors on the right-hand side of the equation are positive and constant under given pressure and temperature conditions). In addition, we can observe that the instantaneous rate of absorption is directly proportional to the absorptive surface area. This explains the key role of anatomical adaptations that amplify the effective surface area  for increased absorption (or reabsorption) in certain organs. For instance, plicae circulares together with villi and microvilli amplify the mean total mucosal surface of the digestive tract to about 32 m2. This concept has also been exploited for the design of transdermal dosage forms (see the example below).

Example Nicotine replacement therapy is one of the most frequently used treatments for cessation of smoking. It is available in a diverse range of dosage forms, including gum, sublingual tablets, nasal sprays, inhalers, and transdermal patches. Table 3.1 shows the impact of nicotine transdermal patch surface area on nicotine bioavailability (the table includes mean data from six healthy male smokers +/− the standard deviation). Bioavailability is measured using the area under the curve in a plot of drug concentration in blood plasma versus time (AUC) and peak plasma concentration (Cmax). AUC (from zero to infinity) is an indicator of the total amount of drug that has reached general circulation (total drug exposure over time); Cmax reflects both the quantitative and kinetic aspects of bioavailability. tmax represents the time at which Cmax is observed, and it speaks of the kinetic component of bioavailability (the smaller tmax, the faster the absorption kinetics). The data shown in the table has been extracted from Sobue et al. (2006). Note that the bioavailability for a surface area of 40 cm2 roughly doubles that for a surface area of 20 cm2, while tmax increases for 20  cm2, which is consistent with a lower initial rate of absorption. The reader may visit Chaps. 8 and 18 for an extensive discussion of bioavailability. Table 3.1  Comparison of nicotine bioavailability for 20 and 40 cm2 nicotine transdermal patches. The mean +/− s.d. of each parameter is shown Transdermal patch 20 cm2 Transdermal patch 40 cm2

AUC (ng ⋅ h/ml) 155.13 +/− 29.51

Cmax (ng/ml) 9.15 +/− 1.79

tmax (h) 13.3 +/− 3.4

280.96 +/− 31.08

17.25 +/−1.60

8.2 +/− 5.2

38

A. Talevi and C. L. Bellera

Considering that the magnitude of the diffusion rate (i.e., the rate of appearance of the solute at the acceptor side of the membrane) is the same than the rate of disappearance from the donor side (i.e., the site of absorption), then if we call the mass of solute in the donor compartment Q1: dQ1  P·S·C dt



(3.8)

And now, if both sides of the equation are divided by the volume of the donor compartment V: dC1 P·S  ·C dt V



(3.9)

Questions Is the (C2-C1) factor in Eq. (2.4) smaller or greater than zero? Answers Since diffusion is proceeding from the site of absorption (donor compartment), C1 is necessarily greater than C2, and ΔC, i.e., (C2-C1) is smaller than zero. Accordingly, the alluded factor is smaller than zero.

Questions When will the diffusion end? Answers In principle, if the considered system was a closed system, diffusion would end when C2 equals C1. This is what would occur in an in vitro absorption experiment if, after the experiment begins, no mass of the test compound is added to the donor compartment or removed from the acceptor compartment. At that point of dynamic equilibrium (that is, when identical concentrations are found at both sides of the diffusion barrier) the masses of solute in each compartment (Q1 and Q2) might well be very different, because they depend on the volumes of these compartments. If we now think of absorption in vivo and for a moment we consider that the acceptor compartment represents the intravascular fluid, it can be accepted that the volume of the acceptor compartment would usually be much larger than the volume of the donor compartment (the site of absorption). Therefore, in a hypothetical equilibrium condition, this acceptor compartment would accommodate much more solute molecules than the donor compartment, so that both concentrations are equal. As discussed in Sect. 3.5.1, this and other factors will contribute to a more complete and efficient absorption of the administered dose.

3  Drug Absorption

39

Important Depending on the route of administration, the lipid bilayer of the cell membrane of epithelial or endothelial cells is often considered to be the fundamental barrier that a small molecule must overcome to reach the intravascular space. Such an assumption is generally well-founded, although the number of diffusional barriers that the drug must overcome greatly exceeds the bilayer. For instance, if we consider oral absorption, the intestinal epithelium is covered by a mucus layer rich in mucins, with gel-like properties (Johansson et  al. 2013). After crossing the apical membrane, the drug must cross the cytosol, the basolateral membrane, connective tissue, interstitial fluid, and the cell membrane of the endothelial cell of the blood capillary. The drug will then be convectively carried to the heart, finally reaching general circulation. It can be observed that the aforementioned (sub)processes are serial. If any of them is slower than the others, it will be the limiting step and govern the kinetics of the entire absorption process. In general, it is safe to assume that crossing the cell membranes is such rate-limiting step, because the hydrophobic core of the cell membrane is much more viscous than water and the bulk cytoplasm (Bicknese et al. 1993; Adrien et al. 2022).

3.5.1 Sink Condition and Absorption Following the previous discussion, if the intravascular fluid was considered as the acceptor compartment (as drug absorption concludes when the drug molecules reach systemic circulation), it can be admitted that the volume of the acceptor compartment is much larger than the volume of the site of absorption. The diffusion process linked to absorption will conclude once the molecule has reached the capillary bed that irrigates the site of absorption (from there, the blood flow will take it to systemic circulation by convection, removing drug molecules away from the site of absorption). Furthermore, while only free drug concentrations are relevant to establish the concentration gradient, a substantial amount of the drug in the intravascular compartment will generally appear bound to plasma proteins (this is further discussed in Chap. 4). From all these considerations, it is reasonable to assume that C2 will tend to be much smaller than C1 and can be neglected in most in vivo situations. This is regarded as sink condition and allows simplification of Eq. 3.9 to:



dC1 P·S  ·C1 dt V

(3.10)

which corresponds to a first-order kinetics equation. We can now group together P, S and V into a kinetic absorption constant linked to absorption by free diffusion through the membrane, ka,m (O’Shea et al. 2022):

40



A. Talevi and C. L. Bellera

dC1  ka , m  C1 dt

(3.11)

This simple step has very important implications. The resulting expression (3.11) clearly corresponds to a first order kinetics equation: under sink conditions, the instantaneous rate at which the concentration of drug decreases in the site of absorption is at any moment proportional to the remaining concentration of the drug in the site of absorption. The sink condition positively impacts on the extent of absorption since it helps sustaining a concentration gradient (which is the driving force of the diffusion process) and thus favors a complete absorption of the administered dose. There are some scenarios for which the sink condition may not apply, consequently reducing the rate of absorption. Exercise, for instance, redistributes blood flow away from the gastrointestinal tract toward the active muscles and lungs, resulting in a decrease in the splanchnic-hepatic blood flow (splanchnic blood flow is reduced by half during moderate exercise and up to 80% during intensive exercise (Koivisto and Felig 1978)). Although this is rarely of clinical significance, it might negatively affect the kinetics oral absorption, especially for highly permeable drugs (Clausen 1977). An opposite trend may be observed in the case of intramuscular or subcutaneous administration, as drug absorption from exercising tissue may be increased during exercise. The absorption of subcutaneous insulin, for instance, is increased during moderate or intensive leg exercise (Kemmer et al. 1979). Figure 3.6 represents the instantaneous rate of absorption versus the concentration of drug at the site of absorption for a drug absorbed through simple diffusion: absorption by free diffusion can be regarded as a linear process. A similar behavior is expected for a passively absorbed drug when, in the framework of in vitro permeability studies (e.g., permeability studies using Caco-2 or MDCK cell monolayers), the initial rate of absorption is plotted versus the initial concentration of the drug using different concentration levels.

Fig. 3.6  Graph of instantaneous rate of absorption versus the (instantaneous) concentration of the drug in the site of absorption, for a drug being absorbed through a linear process

3  Drug Absorption

41

Important Free diffusion is a linear, nonsaturable transport mechanism. Notably, the concentration gradient across the barrier was assumed to be linear and constant (time-independent) from which Eq. (3.1) could be rewritten as (3.2) and then further simplified by assuming that the drug concentration in the receiver compartment can be neglected versus the concentration in the donor compartment under sink conditions. Is this assumption reasonable in a biopharmaceutical framework? Can the boundary conditions (drug concentrations at the boundaries of the barrier) remain constant in the face of a perpetual flux? No, they cannot (Friedman 2008a). The transport of a solute from the donor compartment to the receiver compartment implies a first-order decrease in the drug concentration in the donor compartment, and a concomitant increase in its concentration in the receiver compartment. Therefore, the concentration gradient cannot be considered constant (Brodin et al. 2009). Nevertheless, if the boundary conditions change slowly compared with the rate at which the transport process can adapt to that change, transport can be regarded as quasi-steady, which has proven an adequate assumption for describing biological transport in the presence of changing boundary conditions (Friedman 2008a). When estimating drug permeability in  vitro using closed systems (where no mass of the test compound is added or removed after the start of the experiment) experiments are performed under approximate sink conditions. The concentration in the receiver compartment is initially zero and is assumed to increase insignificantly during the time course of the experiment. For the sake of precision, in the framework of this type of experiment it is suggested that the concentration gradient should not change by more than 10% during the experiment (Brodin et al., 2009). Moreover, no hydrodynamic or electrical gradient should be present (the concentration gradient of the test molecule should be the only gradient at play), unstirred water layers surrounding the permeability barrier in question should be minimized by stirring, and it should be ensured that the transport of the test compounds depends entirely on free diffusion (verifying the linearity between the flux and the concentration gradient over a wide concentration range).

3.5.2 The Impact of pH on Drug Absorption The previously presented equations are valid for describing the diffusion of noncharged solutes. Studying the diffusion of charged species could be relevant in the field of pharmaceutical sciences, because most drugs are either weak acids or weak bases that are commonly ionized under physiological conditions. However, the mathematical treatment of the diffusion of charged solutes under the influence of an electrical potential gradient is much more complex. Under such circumstances, the driving force for mass transport is a combination of the electrical and concentration gradients (electrochemical potential). The flux equation for electrolytes is the

42

A. Talevi and C. L. Bellera

electrodiffusion or Nernst-Planck equation; its general solutions are in fact so complex that they have rarely been applied to biological systems. Although it is beyond the scope of this chapter to provide mathematical equations that describe the diffusion of ionic species, we next provide a general discussion on how the ionization state of a weak acid or base may affect its absorption by free diffusion. As mentioned, most known drugs are weak acids or bases. For a weak acid (e.g., aspirin, pKa = 3.5), the ionization equilibrium can be written as: HA  aq   H 2 O  l   H 3 O   aq   A   aq 





Following Le Châtelier’s principle, in an alkaline medium that consumes H3O+, the equilibrium shifts to the right, thus increasing the concentration of the charged A−species. Conversely, in acidic media the equilibrium shifts to the left, increasing the concentration of non-charged HA species. Interestingly, while the solubility in an aqueous media is higher for charged species, the permeability of charged species through the cell membrane by passive diffusion is low. Moreover, the electrical potential gradient will also affect the transport of charged species, in which case the Nernst-Planck equation will be useful for determining the flux across the barrier in question. From the previous equilibrium, the corresponding acid ionization constant Ka is given by: H 3 O    A   Ka   HA 



(3.12)

Taking the logarithm of both sides of the equation easily leads to the Henderson-­ Hasselbalch equation: A  log10 Ka  log10 H 3 O    log10   HA 



(3.13)

Then: A  pKa  pH  log10   HA 



(3.14)

Finally:



 A   pH  pKa  log10 HA 

(3.15)

3  Drug Absorption

43

A very similar development can be realized for a weak base: B  aq   H 2 O  l   BH   aq   OH   aq 





Here, the base ionization constant is: OH    BH   Kb    B



(3.16)

This leads to:



 BH   pOH  pKb  log10   B

(3.17)

The reader should bear in mind that pH + pOH equals 14. Although the effects of pH variations on solubility and permeability are relatively straightforward for drugs with a single ionizable group within the physiologically relevant range of pH, the analysis becomes more complicated when addressing drugs with many ionizable functions (Fig. 3.7), where such ionizable groups will gain protons in a serial manner when the pH gradually decreases, and vice versa, or in some cases even zwitterionic species with no net charge can be observed. The former analysis provides the basis for the classic pH Partition Theory. The theory states that because most drugs are weak electrolytes and the non-ionized species are much more likely to freely diffuse across biological barriers, weak bases will preferably distribute to acidic body compartments where the pH is below the pKa of the conjugate acid, and weak acids are preferably distributed to

Fig. 3.7  Logarithm of the solubility vs. pH for compounds with multiple ionizable groups. Reproduced under permission of Elsevier from the European Journal of Pharmaceutical Sciences, Vol 48, Shoghi E, Fuguet E, Bosch E, Ràfols C. Solubility-pH profiles of some acidic, basic, and amphoteric drugs, pages 291–300, Copyright (2013)

44

A. Talevi and C. L. Bellera

compartments whose pH is above their pKa. Note that the total partitioning of ionizable compound between the lipidic and the aqueous media (Eq. 3.4) tends to be low whenever the pH of the aqueous media favors the occurrence of the charged species. The tendency of weak electrolytes to accumulate in compartments where charged species predominate is also called ion trapping. The pH Partition Theory has some limitations (for instance, the permeation of charged species across biological barriers cannot always be neglected); however, it still provides a useful setting to anticipate, for instance, the permeability of orally delivered drugs in different segments of the gastrointestinal tract. This also explains why the distribution coefficient D (which considers the water:octanol partitioning of a drug at a specific pH) or its logarithmic form logD are more appropriate predictors of passive absorption than the partitioning of the non-ionized species only. Note that the same abbreviation is often used for the distribution coefficient and the diffusion coefficient, although they are very different entities. 

Example Clinically significant drug-drug interactions involving antiretroviral medications occur in up to 40% of HIV patients on treatment; these interactions can lead to diminished efficacy and increased risk of drug resistance (Lewis et al. 2016). Acid-reducing agents are commonly used co-medications by HIV-1-­ infected patients receiving antiretroviral treatment, and several of them are available over-the-counter in most countries (van Lunzen et al. 2007). Several reports have demonstrated significant drug-drug interactions between acid-reducing medications and some antiretrovirals. For instance, Tomilo et al. (2006) reported a mean reduction of 94% in atazanavir bioavailability in nine healthy volunteers receiving this antiretroviral drug coadministered with lansoprazole (Table  3.2). The loss of solubility of atazanavir at increased pH values is considered to be responsible for this effect. In some cases, the effect of the interaction between antiretroviral medications and acid-reducing agents can be mitigated by temporal separation of dose administration, whereas some combinations of antiretrovirals and acid-­reducing agents are directly contraindicated (Falcon and Kakuda 2008; Lewis et al. 2016). Table 3.2  Negative effect of lansoprazole coadministration on atazanavir bioavailability Atazanavir administered alone Atazanavir administered after lansoprazole Mean +/− S.D

AUC0–24 (μM ⋅ h) 16.3 +/− 9.0 0.85+/− 1.8

Cmax (μM) 3.2 +/− 1.7 0.13 +/− 0.19

tmax (h) 2.8 +/− 0.75 3.1 +/− 1.0

3  Drug Absorption

45

3.6 Facilitated Diffusion If we pay close attention to Eq. (3.4), it is clear that free diffusion is not favored for certain chemical compounds. This is the case for large and/or highly hydrophilic molecules. Large molecules tend to possess small diffusion coefficients, whereas highly hydophilic molecules tend to have small partition coefficients. How does the cell handle the transport of such molecules (e.g., glucose from the intestinal lumen) through the cell membrane, when required? Facilitated diffusion (also known as passive-mediated transport) is the downhill movement of molecules or ions (from a region of high chemical or electrochemical potential to a region of low chemical or electrochemical potential) across a biological membrane via specific transmembrane integral proteins. There are essentially two types of protein structures that facilitate diffusion. Channel proteins form open pores in the membrane, allowing small molecules of an appropriate size and charge to pass freely through the lipid bilayer. They create paths for diffusion with resistance below that of the bilayer in which they are embedded. On the other hand, carrier proteins bind specific molecules to be transported and then undergo spontaneous conformational changes that allow the molecule to pass through the membrane and be released on the opposite side.

Important In evolutionary terms, transporters that facilitate the absorption of drugs have emerged to mediate the traffic and compartmentalization of physiologic compounds that, owing to their physicochemical features (hydrophilicity and/or size) would otherwise not be able to permeate through lipophilic membranes, including a wide spectrum of nutrients and waste products. Polar amino acids, peptides, nucleosides, sugars, and vitamins are some physiological compounds that use carrier-mediated transport. To be absorbed, drugs that mimic or resemble these physiological compounds can “hijack” transporters. Classical examples of drugs that use uptake carriers include 5-­fluorouracil (anticancer drug), levodopa (antiparkinsonian) and methyldopa (antihypertensive). For example, the resemblance of levodopa and methyldopa to the amino acid tyrosine is shown in Fig. 3.8. Furthermore, amino acids have been increasingly used in prodrug design: esterification of a drug with an amino acid may allow to hijack intestinal amino acid and peptide transporters, as in the case of L-valyl ester prodrugs of poorly permeable antiviral drugs such as valaciclovir and valganciclovir. Carrier-mediated transport is of particular significance for the absorption of some oral medications, as high levels of relevant transporters that facilitate nutrient uptake are found throughout the intestine (more specifically, along the small intestine, the most specialized absorptive region in the human body).

46

A. Talevi and C. L. Bellera

O

O

O HO

HO

HO

Levodopa

OH

OH

OH NH2

CH3 NH2

HO

NH2 HO

Tyrosine

Methyldopa

Fig. 3.8  The chemical similarity between the therapeutic agents levodopa and methyldopa and tyrosine explain why these drugs can use, for their absorption and distribution, carriers that mediate the uptake of amino acids

A very important difference between facilitated diffusion and free diffusion is that, in the case of the former, the maximal number of solute molecules that can be transported across the membrane per unit time depends on the number of available copies of the channel or carrier that are expressed at the permeability barrier. In other words, facilitated diffusion is saturable. Additionally, the temperature dependence of carrier-mediated transport is more pronounced than that of free diffusion, for which the dependence on temperature is mild (carrier transport rates increase more rapidly with temperature). On the other hand, the temperature dependence of the conductance of an open aqueous channel is not so different from that of a film of water (Friedman 2008b). Therefore, variation of transport rate upon temperature changes has been used as experimental evidence of mediated transport. The transport of solutes via carriers and channels has been the subject of extensive kinetic studies, and mathematical models of facilitated diffusion vary depending on the range of assumptions regarding the biophysics of the transport process. However, based on the scope of this volume we may say that kinetics of facilitated diffusion can be appropriately described by the Michaelis-Menten equation if some assumptions are met. For carrier-mediated transport we will assume that the carrier transports a single solute, which binds and unbinds rapidly and reversibly to the substrate, and that the concentration of the substrate in the receiving compartment (C2) is negligible compared to that of the donor compartment (C1) (sink condition). Then:



dC1 Vm·C1  dt Km  C1

(3.18)

where Vm is the maximum rate of transport, which depends on the number of available copies of the carrier and the rate with which the carrier can flip between its two conformational states; Km represents the Michaelis-Menten constant, which is an inverse measure of affinity (note that it equals the substrate concentration at which half the available copies of transporters are occupied by substrate and, thus, half of Vm is observed). At very low substrate concentrations, much below the saturation condition, C1 can be neglected versus Km thus resulting in apparent first order

3  Drug Absorption

47

Fig. 3.9 Michaelis-­ Menten saturation curve for a drug that exhibits carrier-mediated absorption

kinetics (the apparent term denotes that the kinetics is not strictly linear over the entire concentration range, but only at small substrate concentrations):



dC1 Vm·C1  dt Km

(3.19)

Above saturation conditions (in which, at a given moment, every available copy of the transporter is interacting with a substrate molecule), Km could be neglected, and the transport process assumes an apparent zero order kinetics:



dC1  Vm dt

(3.20)

The Michaelis-Menten saturation curve is shown in Fig. 3.9. It should be emphasized that, strictly speaking, Michaelis-Menten equation does not fully represent facilitated diffusion, which is intrinsically dependent on substrate concentrations on both sides of the plasma membrane. The approach is, however, useful to describe apparent transport properties under specific conditions (e.g., initial uptake rates, sink condition) (Panitchob et  al. 2015). More complex mechanistic models are required to capture transporter behavior under diverse physiological conditions.

3.7 Active Transport In active transport, the cell spends energy to move a substrate across biological barriers. The solute can be transported, in this way, uphill, from a phase of low chemical (or electrochemical) potential to a phase of higher potential. In primary active transport, the energy from ATP hydrolysis drives the movement of the substrate

48

A. Talevi and C. L. Bellera

Fig. 3.10 Relationship between the instantaneous rate of absorption and drug concentration at the absorption site, for drug transport occurring by simple diffusion plus active transport or facilitated diffusion

directly (this is the case, for example, of ABC transporters), whereas in secondary or tertiary active transport the cell spends ATP indirectly by coupling the transport of a substrate to the downhill co-transport of a co-substrate (this is the case for the solute carrier group of membrane transport proteins). The co-substrates may move in the same direction (symporters) or in the opposite direction (antiporters) to the solute. For example, active transport of amino acids dissipates sodium and proton gradients and sometimes it also makes use of the downhill movements of other amino acids (Bröer 2008). Michaelis-Menten analysis was developed to model the kinetics of soluble enzymes that bind the substrate directly from an aqueous phase; however, it applies well enough to membrane transporters that bind their substrate directly from the water phase or even to membrane transporters whose substrate resides in the lipid bilayer of the plasma membrane (Bentz et  al. 2005). Kinetic description of co-­ transport of solutes across membranes often requires more complicated models (Friedman 2008c; Korla and Mitra 2014). Notably, some transporters collaborate with drug absorption, whereas others oppose it. The former tends to display relatively narrow substrate specificity; the latter, in contrast, display a wide substrate specificity. Consider that uptake transporters have evolved to help moving specific molecules of physiologic value through membranes; conversely, efflux transporters have evolved to protect cells (or the organism) from non-physiologic compounds (xenobiotics) and to dispose waste products. Uptake and efflux transporters are discussed separately in a specific chapter in this volume. Drug absorption may occur through a combination of transport mechanisms acting in parallel, e.g., carrier-mediated transport plus free diffusion. In such a case, the instantaneous rate of absorption results from the addition of the mathematical equations that explain each of the individual processes involved in absorption. Figure 3.10 illustrates the instantaneous rate of absorption versus concentration at the site of absorption when transport occurs though a combination of simple diffusion plus carrier-mediated transport.

3  Drug Absorption

Example Parkinson’s disease is one of the most common neurodegenerative disorders and is characterized by motor symptoms (tremor, rigidity, slow movement, imbalance) and a wide variety of non-motor complications (cognitive impairment, sleep disorders, etc.). It is characterized by the loss of dopaminergic neurons, with consequent impairment of dopaminergic neurotransmission. Levodopa, an oral dopamine precursor, is used to restore dopamine levels at the central nervous systems exogenously. Thus far, it is the most effective treatment for patients with Parkinson’s disease. Levodopa is rapidly metabolized to dopamine peripherally by amino acid decarboxylase. This is not desirable for two reasons: one, dopamine cannot penetrate the blood-brain barrier; two, peripheral conversion of levodopa to dopamine causes significant gastrointestinal side effects, such as nausea and vomiting. Therefore, levodopa is commonly co-administered with amino acid decarboxylase inhibitors such as carbidopa and benserazide, which confines the conversion of levodopa to dopamine to the central nervous system. Over time, the patients are likely to develop reduced response to levodopa and develop motor complications such as dyskinesia and motor fluctuations (abrupt shifts between “on” and “off” states in which the symptoms are controlled and uncontrolled, respectively). Motor fluctuations in patients in the advanced stages becomes closely connected with the rapid rise and fall in levodopa levels after each dose. Fluctuating motor performance is a major cause of disability in patients with Parkinson’s disease, and drastically impairs their quality of life. Therefore, maximizing the therapeutic efficacy of levodopa is of the utmost importance. Levodopa is molecularly similar to some dietary amino acids and is absorbed via members of the solute carrier family (Camargo et al. 2014). This transport is also responsible for the penetration of levodopa into the central nervous system, where it exerts its therapeutic action. The presence of large neutral amino acids from the diet either in the intestinal lumen or in blood may reduce levodopa absorption and distribution to the brain, in that order, since they compete with levodopa for the transporter. Accordingly, protein-restricted diets are effective in ameliorating motor fluctuations (Wang et al. 2017; Barichella et al. 2017). These diets include: low-protein diets (in which protein consumption is restricted to 0.5–0.8 g protein/per kilogram of body weight per day) and protein-redistribution diets (in which the patient consumes low-protein food such as cereal products, fruits and vegetables for breakfast and lunch, and is allowed high-protein food including eggs, legumes, fish and meat for dinner. This maximizes daytime motor function, since diurnal fluctuations are considered to limit the quality of life more than fluctuations at night).

49

50

A. Talevi and C. L. Bellera

Case Study A 67-year-old patient with a 23-year history of Parkinson’s disease and type 2 diabetes presented to the emergency department on New Year’s Day. She had hosted a large New Year’s Eve party in the evening. She was on levodopa/ carbidopa therapy as well as two dopamine agonists (pramipexole and amantadine) and metformin. She had taken her medication at 10  p.m. on New Year’s Eve, although it was unclear whether she had taken her tablets at 11 a.m. A large protein-rich Chinese meal was consumed shortly before her last confirmed treatment dose. On New Year’s Day she slept at 11 a.m. after cooking lunch. At 2 p.m. her son could not rouse her and called an ambulance. Blood sugar was 9.0 mmol/l. She had a Glasgow Coma Scale score of 5/15. Her blood pressure was 130/78 mmHg, pulse 84 beats/minute and blood oxygen saturation 98% on room air. No focal neurological deficit was observed. A tomographic scan revealed a mild degree of cerebral atrophy but was otherwise normal. Routine hematology and biochemistry studies were normal. Peripheral white blood cell count was normal. Treatment with intravenous acyclovir was commenced to cover possible viral encephalitis Shortly before admission, she experienced an increased frequency of falls. The patient recovered rapidly following reinstatement of dopaminergic therapy via nasogastric tube. Considering that severe motor ‘wearing off’ phenomena can be life threatening, in extreme cases resembling neuroleptic malignant syndrome with confusion, hyperthermia and rhabdomyolysis, discuss possible causes to the patient’s condition. The results of hematology and biochemistry studies and other tests performed on the patient should be taken into consideration. Based on a case report by Arulanantham et al. (2014).

Overview –– In general terms, drug molecules must be in solution in the site of absorption for absorption to occur (with the possible exception of drugs included in some last-generation pharmaceutical carriers, such as nanocarriers). –– The vast majority of drugs are absorbed by simple diffusion and use the transcellular path to cross epithelia and/or endothelia. –– The same mass transfer mechanisms studied in this chapter will explain the drug distribution process and the uptake of drugs by organs responsible for drug elimination. –– For a drug to be absorbed through carrier-mediated transport, it may usually present molecular similarity to natural/physiological substrates of the involved carrier, such as amino acids, peptides, sugars or nucleotides. –– The kinetics of simple diffusion can be modeled using Fick’s first law. In contrast, the kinetics of facilitated/carrier-mediated transport can usually be described by the Michaelis-Menten equation, with exceptions.

3  Drug Absorption

51

–– Simple diffusion is a nonsaturable (linear) process. Facilitated diffusion and active transport require transmembrane proteins and are saturable processes. –– Since most drugs are either weak acids or weak bases, their solubility and permeability are highly dependent on the pH of the absorption site. Moreover, the electrochemical gradient should be considered to study the absorption of charged species rigourously. –– Ionic species tend to be more soluble than neutral species; conversely, charged species are less permeable.

References Adrien V, Rayan G, Astafyeva K et al (2022) How to best estimate the viscosity of lipid bilayers. Biophys Chem 281:106732 Al-Mohizea A, Zawaneh F, Alam MA et al (2014) Effect of pharmaceutical excipients on the permeability of p-glycoprotein substrate. J Drug Deliv Sci Technol 24:491–495 Alberts B, Johnson A, Lewis J et al (2002) Membrane structure. In: Alberts B, Johnson A, Lewis J et al (eds) Molecular biology of the cell. Garland Science, New York Arulanantham N, Lee RW, Hayton T (2014) Lesson of the month 2: a case of coma in a Parkinson's patient: a combination of fatigue, dehydration and high protein diet over the new year period? Clin Med (Lond) 14:449–451 Barichella M, Cereda E, Cassani E et  al (2017) Dietary habits and neurological features of Parkinson’s disease patients: implications for practice. Clin Nutr 36:1054–1061 Bazzoni G (2006) Endothelial tight junctions: permeable barriers of the vessel wall. Thromb Haemost 95:36–42 Bentz J, Tran TT, Polli JW et  al (2005) The steady-state Michaelis-Menten analysis of P-glycoprotein mediated transport through a confluent cell monolayer cannot predict the correct Michaelis constant km. Pharm Res 22:1667–1677 Bicknese S, Periasamy N, Shohet SB et  al (1993) Cytoplasmic viscosity near the cell plasma membrane: measurement by evanescent field frequency-domain microfluorimetry. Biophys J 65:1272–1782 Brodin B, Steffansen B, Nielsen CU (2009) Passive diffusion of drug substances: the concepts of flux and permeability. In: Steffansen B, Brodin B, Nielsen CU (eds) Molecular biopharmaceutics: aspects of drug characterization. Drug Delivery and Dosage Form Evaluation, Pharmaceutical Press, London, pp 135–152 Broër S (2008) Amino acid transport across mammalian intestinal and renal epithelia. Physiol Rev 88:249–286 Camargo SM, Vuille-dit-Bille RN, Mariotta L et al (2014) The molecular mechanism of intestinal levodopa absorption and its possible implications for the treatment of Parkinson's disease. J Pharmacol Exp Ther 351:114–123 Censi R, Di Martino P (2015) Polymorph impact on the bioavailability and stability of poorly soluble drugs. Molecules 20:18759–18776 Chan TC, Li HT, Li KY (2015) Effects of shapes of solute molecules on diffusion: a study of dependences on solute size, solvent, and temperature. J Phys Chem B 119:15718–15728 Clausen JP (1977) Effect of physical training on cardiovascular adjustments to exercise in man. Physiol Rev 57:779–815 Dahan A, Beig A, Lindley D et al (2016) The solubility-permeability interplay and oral drug formulation design: two heads are better than one. Adv Drug Deliv Rev 101:99–107 Davis SS (2005) Formulation strategies for absorption windows. Drug Discov Today 10:249–257

52

A. Talevi and C. L. Bellera

Falcon RW, Kakuda TN (2008) Drug interactions between HIV protease inhibitors and acid-­ reducing agents. Clin Pharmacokinet 47:75–89 Friedman MH (2008a) Free diffusion. In: Principles and models of biological transport, Springer, New York Friedman MH (2008b) Facilitated diffusion: channels and carriers. Principles and models of biological transport, Springer, New York, In Gidal BE (2006) Drug absorption in the elderly: biopharmaceutical considerations for the antiepileptic drugs. Epilepsy Res 68(Suppl 1):S65–S69 González-Canudas J, García-Aguirre LJ, Medina-Nolasco A et al (2019) Bioequivalence evaluation of two oral formulations of Dexketoprofen-trometamol (solution and tablets) in healthy subjects: results from a randomized, single-blind, crossover study. Trends Med 19:2–5 González-Mariscal L, Betanzos A, Nava P et al (2003) Tight junction proteins. Prog Biophys Mol Biol 81:1–44 González-Mariscal L, Posadas Y, Miranda J et al (2017) Strategies that target tight junctions for enhanced drug delivery. Curr Pharm Des 22:5313–5346 Gursoy RN, Benita S (2004) Self-emulsifying drug delivery systems (SEDDS) for improved oral delivery of lipophilic drugs. Biomed Pharmacother 58:173–182 Hodaei D, Baradaran B, Valizadeh H et al (2014) The effect of tween excipients on expression and activity of p-glicoprotein in Caco-2 cells. Pharm Ind 76:788–794 Hoheisel C (1988) Effect of the molecule shape on the diffusion in molecular fluids. Molecular dynamics calculation results for SF6–C6H6 and a related model solution. J Chem Phys 89:7457–7461 Johansson MEV, Sjövall H, Hansson GC et al (2013) The gastrointestinal mucus system in health and disease. Nat Rev Gastroenterol Hepatol 10:5352–5361 Kinoshita R, Ohta T, Shiraki K et al (2017) Effects of wet-granulation process parameters on the dissolution and physical stability of a solid dispersion. Int J Pharm 524:304–311 Kemmer FW, Berchtold P, Berger M et al (1979) Exercise-induced fall of blood glucose in insulin-­ treated diabetics unrelated to alteration of insulin mobilization. Diabetes 28:1131–1137 Koivisto VA, Felig P (1978) Effects of leg exercise on insulin absorption in diabetic patients. N Engl J Med 298:79–83 Kováčik A, Kopečná M, Vávrová K (2020) Permeation enhancers in transdermal drug delivery: benefits and limitations. Expert Opin Drug Deliv 17:145–155 Korla K, Mitra CK (2014) Kinetic modelling of coupled transport across biological membranes. Indian J Biochem Biophys 51:93–99 Lee Y, Choi SQ (2019) Quantitative analysis for lipophilic drug transport through a model lipid membrane with membrane retention. Eur J Pharm Sci 134:176–184 Leung SS, Mijalkovic J, Borrelli K et  al (2012) Testing physical models of passive membrane permeation. J Chem Inf Model 52:1621–1636 Lewis JM, Stott KE, Monnery D et al (2016) Managing potential drug-drug interac-tions between gastric acid-reducing agents and antiretroviral therapy: experience from a large HIV-positive cohort. Int J STD AIDS 27:105–109 Li W, You L, Schaffler MB et al (2009) The dependency of solute diffusion on molecular weight and shape in intact bone. Bone 45:1017–1023 Maher S, Brayden DJ, Casettari L et al (2019) Application of permeation enhancers in oral delivery of macromolecules: an update. Pharmaceutics 11:41 O'Shea JP, Augustijns P, Brandl M et al (2022) Best practices in current models mimicking drug permeability in the gastrointestinal tract - an UNGAP review. Eur J Pharm Sci 170:106098 Panitchob N, Widdows IP, Crocker MA et  al (2015) Computational modelling of amino acid exchange and facilitated transport in placental membrane vesicles. J Theor Biol 365:352–364 Rezai T, Bock JE, Zhou MV et al (2006) Conformational flexibility, internal hydrogen bonding, and passive membrane permeability: successful in silico prediction of the relative permeabilities of cyclic peptides. J Am Chem Soc 128:14073–14080 Ribeiro RP, Coimbra JTS, Ramos MJ et al (2017) Diffusion of the small, very polar, drug piracetam through a lipid bilayer: an MD simulation study. Theor Chem Accounts 136:46

3  Drug Absorption

53

Russell TL, Berardi RR, Barnett JL et al (1993) Upper gastrointestinal pH in seventy-nine healthy, elderly, north American men and women. Pharm Res 10:187–196 Sandri G, Bonferoni MC, Ferrari F et al (2014) The role of particle size in drug release and absorption. In: Merkus H, Meesters G (eds) Particulate Products. Springr, Cham, pp 323–341 Santos JV, Batista de Carvalho LA, Pina ME (2010) The influence of the compression force on zidovudine release from matrix tablets. AAPS PharmSciTech 11:1442–1448 Seddon AM, Casey D, Law RV et al (2009) Drug interactions with lipid membranes. Chem Soc Rev 38:2509–2519 Singer SJ, Nicolson GL (1972) The fluid mosaic model of the structure of cell membranes. Science 175:720–731 Singh I, Morris AP (2011) Performance of transdermal therapeutic systems: effects of biological factors. Int J Pharm Investig 1:4–9 Sobue S, Sekiguchi K, Kikkawa H et  al (2006) Comparison of nicotine pharmacokinetics in healthy Japanese male smokers following application of the transdermal nicotine patch and cigarette smoking. Biol Pharm Bull 29:1068–1073 Soriano-Meseguer S, Fuguet E, Port A et al (2020) Optimization of experimental conditions for skin-PAMPA measurements. ADMET DMPK 8:16–28 Sultatos L (2011) Drug absorption from the gastrointestinal tract. In: Enna SJ, Bylund DB (eds) xPharm: the comprehensive pharmacology reference. Elsevier, Boston Tompkins L, Lynch C, Haidar S et al (2010) Effects of commonly used excipients on the expression of CYP3A4 in colon and liver cells. Pharm Res 27:1703–1712 Tomilo DL, Smith PF, Ogundele AB et al (2006) Inhibition of atazanavir oral absorption by lansoprazole gastric acid suppression in healthy volunteers. Pharmacotherapy 26:341–346 Van Itallie CM, Anderson JM (2004) The molecular physiology of tight junction pores. Physiology (Bethesda) 19:331–338 van Lunzen J, Liess H, Arastéh K et al (2007) Concomitant use of gastric acid-reducing agents is frequent among HIV-1-infected patients receiving protease inhibitor-based highly active antiretroviral therapy. HIV Med 8:220–225 Veber DF, Johnson SR, Cheng HY (2002) Molecular properties that influence the oral bioavailability of drug candidates. J Med Chem 45:2615–2623 Wang L, Xiong N, Huang J et al (2017) Protein-restricted diets for ameliorating motor fluctuations in Parkinson’s disease. Front Aging Neurosci 9:206 Wang CK, Swedberg JE, Harvey PJ et al (2018) Conformational flexibility is a deterinant of permeability for cyclosporin. J Phys Chem B 122:2261–2276 Yang NJ, Hinner MJ (2015) Getting across the cell membrane: an overview for small molecules, peptides, and proteins. Methods Mol Biol 1266:29–53

Further Reading The reader is referred to the extraordinary volume Principles and Models of Biological Transport second Edition (Springer Germany, 2008) for a comprehensive description of free diffusion, facilitated diffusion and, active transport. The wonderful volume by Carsten Ehrhardt and Kwan-Jin Kim (Drug Absorption Studies. In Situ, In Vitro and In Silico models, Springer, 2008) can be consulted for additional insight on the absorption process with a focus on experimental and computational models of absorption. The Molecular Biopharmaceutics volume by Bente Steffansen et al. is highly recommended to study experimental models of drug absorption (Molecular Biopharmaceutics, Pharmaceutical Press, 2010)

Chapter 4

Drug Distribution Alan Talevi and Carolina Leticia Bellera

4.1 Introduction Definition Drug distribution describes the (generally reversible) mass transfer of a drug from one location to another within the body. Once drug molecules have been absorbed (i.e., when they enter systemic circulation), they might extravasate into the interstitial fluid of extravascular tissues and, subsequently, into the intracellular space. Each organ or tissue can receive different levels of the drug at different rates, and a drug can remain in the different organs or tissues for different amounts of time depending on the tissue characteristics. The mass transfer of drugs from the vascular to the extra-vascular space occurs through the endothelial lining of the blood capillaries. Accordingly, the rate of mass transfer is strongly dependent on the features of the capillary bed under consideration. The possible drug transfer mechanisms across endothelia and cell membranes are essentially similar to those previously described for drug absorption. Moreover, distribution will be also influenced by unspecific, reversible binding events to plasma proteins (which will conspire against drug extravasation) and tissue elements (which will promote drug extravasation). A key concept to understand distribution is that if conventional drug dosage forms are considered, only free, unbound drug molecules can cross endothelia and thus engage in the pseudo-equilibrium of distribution in a direct manner.

A. Talevi (*) · C. L. Bellera Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_4

55

56

A. Talevi and C. L. Bellera

In the previous chapter, we discussed the mass transfer of a drug from the site of administration/absorption to the systemic circulation. We will use the term drug distribution to describe the phenomena and processes involved in the exchange of drug molecules between the blood and the rest of the tissues (extravascular tissues). As absorption develops, the drug molecules progressively arrive at the blood, resulting in a concentration gradient across the blood vessel walls. This gradient explains the diffusion of drug molecules from the vascular space to the interstitial compartment, in the first place, and towards the intracellular fluid, later. Exchange is facilitated at the level of blood capillaries, the smallest blood vessels that are a few micrometers in diameter and possess a wall that is one endothelial cell thick. In addition to the facilitated diffusion provided by this narrow diffusion barrier, capillaries are part of a capillary bed, an interweaving network of small vessels supplying tissues and organs, and presenting an impressive surface area. Drug distribution is thus an eminently passive process, except for drug molecules recognized by transporters involved in facilitated or  active transport. Molecules that are substrates for transporters can use them to extravasate and/or enter the cells of extravascular tissues. Exchange is a key term here, reflecting the reversible (two-way) nature of the process. While it is true that the initial sense of the drug transfer will be from the blood to the extravascular tissues, if occasionally the free drug levels in blood fall, a net return of drug molecules from the extravascular tissues to the vascular space will be observed. Underlying the previous description is the idea of distribution equilibrium, a concept that is frequently found in the literature when approaching the topic of drug distribution. However, can a true distribution equilibrium be achieved? Can equilibrium be achieved within living beings? if we consider living organisms from a thermodynamic perspective, we will recognize an open system subjected to constant exchange of heat and matter with its environment. Living organisms are thus highly dynamic systems, exposed to permanent change, and invest high amounts of energy to maintain balance (homeostasis is a thermodynamic stationary state of non-­ equilibrium) (Recordati and Bellini 2004). Biological systems reach a state of thermal and mechanical equilibrium with the external environment only when they are dead. In addition to strict thermodynamic considerations, there are also pharmacokinetic reasons to regard distribution equilibrium as impossibility. Once the drug has been absorbed, it will be continuously subjected to elimination. Moreover, drug delivery (i.e., drug administration) is typically performed in a discontinuous, intermittent manner (through the administration of subsequent dosage units). Only in those occasions where the drug enters the body in a continuous manner and follows zero-order absorption  kinetics (that is, at a constant rate  of  absorption) will we approach a true state of distribution equilibrium (Fig. 4.1). In light of the previous discussion on the thermodynamics of living systems, the resulting steady state will frequently and more appropriately be referred to as a pseudo-equilibrium. Under such circumstances, at some point after starting administration of the drug, a steady state is reached in which the rate of drug inflow to the blood equals the rate of drug elimination from the blood, and no net change is observed in drug levels as long as the in-flow and elimination rates remain unchanged. This will likely occur in just a

4  Drug Distribution

57

Fig. 4.1  Plasma concentration versus time profile for a drug that reaches systemic circulation with zero- order kinetics (i.e., at constant rate). The so obtained steady state is possibly the closest scenario to a distribution equilibrium

few therapeutic settings: a) continuous intravenous infusion of a drug and b) when modified-release drug delivery systems that achieve sustained release are used. With immediate-release dosage forms, the steady state will only be met through the administration of multiple and identical doses at regular time intervals. In such cases, the steady state results from a balance between the mean absorption rate and the mean elimination rate during a dosing interval and will consist of fluctuations of the plasma drug levels between practically constant maximal and minimal concentrations. It then consists of a (peudo) steady state, in which plasma levels oscillate within an approximately fixed range of concentrations (Fig. 4.2).

4.2 Factors Impacting on [The Extent and Kinetics of] Drug Distribution Several factors influence drug distribution. First, we may consider some physicochemical properties of the drug that can directly or indirectly affect its capacity to cross biological barriers. The general mechanisms of drug transport across biological barriers have already been discussed in Chap. 3. Among the salient factors intrinsic to the drug, molecular size (the larger a molecule, the more difficult its passive diffusion) and polarity (polar molecules will face difficulties in moving through simple diffusion, and their movement will also be reduced in epithelia or endothelia where the paracellular route is precluded, e.g., at the blood-brain

58

A. Talevi and C. L. Bellera

Fig. 4.2  Plasma concentration versus time profile following multiple doses of an intravenous bolus at regular dosing intervals

barrier) are of major importance. The degree of ionization is also key because the net charge of the molecule depends on both the ionization constant(s) of the drug and the pH of the physiologic environment (Ito et al. 2013). The affinity of a drug for drug transporters should also be considered when studying drug distribution. In fact, a substantial proportion of drugs predominantly exhibit a net charge at pH 7.0 (that is, close to the blood pH). Cellular uptake of charged species is thus likely to occur via mediated transport (see Chap. 19 for further details). On the other hand, a group of physiological factors that have an impact on drug distribution can be listed, which are, in some cases, subjected to variation in an individual in both health and disease states. These factors include the expression levels of drug transporters in different tissues, the characteristics of the capillary beds that supply different tissues and organs, including the permeability of tight junctions between adjacent endothelial cells and the presence of pores and discontinuities, and tissue or organ perfusion (blood flow to a given organ, that is, the rate at which blood is delivered to a given tissue or organ, measured as the volume of blood per unit of time per unit of tissue mass). It should be noted that the fraction of cardiac output received by an organ dynamically changes depending on its metabolic demand. For instance, the blood flow to different organs changes during exercise, in the resting state, in the postprandial state, etc. The affinity between the drug and tissue (which depends on the unspecific interactions that occur between the drug and tissue elements) should also be taken into consideration. The reader should have in mind that different types of capillary beds exist, each with different degrees of “leakiness”, as can be seen in the following definitions and in Fig. 4.3.

4  Drug Distribution

Fig. 4.3  Types of capillaries

Definition Continuous capillaries, the most widely distributed type of capillaries, are found in the muscle, skin, lungs, adipose tissue, and nervous system, and are characterized by a complete endothelial lining with tight junctions between endothelial cells and an intact basement membrane. The intercellular tight junctions usually display intercellular clefts that allow for the exchange of water and other small polar molecules between blood plasma and interstitial fluid. Continuous capillaries not associated with the brain are rich in transport vesicles that participate in endocytosis and/or exocytosis. Those in the brain are part of the blood-brain barrier. Fenestrated capillaries: this type of capillaries have pores (or fenestrations, from Latin fenestra, i.e., “window”) in addition to tight junctions in the endothelial lining. This makes the capillary permeable to the larger molecules. They are found in the glomerulus of the kidney, the intestinal villi and the choroid plexus of the brain, and also in many endocrine structures. As with continuous capillaries, they have an intact basement membrane. Sinusoid capillaries (or simply sinusoids), the least common type of capillary, are flattened and have extensive intercellular gaps and incomplete basement membranes, in addition to intercellular clefts and fenestrations. This allows for the passage of large molecules and structures, including plasma proteins and even cells. Blood flow through sinusoids is very slow, allowing more time for solute exchange. Sinusoids are found in the liver, spleen, bone marrow, lymph nodes, and some endocrine glands. Blood-brain barrier: The endothelial lining of capillary beds that perfuse the brain present especially tight tight junctions with no intercellular clefts, plus a thick basement membrane and astrocyte extensions called end feet. In combination with the high expression of efflux transporters, these structures pose a highly selective permeability barrier that provides precise control over which substances can enter or exit the brain.

59

60

A. Talevi and C. L. Bellera

4.3 Different Pharmacokinetic Models Describe Drug Distribution in Different Ways It is convenient to note that different pharmacokinetic models will describe the distribution process in different ways, both conceptually and mathematically. For instance, classical compartmental models depict the body as one (one-compartment model, Fig. 4.4), two (two-compartment model, Fig. 4.5), or more pharmacokinetically uniform compartments and assume that the drug is distributed uniformly and instantaneously within each compartment, although movement from one compartment to others is not instantaneous but follows first-order kinetics. Definition In the framework of classical compartmental modeling, a pharmacokinetic compartment is an abstract space (with no direct relationship to anatomical spaces, tissues, or fluids) in which the drug is assumed to be distributed homogeneously and instantaneously. Whereas the distribution within a compartment is assumed to be instantaneous, drug exchange between compartments is not. Organs and tissues with similar pharmacokinetic behavior (e.g., similar distribution kinetics) are pooled into one compartment. Although it is difficult to imagine that any drug will be distributed uniformly across different organs or tissues, and it is even harder to believe that a mass transfer process may be instantaneous, such assumptions or simplifications lead to models that are relatively simple from a mathematical perspective.

As empirical models, classic compartmental pharmacokinetic models are useful for data description and interpolation and may admit some rough physiological parameterization. However, they do not have a direct physiological interpretation, which makes it difficult to predict kinetic profiles when underlying physiological changes occur (Aarons 2005; Espié et al. 2009). In contrast, physiologically based pharmacokinetic models strive to be mechanistic by mathematically transcribing the anatomical, physiological, physical, and bio-chemical descriptions of the phenomena involved in the complex ADME processes. They describe the pharmacokinetics of drugs within the body in relation to blood flow, tissue volume, route of Fig. 4.4  Classic one-­ compartment pharmacokinetic model. Ka is the absorption rate constant and K is the elimination rate constant

4  Drug Distribution

61

Fig. 4.5  Classic two-compartment pharmacokinetic model with first-order absorption and elimination. 1 and 2 are the central and peripheral compartments, respectively. Ka, K10, K12, and K21, represent, in that order, the kinetic constants associated with absorption, elimination, drug transfer from compartment 1 to compartment 2 and drug transfer from compartment 2 to compartment 1. In this context, Ka has also been represented as K01

administration, interactions with tissues or organs, and biotransformation pathways. In essence, they are multi-compartment models but differ from classical pharmacokinetic models in that the compartments represent actual tissue and organ spaces. The biological and mechanistic bases of physiologically based models allow a better extrapolation of the kinetic behavior of drugs with respect to dose, route, species, individuals from different populations (males to females, adults to children, non-­ pregnant to pregnant women, etc.), and even in vitro to in vivo (Bouvier d’Yvoire et  al. 2007). An example of a physiologically based model is shown in Fig.  4.6. Further information on classic and physiologically based compartmental models is provided in Chaps. 8 and 9, respectively.

Important In contrast to classic compartmental models, physiologically based pharmacokinetic models represent the body using pharmacokinetically relevant anatomical compartments and rely on physiological data. They present an extended domain of applicability and facilitated transposition/extrapolation. Some of these extrapolations are parametric; only changes in input or parameter values are needed to achieve the extrapolation (e.g., dose extrapolations). Other transpositions are nonparametric; a change in the model structure itself is needed (e.g., when extrapolating to a pregnant female, equations for the fetus should be added).

From the previous discussion, it can be seen that physiologically based models better capture the factors that influence drug distribution. In the context of this type of

62

A. Talevi and C. L. Bellera

Fig. 4.6  Physiologically based model describing the pharmacokinetics of doxycycline in swine after intramuscular administration. Q represents the blood flow. Clre and Clhe are the renal and hepatic clearances, respectively. F is the absolute bioavailability of doxycycline after intra-­ muscular administration, and Ka is absorption rate constant. Kbile, Kra and Kic are the biliary excretion rate constant, the intestinal reabsorption rate constant, and the intestinal elimination rate constant, respectively. Qtot is cardiac output. The figure is based on the model reported in Yang et al. (2012) Food Additives & Contaminants: Part A 29: 73–84

pharmacokinetic modelling and depending on the rate of drug uptake by the tissues or organs, we may classify the distribution as a) perfusion-limited or blood flowlimited and b) permeability-limited or membrane-limited (Espié et  al. 2009; Korzekwa et al. 2022). In the first situation, it is assumed that the drug dis-tributes instantaneously from the capillaries into the interstitial fluid and tissue cells so that the rate-limiting step governing the movement of drug molecules into and out of the organ is the blood flow to the organ, and no concentration gradients are likely within the considered physiological compartment. This may be a reasonable assumption in the case for small lipophilic drugs. Such drugs are rapidly distributed to highly perfused organs, such as the brain, kidneys, lungs, and liver. Table 4.1 presents a list of the blood flow to different organs. In the second situation (membrane-limited distribution), membrane permeability limits the distribution within the tissue/organ, and the tissue may be divided into sub-compartments with a permeability-rate-limited transfer between them (Nestorov 2003). The use of a membrane-limited organ model is indicated if tissue drug concentrations do not decline in parallel with the plasma drug concentrations. This occurs, for example, when the affinity of a tissue for the compound is high; when large and physically distributed tissues are modelled (such as muscles or the skin) and when the kinetic of large molecules are modelled and thus diffusion and transport impediments are anticipated. Figure 4.7 schematically illustrates the difference between perfusion-limited and permeability-­ limited distribution.

4  Drug Distribution Table 4.1  Regional blood flow distribution in a 70 kg healthy subject at rest. (Data extracted from Calzia et al. (2005)

63

Organ Kidneys Liver Heart Brain Muscle Skin

Blood flow (L/kg/min) 4.0 0.9 0.8 0.5 0.08 0.03

Fig. 4.7  Perfusion-limited (a) and permeability-limited (b) distribution. A thick double-headed arrow denotes a fats drug exchange, whereas a narrow arrow denotes a limited, slow exchange resulting in a persisting concentration gradient

4.4 Real and Apparent Volumes of Distribution Definition The real volume of distribution of a drug consists in the volume of body water in which the drug is distributed (including intracellular and extracellular fluids). Total body water constitutes 50–70% of the total body weight; two-thirds of it corresponds to intracellular fluid, whereas the remaining one-third corresponds to extracellular fluid (75% extravascular interstitial fluid and 25% plasma) (Dineen and Schumacher 2009). An adult has 500–700 mL of water per kilogram of body weight, depending on factors such as sex or age). Therefore, the upper bound for the real volume of distribution was approximately 42 L for a person weighing 70 kg. In such individuals, no drug will have a real volume of distribution significantly above that value. Only a drug that accesses the entire body fluid will achieve the upper bound of the real volume of distribution. Conversely (i.e., if the drug is not fully distributed to a given tissue or organ), the real volume of distribution will be below the corresponding upper limit. Table  4.2 presents some approximate values of the body fluid volumes. After reaching systemic bioavailability, drug molecules rapidly disperse within the plasma pool. Accordingly, the lower bound of the

64

A. Talevi and C. L. Bellera

Table 4.2 Approximate physiological volumes for a 70 kg healthy subject

Fluid Plasma Blood Interstitial fluid Extracellular water Intracellular water Total body water

Volume (L) 3 5 12 15 27 42

real volume of distribution corresponds to the plasma volume, corresponding to a drug that can neither extravasate nor penetrate the blood cells. Similar to other reference body volumes, the plasma volume is also subjected to significant variation across sex, age, body surface area or health state. For instance, young male adults weighing 60–65 kg have plasma volumes of approximately 3000 mL (Yiengst and Shock 1962), and the mean plasma volume tends to be smaller in females (Wadsworth 1954; Brown et al. (1962). On the other hand, significant increase in plasma volume has been reported in patients with chronic heart failure (Fudim and Miller 2018). The apparent volume of distribution Vd, in contrast is (in the context of the one-compartment model) a proportionality constant between the total amount of drug in the body and drug plasma concentration at a given time. In other words, it is the theoretical volume that would be necessary to contain the total amount of a drug in the body if its concentration was homogeneous and identical to the concentration measured in plasma. In mathematical terms: Vd =



A Cp

(4.1)

where A is the total amount of drug in the body and Cp represents the total drug plasma concentration, both of which are time-dependent variables since they are subjected to constant changes owing to absorption and elimination processes. However, as in the one-compartment model distribution is assumed to be instantaneous, the ratio between both variables (i.e., Vd) will be constant (time-independent). Importantly, the plasma drug concentration in the denominator of Eq. 4.1 is the total concentration, including bound and unbound drug molecules. Vd can be estimated after intravenous administration of the drug (Kwon 2002). If the plasma concentration versus time profile follows a monoexponential decay after the administration of an intravenous bolus, the apparent volume of distribution can be estimated by dividing the intravenous dose D by the estimated drug concentration at time 0 (Cp0), determined by back-­ extrapolating to the intercept with the Y-axis. Vd =

A C p0

(4.2)

4  Drug Distribution

65

Frequently, the apparent volume of distribution will be simply referred as “volume of distribution” in the specialized literature. As discussed in Chap. 8, other volumes of distribution may be computed, as snapshot plasma drug concentrations may be assessed under different conditions (Toutain and Bousquet-Mélou 2004). It has previously been mentioned that the apparent volume of distribution is an imaginary entity that does not refer to any identifiable compartment in the body. Therefore, its physical interpretation is not unequivocal. Two drugs with different distribution patterns may have similar Vd values.

4.5 The Source of Disagreement: Discrepancies Between Apparent and Real Volumes Table 4.3 shows the approximate values of Vd for a selection of drugs. It is evident that Vd does not have any anatomical meaning; for some compounds (chloroquine, chlorpromazine, digoxin), the value of this parameter greatly exceeds the total volume of body water. The origin of the discrepancies between the real and apparent volumes is the lack of homogeneity in drug levels across different organs or tissues, even if the steady state has been reached. Such a lack of uniform distribution mainly arises from the affinity of the drug for drug carriers that perform active transport and/or nonspecific interactions between the drug and tissue constituents. The effect of active transport on drug concentrations is understandable: cells invest energy in maintaining chemical/electrochemical gradients across membranes, often transporting their substrates uphill. Accordingly, for transporter substrates the free drug hypothesis (see Chap. 1) does not apply and free unbound drug concentrations at both sides of the biological barrier will not be equal. Regarding nonspecific interactions with tissue elements, such elements may be located within or outside the vascular space. In the first case, the most relevant interaction is the one established between the drug and plasma proteins. In the second, lipids and proteins are common binding partners at extravascular tissues, although the binding partners vary from drug to drug, as further discussed in subsection 4.5.2. For instance, chloroquine and its analogs (e.g., hydroxychloroquine) have been shown to be differentially sequestered by different tissues (its concentration in the liver, spleen and adrenal gland is 6000–80,000 times that of plasma) (Browning Table 4.3 Approximate apparent volumes of distribution for a selection of drugs, for a person of 70 kg

Drug Chloroquine Chlorpromazine Digoxin Diazepam Phenytoin Phenobarbital Aspirin Ibuprofen

Vd (L/70 kg) 15,000 1,400 490 63 49 42 10.5 8.4

66

A. Talevi and C. L. Bellera

2014) and it has been demonstrated that they bind to melanin, DNA, ferriprotoporphyrin, and phospholipids, and that they also accumulate in lysosomes (MacIntyre and Cutler 1986; Schroeder and Gerber 2014). The influence of drug binding to plasma proteins and extravascular tissue components can be examined through the series of equilibria displayed in Fig. 4.8. A key concept worth remembering to gain a basic understanding of drug distribution is that, following the free drug hypothesis, only free (unbound) drug molecules directly participate in the distribution process and tend to equilibrate at steady state. Le Chatelier’s principle is also useful to anticipate the effect of a change in the biological system (i.e., drug displacement) on drug distribution: when a system at equilibrium is subjected to change (e.g., in the concentration of the drug or the binding partner), the system readjusts itself to counteract the effect of the applied change, evolving to a new equilibrium state. Case Study Sixteen women receiving hydroxychloroquine, chloroquine or a combination of both agents were monitored over 7 years. Through ophthalmologic examinations were undertaken in all 16 patients. 9 of the patients were taking doses above than the recommended dosage, adjusted by lean body weight. 10 of the patients presented difficulties with reading. Fundus photography revealed changes ranging from mild retinal pigment epithelial changes to bull’s-eye maculopathy, with only 3 patients having a normal-appearing macula. When subjected to full-field electroretinograms, 6 patients showed evidence of generalized retinal dysfunction with reduced rod and cone responses. 15 patients underwent multifocal electroretinography and had evidence of bilateral macular cone dysfunction. How can evidence for retinal issues be explained? Based on data from Michaelides et al. (2011).

4.5.1 Binding to Plasma Proteins and Blood Cells The high concentrations of plasma proteins along with the propensity of most drugs to form reversible complexes with them, explain the great interest in understanding the impact of this process on pharmacokinetics and pharmacodynamics. The major drug-binding components in plasma are albumin, α-acid glycoprotein (AAG), and lipoproteins such as γ-globulin (Bohnert and Gan 2013). Some drugs show an affinity for blood cells (Kalamaridis and DiLoretto 2014). However, for ease of clinical analysis, plasma has historically been the most commonly analyzed fluid (Rowland and Tozer 1994), and binding to blood cells (in particular, erythrocytes) has been frequently neglected in pharmacokinetic studies (Hinderling 1997). By analyzing the equilibria in Fig. 4.8, it can be seen that since the free drug in plasma establishes a comparatively rapid equilibrium with the extravascular free drug, but the bound drug does not, drug binding to plasma proteins will result in total

4  Drug Distribution

67

Fig. 4.8  A series of equilibria can be used to understand the influence that binding to plasma protein and extravascular tissue elements have on distribution

plasma drug levels that overestimate the extravascular drug levels. Accordingly, the Vd for drugs with high affinity for plasma proteins (i.e., with high bound fractions) will tend to be small, especially if the affinity for plasma proteins is not compensated by a similar or higher affinity for tissue elements. In most cases, small Vds indicate the high affinity of the drug for plasma proteins and low affinity to extravascular tissues. Consider the free (or unbound) fraction of a drug in plasma: fu =

C pfree Cp

(4.3)

where Cpfree is the free drug plasma concentration and Cp is the total drug plasma concentration (free plasma drug concentration plus bound plasma drug concentration, Cpbound). For a drug with a single family of binding sites in its binding partner (i.e., binding sites characterized by the same affinity constant) Cpbound is provided by the following general equation (Toutain and Bousquet-Mélou 2002) (note its similarity to the Michaelis-Menten equation): Cbound 

Bmax  C pfree K D  C pfree

(4.4)

Bmax denotes the maximal binding capacity, which is directly related to the molar concentration of the binding protein and the number of binding sites per copy of the

68

A. Talevi and C. L. Bellera

binding protein, and KD is the equilibrium dissociation constant of the complex between the drug and the plasma proteins, which is the inverse of the affinity constant KA: KA 

C pbound C pfree  P

(4.5)

P represents the unoccupied plasma protein concentration. Now consider the total drug concentration, that following expression 4.4 might be expressed as: C p  C pfree 

Bmax  C pfree K D  C pfree

(4.6)

After factorization and taking the inverse of the previous equation: fu 

K D  C pfree Bmax  K D  C pfree

(4.7)

When a drug binds to more than one type of binding site with different affinities, several equilibria are involved, each of which is characterized by distinctive dissociation constants (Koch-Weser and Sellers 1976). Naturally, when the drug binds to only one binding site per binding partner molecule, the maximum molar concentration of drug that may bind the correspondent protein is equal to the molar concentration of the binding partner. The maximum possible binding capacity of a plasma protein (in relation to a given drug) equals the number of binding sites for that drug per protein molecule times the molar concentration of the protein. Several experimental methods are routinely used to study drug binding to plasma proteins, either qualitatively or quantitatively; among them, equilibrium dialysis, ultrafiltration and ultracentrifugation are used to quantify the unbound drug concentration and determine the unbound fraction, with equilibrium dialysis being considered the gold standard (Vuignier et al. 2013).

Important In most therapeutic situations, the free fraction is relatively constant in the range of concentrations obtained in therapeutic scenarios, that is, independent of the drug concentration (Rowland and Tozer 1994; Toutain and Bousquet-­ Mélou 2002). This results from the rather high KD values associated with the drug-protein complex formation equilibrium (considering that an appreciable free fraction exists for drugs, even in the presence of a significant molar excess of plasma proteins, which denotes a high dissociation constant). In such circumstances, Eq. 4.7 takes the following approximate form: fu 

KD Bmax  K D

(4.8)

This is especially true for drugs that bind to albumin due to the molar excess of albumin in comparison with the total plasma concentration range

69

4  Drug Distribution

customary for most drugs (e.g., < 25  μM). In contrast, the free fraction of drugs binding to AAG is generally observed to increase with increasing total drug concentration in the therapeutic range of many drugs (Trainor 2007). Furthermore, drugs that bind to AAG are more prone to significant inter- and intra-individual variations in their free fractions. Drugs whose free fraction is independent of the drug concentration in the therapeutic range exhibit linear binding behavior. However, some drugs display nonlinear behavior within the therapeutic range. For instance, valproic acid (an anticonvulsant drug that is also used to prevent migraines and to treat manic or mixed episodes associated with bipolar disorder) may achieve total plasma levels above the plasma albumin concentration. Table 4.4 displays the therapeutic plasma levels of different drugs, along with their approximate bound fractions and binding partners. Only drugs whose therapeutic levels are similar to those of their binding partner are subjected to nonlinear binding behavior. If the reader notes that appreciable free plasma levels of the drugs are observed despite the high levels of plasma proteins, high KD values of the drug-protein complex can also be envisioned. Table 4.4  Some examples of drugs with linear and nonlinear behaviors in their binding to plasma proteins. Whether binding follows a linear or nonlinear behavior essentially depends on the therapeutic range of the drug and the physiological concentration of its binding partner. Drugs that bind to AAG exhibit nonlinear behavior at lower therapeutic concentrations. Note that the information in the table may vary significantly in physiological and pathological conditions where the plasma protein levels are modified or increased concentrations of endogenous displacing agents (e.g., fatty acids and bilirubin) occur Therapeutic Drug range (μM)a Carbamazepine 21.2–50.8

Bound fractionb 0.76

Digoxin

0.25

0.001–0.003

Lidocaine

6.4–21.3

Tacrolimus

0.004–0.025

Theophylline Valproic acid

55.5–111.0 346.7–693.4

Binding partner/s Albumin (70–80%) AAG (20–30%) Albumin

0.60–0.80 AAG (40–60%), Albumin (20%) 0.99 Albumin (44–57%) AAG (39%) 0.40 Albumin 0.80–0.90 Albumin

Nonlinear behavior? Yes (AAG) (MacKichan and Zola 1984)

No (Evered 1972; Tillement et al. 1980) No (Routledge et al. 1985)

Nob

No (Buss et al. 1985) Yes, only significant for the highest therapeutic concentrations (Gugler and Mueller 1978; Gómez Bellver et al. 1993)

Data extracted from MedlinePlus. Therapeutic drug levels. Reviewed May 21, 2017. Available at http://www.nlm.nih.gov/medlineplus/ency/article/003430.htm. Accessed: Aug 14, 2023. Interpretation of reported drug concentrations must consider the timing of the serum sample obtained for drug concentration determination b Data extracted from DrugBank (https://www.drugbank.ca). Accessed: May 21, 2018 a

70

A. Talevi and C. L. Bellera

4.5.1.1 Binding Partners Plasma is known to contain more than 60 different proteins, but only a few are relevant in drug binding phenomena: albumin, AAG, and, to a lesser extent, lipoproteins (some other plasma proteins might also be relevant for the pharmacokinetics of very specific drugs) (Kishore Deb et al. 2018). The main features of each of these proteins and their contributions to drug binding are discussed in the following subsections. Albumin Human serum albumin is the major component of plasma proteins (approximately 50% of the total protein pool). It is synthesized by the liver and has multiple functions; among them, it acts as a transporter protein for a diversity of physiologic compounds and metals (e.g., fatty acids and bilirubin), provides most of the blood pH buffer and antioxidant capacity, and plays a crucial role in the regulation of oncotic pressure (Fanali et al. 2012; Levitt and Levitt 2016). Structurally, it consists of a single polypeptide chain with a molecular weight around 66 kDa, which is present in the plasma of healthy individuals in concentrations similar to 600 μM (Bohnert and Gan 2013; Kishore Deb et al. 2018). This means that, even in the case of complexes with a 1:1 stoichiometry, it is uncommon that all the available binding sites in albumin get simultaneously occupied (and there are cases when the stoichiometry is far from 1:1; for instance, albumin can bind up to nine equivalents of fatty acids (Kishore Deb et al. 2018). Only relatively few drugs with very high bound fractions, which are administered at high doses, can saturate albumin under normal conditions (McElnay 1996). Hypoalbuminemia has been observed in many health conditions due to reduced biosynthesis; increased catabolism; renal, gastrointestinal, and/or renal clearance; and modified distribution due to increased capillary permeability (Levitt and Levitt 2016). Lower serum albumin concentrations can be precipitated by several factors, such as burns, neoplastic disease or age (see Table 4.4 for some physiological and pathological states that modify binding to plasma proteins). Albumin preferentially binds to acidic and neutral drugs, but it may also bind to basic drugs (e.g., diazepam) (Kishore Deb et al. 2018). It has multiple binding sites, although two appear to predominate and have been better characterized: site I (also known as the warfarin binding site) and site II (also termed the benzodiazepam binding site). Despite the large molecular size of albumin, it is not exclusively confined to the plasma compartment and can extravasate. Nevertheless, the rate at which the albumin-­drug complex extravasates is much lower than that of the free drug and extravasation of bound drugs is frequently disregarded in pharmacokinetic analysis. α1-acid glycoprotein AAG is an acidic, heavily glycosylated protein with a molecular weight around 41 kDa and with a preference for basic and neutral drugs. It shows quite variable plasma concentrations in healthy subjects (from around 9 to 30 μM) (Trainor 2007) and, as it is an acute phase protein, its concentration can substantially increase in response to health conditions (Trainor 2007; Kishore Deb et al. 2018). Owing to this natural variation and its lower plasma concentration, AAG is more prone to

4  Drug Distribution

71

participate in clinically relevant drug-drug and drug-disease interactions (Bohnert and Gan 2013). Lipoproteins Lipoproteins (very low-density lipoproteins, low-density lipoproteins, and high-­ density lipoproteins) are complexes of lipids and proteins arranged into a macromolecular structure, which act as plasma soluble transporters of lipids in circulation. Predictably, they can bind lipophilic drugs (predominantly basic and neutral compounds) (Kishore Deb et al. 2018). Although the mass concentrations of lipoproteins are similar to those of AAG, their molar concentrations are much lower because of their higher molecular weights, ranging from approximately 100 to 500  kDa. Disease state can have a significant impact on plasma lipoprotein profiles, which can eventually result in altered therapeutic outcomes (Wasan and Cassidy 1998). However, their role in drug disposition is much less studied and understood than that of albumin and AAG (Table 4.5).

4.5.2 Binding to Extravascular Tissue Elements Binding to specific receptors is a crucial step in drug pharmacodynamics. However, drug disposition is critically influenced by non-specific interactions with intravascular and extravascular tissue components. This will naturally be a critical determinant in the choice of dosing schemes; as only the unbound drug is available to engage in Table 4.5 Some physiological and disease states that may alter drug binding to different plasma proteins

Decrease Albumin Hepatic cirrhosis Age (neonate or elderly) Pregnancy End stage renal disease Cystic Fibrosis Nephrotic syndrome Acute pancreatitis α1-Acid glycoprotein Nephrotic syndrome Oral contraceptives

Lipoprotein Hyperthyroidism Trauma

Increase Exercise Schizophrenia Hypothyroidism Neurotic disorders Psychosis

Myocardial infarction Trauma Crohn’s disease Surgery Rheumatoid arthritis Celiac disease Renal insufficiency Diabetes Hypothyroidism Nephrotic syndrome

72

A. Talevi and C. L. Bellera

pharmacologically productive interactions with the pharmacological target, the delivered dose should compensate for non-specific binding to achieve effective drug levels at the site of action. Since the early reports on the binding of small molecules to plasma proteins in the 1930s, plasma protein binding has attracted much interest from the scientific community; however, mostly due to methodological limitations, binding to extravascular tissues was neglected for several decades, until the 1970s (Bickel 1980; Fichtl et  al. 1991). As noted by Goldstein (1949) in those days “The difficulties inherent in studying interactions with the proteins of fixed tissues present a sharp contrast to the readiness with which those of the plasma lend themselves to investigation.”. Although this situation has shown real improvements, many of the current methods used to characterize tissue binding still have substantial limitations. In the context of physiologically based pharmacokinetic modelling, the affinity of drugs for a given tissue is often reflected by tissue-to-plasma partition coefficients (Kp). The identity of the binding partners at extravascular tissues greatly varies from drug to drug; however, muscle, fat, liver, kidney, and lung are some tissues or organs for which drugs often have comparatively high binding affinities (Plomp et al. 1987; Fichtl et al. 1991; Clausen and Bickel 1993; Kishore Deb et al. 2018; Björkman 2002). In addition to the affinity for the extravascular binding sites, the relative abundance of the binding partners outside the vascular compartment must also be considered. 98% of body protein is found in the extravascular space (Bickel 1980), which contributes to proteins being among the major tissular binding partners. For instance, cardiac glycosides and other drugs bind to contractile proteins in the heart (Fichtl et al. 1991); imipramine and quinidine exhibit affinity for cytosolic ligandin (McElnay 1996). Polyaromatic drugs containing planar fused ring systems, such as anthracyclines, usually bind strongly to DNA (Bischoff and Hoffmann 2002). The affinity of chloroquine and its analogs for melanin has already been described. Adipose tissue should be considered an important potential storage tissue, since it contains a very high percentage of stored lipid in the form of triglycerides, and, together with muscle, it has a comparatively large mass in mammals and can sequester large amounts of lipophilic drugs (Bickel 1984). This should be considered to adjust doses in obese patients and after significant weight loss (May et al. 2020; Niederberger and Parnham 2021). Drugs can also partition into cell membranes or become trapped within organelles owing to pH or electro chemical gradients. In particular, basic drugs accumulate in lysosomes, endosomes, the Golgi apparatus, and secretory vesicles, owing to the low intra-organelle pH (one to two units below the pH in the cytosol) (Vuignier et al. 2010). Upon entering the organelle, the drugs become protonated and are thus less likely to return to the cytosol, which can be exploited therapeutically. For instance, chloroquine exerts its therapeutic effects in the lysosomes of parasites of the Plasmodium genre).

4  Drug Distribution

73

4.6 Pharmacokinetic Impact of Binding to Plasma Proteins and Extravascular Tissues Drug binding to plasma proteins and extravascular tissue may affect drug kinetics by acting at three levels: absorption, distribution, and elimination (McElnay 1996). Regarding absorption, via their influence on maintaining a high concentration gradient of free drug between the absorption site and plasma, both plasma and tissue binding will reinforce the sink condition, thus favoring more rapid absorption of drugs. Plasma and tissue protein binding have opposing effects on drug distribution. Because plasma proteins are largely retained within the plasma compartment, drugs that are highly bound to plasma proteins tend to have low Vds (warfarin is an archetypical example). In contrast, drugs that are highly bound to tissues tend to have high Vds (especially if the tissue to which the drug is bound is abundant in the body). For drugs that are highly bound to both the plasma and tissues, the value of Vd depends on the relative binding at both sites. In the case of tricyclic antidepressants, for instance, nearly all the drug in the plasma is bound to albumin; however, due to extensive tissue binding, only a small fraction of the total drug in the body remains in the vascular space (Koch-Weser and Sellers 1976). Let us now discuss the influence of tissue and plasma protein binding on the kinetics of elimination. If a drug is highly bound to tissues, it will be protected from elimination, as the drug outside the vascular compartment is not being delivered to the elimination organs (mainly the liver and kidneys). Thus, the tissue-bound drug acts as a drug reservoir that replenishes drug in plasma as the drug is removed from the plasma compartment by elimination mechanisms, thus increasing elimination half-lives. On the other hand, for those drugs with low extraction ratios in the elimination organs, plasma protein binding will protect the drug molecules from elimination since only the unbound drug will we uptaken by such organs. This situation is mostly observed for those drugs that move from the serum to the elimination organs through passive processes (e.g., by passive uptake to the biotransformation sites in the hepatocytes). For these drugs, the duration of action directly correlates with the bound fraction. In contrast, for drugs highly extracted by the elimination organs, protein binding has the opposite effect (Koch-Weser and Sellers 1976; McElnay 1996): by keeping more drug in the blood and increasing its delivery to the highly perfused elimination organs, protein binding will tend to shorten the duration of drug action. This is typically the case for drugs that exhibit high affinity for elimination systems, including metabolizing enzymes and drug transporters that promote uptake or excretion by eliminating organs. In such cases, eliminated free drug molecules are rapidly replaced by additional free drug molecules that result from the dissociation of the drug-protein complex. This topic will be resumed in those chapters dedicated to elimination processes and drug transporters.

74

A. Talevi and C. L. Bellera

Important For drugs with low extraction ratios in elimination organs, both plasma protein binding and tissue binding protect the drug from elimination processes, contributing to a long duration of action. Conversely, for those drugs with high extraction rations in the liver and/or the kidneys, plasma protein binding conspires against the duration of action, promoting fast elimination.

4.7 Some Therapeutic Considerations Related to Displacement from Plasma Protein and Tissue Binding Sites Displacement of a drug from its binding sites (either in plasma proteins or tissue components) by competition with other drugs or endogenous compounds has long been a concern among the healthcare community because of the risk of adverse drug reactions. The equilibria involved in such competition are represented in Fig. 4.9, and the effect of displacement on the total plasma drug concentration and the free drug concentration in the vascular and extravascular compartments can be easily predicted by applying the Equilibrium Law. The potential relevance of such interactions is vividly illustrated by the well-known interactions between tolbutamide and sulfonamides. The clinical significance of drug-drug interactions has been a subject

Fig. 4.9  Equilibria describing displacement of a drug from a binding protein by another exogenous (e.g., a drug) or endogenous compound. The displacing agent is also termed “perpetrator”, and the displaced agents is also termed “victim”

75

4  Drug Distribution

of much scientific debate (Sellers 1979; MacKichan 1989; Rolan 1994; Benet and Hoener 2002), and it is now generally accepted that any increment in the free drug plasma concentrations tends to be transient. Examples of relevant interactions involving displacement from protein binding, in general, also involve additional drug-drug interaction mechanisms occurring simultaneously (in particular, inhibition of saturable eliminating systems, such as metabolizing enzymes). There are two reasons for this finding. First, the transient increase in unbound drug concentrations due to displacement will be mitigated by diffusion of part of the sudden excess of free drug to the extravascular compartment (also note that the extravascular tissues usually have a large sink capacity). Second, immediately after displacement occurs, higher amounts of free drug are presented to the elimination organs. If linear elimination is observed, the increment of unbound drug will be rapidly compensated for by an increased elimination rate. In other words, elimination and distribution to the extravascular space buffer the initial excess of free drug molecules in the vascular compartment. For all practical purposes, the (mean) unbound levels of the displaced drug will soon return to the pre-displacement levels (Rolan 1994). The total drug concentration in plasma decreases due to the redistribution of a fraction of the displaced drug to extravascular tissues, and the apparent volume of distribution tends to increase. Therefore, drug displacement from plasma proteins is likely to be a concern only when several conditions are simultaneously met (Rolan 1994): a) the displaced drug has a low Vd, which indicates a reduced sink capacity of the extravascular tissues; b) the extraction ratio of the drug by eliminating organs is low, so that the increase in free drug concentrations is not fully compensated by an increase in the elimination rate; c) the displaced agent has a narrow therapeutic window, so that a small increase in the free drug levels at the site of action or at other tissues may result in significant toxicity; and d) the displaced agent has a high bound fraction (> 0.9), so that the displacement results in a relatively high increment in free drug levels (Table 4.6 illustrates this effect). Condition b) may occur if the victim (displaced) drug has low intrinsic clearance or if the metabolism of the victim drug is inhibited by the displacing agent. For example, Table 4.6  Effect of drug displacement on the transient free drug levels, depending on the bound fraction  (example). Displacement of highly bound drugs is more relevant from the clinical viewpoint, since it results in a more significant relative increase in the transient free drug levels, which would be later compensated by distribution to the extravascular compartment and elimination mechanisms

Drug A Bound Free Drug B Bound Free

% Drug before displacement

% Drug after displacement

% Increase in free drug concentration

95 5

90 10

+100

50 50

45 55

+10

76

A. Talevi and C. L. Bellera

salicylates and valproate can mutually displace from plasma protein binding, which could increase valproate toxicity, as salicylates also inhibit its β-oxidation (Goulden et al. 1987; Bohnert and Gan 2013). This interaction is more likely to occur with high or prolonged doses of salicylate. A scenario in which the displacement could be dangerous involves fast intravenous infusion of the displacing agent. The initial dramatic increase in the displaced agent unbound concentrations in plasma could translate to higher occupancy of its target(s) despite compensating mechanisms, especially in highly perfused organs (McElnay 1996). A well-known example is the displacement of bilirubin by sulfonamides in neonates). Finally, it is worth mentioning that when therapeutic drug monitoring is performed, clinicians should always be aware that a displacement has occurred and to what extent, in order to avoid unnecessary dose adjustment on the basis of total plasma concentrations of the displaced agents (which fall upon displacement despite the unbound, and therapeutically relevant, drug levels are in most situations not substantially modified) (Rolan 1994). Important Most drug-drug interactions involving drug displacement from plasma proteins are not clinically relevant.

Inversely, displacement of drugs from tissue-binding sites triggers the redistribution of the drug from the tissues into the plasma compartment, thus increasing drug concentrations in the intravascular space (McElnay 1996). The plasma compartment is relatively small in comparison to the extravascular tissue compartment and may not buffer these changes as efficiently at the extravascular tissues do upon displacement of the drug from plasma protein binding. Therefore, a clinically relevant sequela of such a displacement may be more likely. If plasma clearance remains unchanged, the increased elimination rate of the drug will compensate the transiently increased plasma level, which eventually will return to the predisplacement values. The diminished Vd results in a decrease in the elimination half-life, which in turn could produce a greater fluctuation in plasma levels (in the context of multiple dosing). As occurs with displacement from plasma protein binding, a major drug interaction could occur if the displaced drug has a narrow therapeutic window, and the perpetrating drug also inhibits the elimination of the displaced agent. In that case, elimination will not compensate for the increase in the displaced drug plasma concentrations.

4  Drug Distribution

77

References Aarons L (2005) Physiologically based pharmacokinetic modelling: a sound mechanistic basis is needed. Br J Clin Pharmacol 60:581–583 Benet LZ, Hoener BA (2002) Changes in plasma protein binding have little clinical relevance. Clin Pharmacol Ther 71:115–121 Bickel MH (1980) Tissue binding of drugs. Acta Pharm Suec 17:73 Bickel MH (1984) The role of adipose tissue in the distribution and storage of drugs. Prog Drug Res 28:273–303 Bischoff G, Hoffmann S (2002) DNA-binding of drugs used in medicinal therapies. Curr Med Chem 9:312–348 Björkman S (2002) Prediction of the volume of distribution of a drug: which tissue-plasma partition coefficients are needed? J Pharm Pharmacol 54:1237–1245 Bohnert T, Gan LS (2013) Plasma protein binding: from discovery to development. J Pharm Sci 102:2953–2994 Bouvier d’Yvoire M, Prieto P, Blaauboer BJ et al (2007) Physiologically-based Kinetic Modelling (PBK Modelling): meeting the 3Rs agenda. The report and recommendations of ECVAM Workshop 63. Altern Lab Anim 35:661–671 Brown E, Hopper J Jr, Hodges JL Jr et al (1962) Red cell, plasma, and blood volume in the healthy women measured by radiochromium cell-labeling and hematocrit. J Clin Invest 41:2182–2190 Browning DJ (2014) Hydroxychloroquine and chloroquine retinopathy. Springer, New York Buss DC, Ebden P, Baker S, Routledge PA (1985) Protein binding of theophylline. Br J Clin Pharmacol 19:529–531 Calzia E, Iványi ZD, Radermacher P (2005) Determinants of blood flow and organ perfusion. In: Pinsky MR, Payen D (eds) Functional Hemodynamic Monitoring. Springer, Heidelberg Clausen J, Bickel MH (1993) Prediction of drug distribution in distribution dialysis and in vivo from binding to tissues and blood. J Pharm Sci 82:345–349 Dineen S, Schumacher P (2009) Disorders of acid-base, fluids, and electrolytes. In: Alexander L, Eastman AL, Rosenbaum DH et  al (eds) Parkland Trauma Handbook, 3rd edn. Elsevier, Amsterdam Espié P, Tytgat D, Sargentini-Maier ML et  al (2009) Physiologically based pharmacokinetics (PBPK). Drug Metab Rev 41:391–407 Evered DC (1972) The binding of digoxin by the serum proteins. Eur J Pharmacol 18:236–244 Fanali G, di Masi A, Trezza V et al (2012) Human serum albumin: from bench to bedside. Mol Asp Med 33:209–290 Fichtl B, Nieciecki VA, Walter K (1991) Tissue binding versus plasma binding of drugs: General principles and pharmacokinetic consequences. Adv Drug Res 20:117–166 Fudim M, Miller WL (2018) Calculated estimates of plasma volume in patients with chronic heart failure-comparison with measured volumes. J Card Fail 24:553–560 Goldstein A (1949) The interactions of drugs and plasma proteins. Pharmacol Rev 1:102–165 Gómez Bellver MJ, García Sánchez MJ, Alonso González AC, Santos Buelga D, Dominguez-Gil A (1993) Plasma protein binding kinetics of valproic acid over a broad dosage range: therapeutic implications. J Clin Pharm Ther 18:191–197 Goulden KJ, Dooley JM, Camfield PR et al (1987) Clinical valproate toxicity induced by acetylsalicylic acid. Neurology 37:1392–1394 Gugler R, Mueller G (1978) Plasma protein binding of valproic acid in healthy subjects and in patients with renal disease. Br J Clin Pharmacol 4:441–446 Hinderling PH (1997) Red blood cells: A neglected compartment in pharmacokinetics and pharmacodynamics. Pharmacol Rev 48:279–295 Ito S, Ando H, Ose A et al (2013) Relationship between the urinary excretion mechanisms of drugs and their physicochemical properties. J Pharm Sci 102:3294–3301

78

A. Talevi and C. L. Bellera

Kalamaridis D, DiLoretto K (2014) Drug partition in red blood cells. In: Caldwell GW, Yan Z (eds) Optimization in Drug Discovery. Springer, New York Kishore Deb P, Al-Attraqchi O, Raghu Prasad M et al (2018) Protein and tissue binding: Implication on pharmacokinetic parameters. In: Tekade RK (ed) Advances in Pharmaceutical Product Development and Research, Dosage Form Design Considerations. Academic Press, Cambridge Koch-Weser J, Sellers EM (1976) Binding of drugs to serum albumin (first of two parts). N Engl J Med 294:311–316 Korzekwa K, Radice C, Nagar S (2022) A permeability- and perfusion-based PBPK model for improved prediction of concentration-time profiles. Clin Transl Sci 15:2035–2052 Kwon Y (2002) Handbook of essentials pharmacokinetics, pharmacodynamics, and drug metabolism for industrial scientists. Kluwer Academic Publishers, New York Levitt DG, Levitt MD (2016) Human serum albumin homeostasis: a new look at the roles of synthesis, catabolism, renal and gastrointestinal excretion, and the clinical value of serum albumin measurements. Int J Gen Med 9:229–255 MacIntyre AC, Cutler DJ (1986) in  vitro binding of chloroquine to rat muscle preparations. J Pharm Sci 75:1068–1070 MacKichan JJ (1989) Protein binding drug displacement interactions fact or fiction? Clin Pharmacokinet 16:65–73 MacKichan JJ, Zola EM (1984) Determinants of carbamazepine and carbamazepine 10,11-­epoxide binding to serum protein, albumin and alpha 1-acid glycoprotein. Br J Clin Pharmacol 18:487–493 May M, Schindler C, Engeli S (2020) Modern pharmacological treatment of obese patients. Ther Adv Endocrinol Metab 11:2042018819897527 McElnay JC (1996) Drug interactions at plasma and tissue binding sites. In: D’Arcy PF, McElnay JC, Welling PG (eds) Mechanisms of drug interactions. Handbook of experimental pharmacology, vol. 122. Springer, Berlin Michaelides M, Stover NB, Francis PJ et al (2011) Retinal toxicity associated with hydroxychloroquine and chloroquine: risk factors, screening, and progression despite cessation of therapy. Arch Ophthalmol 129:30–39 Nestorov I (2003) Whole body pharmacokinetic models. Clin Pharmacokinet 42:883–908 Niederberger E, Parnham MJ (2021) The impact of diet and exercise on drug responses. Int J Mol Sci 22:7692 Plomp TA, Wiersinga WM, Van Rossum JM et al (1987) Pharmacokinetics and body distribution of amiodarone and desethylamiodarone in rats after oral administration. In Vivo 1:265–279 Recordati G, Bellini TG (2004) A definition of internal constancy and homeostasis in the context of non-equilibrium thermodynamics. Exp Physiol 89:27–38 Rolan PE (1994) Plasma protein binding displacement interactions—why are they still regarded as clinically important? Br J Clin Pharmacol 37:125–128 Routledge PA, Lazar JD, Barchowsky A, Satrgel WW, Wagner GS, Shand DG (1985) A free lignocaine index as a guide to unbound drug concentration. Br J Clin Pharmacol 20:695–698 Rowland M, Tozer TN (1994) Clinical pharmacokinetics. concepts and applications. Lippincott Williams & Wilkins, Philadelphia Sellers EM (1979) Plasma protein displacement interactions are rarely of clinical significance. Pharmacology 18:225–227 Schroeder RL, Gerber JP (2014) Chloroquine and hydroxychloroquine binding to melanin: Some possible consequences for pathologies. Toxicol Rep 1:963–968 Tillement JP, Zini R, Lecompte M, d’Athis P (1980) Binding of digitoxin, digoxin and gitoxin to human serum albumin. Eur J Durg Metab Pharmacokinet 5:129–134 Toutain PL, Bousquet-Mélou A (2002) Free drug fraction vs free drug concentration: a matter of frequent confusion. J Vet Pharmacol Therap 25:460–463 Toutain PL, Bousquet-Mélou A (2004) Volumes of distribution. J Vet Pharmacol Therap 27:441–453

4  Drug Distribution

79

Trainor GL (2007) The importance of plasma protein binding in drug discovery. Expert Opin Drug Discov 2:51–64 Vuignier K, Schappler J, Veuthey JL et al (2010) Drug–protein binding: a critical review of analytical tools. Anal Bioanal Chem 398:53–66 Vuignier K, Veuthey JL, Carrupt PA et  al (2013) Global analytical strategy to measure drug– plasma protein interactions: from high-throughput to in-depth analysis. Drug Discov Today 18:1030–1034 Wadsworth GR (1954) The blood volume of normal women. Blood 9:1205–1207 Wasan KM, Cassidy SM (1998) Role of plasma lipoproteins in modifying the biological activity of hydrophobic drugs. J Pharm Sci 87:411–424 Yang F, Liu HW, Li M et al (2012) Use of a Monte Carlo analysis within a physiologically based pharmacokinetic model to predict doxycycline residue withdrawal time in edible tissues in swine. Food Addit Contam Part A Chem Anal Control Expo Risk Assess 29:73–84 Yiengst MJ, Shock NW (1962) Blood and plasma volume in adult males. J Appl Physiol 17:195–198

Further Reading The reader is advised to complement this chapter with those that focus on drug elimination and drug-drug interactions, in this same volume. If interested in drug interactions related to the distribution process, the reader is strongly advised to read the chapter by McElnay, which is listed among the references

Chapter 5

Drug Metabolism Alan Talevi and Carolina Leticia Bellera

5.1 Introduction: Definition of Xenobiotics Through evolution, the body has developed different strategies to dispose of metabolic waste and to protect itself from exposure to potentially toxic chemicals. It is useful to differentiate chemical compounds with well-defined physiological roles from those that do not have physiological functions. The latter are generically called xenobiotics, and include most drugs, natural and industrial contaminants, other industrial compounds used for technical purposes (i.e., agrochemicals), cosmetics, food additives, food components with no physiological function, recreational and social drugs (Testa and Krämer 2006), along with their biotransformation products. There is also a tendency to view those endogenously produced compounds which are (for medical or non-medical reasons) exogenously provided at high doses as xenobiotics, achieving exposure levels above their physiological range, as xenobiotics (Jenner et al. 1981; Testa and Krämer 2006).

Definition The prefix xeno- comes from Ancient Greek and means “alien.” As in “xenophobia,” it is used to denote strangeness. Accordingly, the word xenobiotic describes a chemical entity that is found within a living organism, but has no

A. Talevi (*) · C. L. Bellera Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_5

81

82

A. Talevi and C. L. Bellera

physiological function. Xenobiotics can be regarded as potentially toxic, and the body has mechanisms to dispose of them. In contrast, physiological compounds have essential biological functions and are indispensable for the survival or well-being of the organism. Among them are water, nutrients, oxygen, minerals, dietary fiber, and vitamins, and their derivatives that are produced via anabolism and catabolism.

The mechanisms available within the body to reduce exposure to non-­ physiological compounds will also limit the bioavailability of a particular category of xenobiotics, which are therapeutic drugs. Essentially, these mechanisms can be broadly classified as: a) preventing xenobiotics from entering the bloodstream or sensitive organs (e.g., the blood-brain barrier limits the distribution of many xenobiotics to the brain); b) physically removing xenobiotic molecules (mainly through bile and urine, but also, sometimes, through perspiration and respiration); and c) metabolizing xenobiotics (biochemical transformation or biotransformation) to produce drug metabolites whose excretion is faster than that of the parent (unchanged) compound.

Important Although drug metabolism and excretion are presented separately, they are highly integrated phenomena. As will be extensively discussed throughout this and the next chapter, metabolism and excretion work in a coordinated manner to provide more efficient disposal of xenobiotics. The ultimate objective of all body elimination systems is to promote the excretion of xenobiotic molecules (either unchanged or as metabolites). The major routes of excretion are bile (those drugs or drug derivatives excreted though bile are eventually excreted in faeces) and urine. Globally speaking, biotransformation processes transform a drug into a considerably more polar biochemical product. Such polar metabolites are not likely to be reabsorbed in either the intestine or renal tubules. Furthermore, they can occasionally be actively secreted into the bile or tubular content. Some textbooks point to enhanced water solubility of drug metabolites as the origin of more efficient excretion. This is only indirectly true because, in general, the drug concentrations in biological fluids are much lower than their solubility. Improved hydrophilicity increases the excretion rate of the metabolite (in comparison with the unchanged drug) because highly polar metabolites are not efficiently reabsorbed or distributed (they experience much greater diffusion barriers) and, on some occasions, are actively secreted into the intestinal, bile canalicular, or tubular lumen.

5  Drug Metabolism

83

It emerges from the previous two chapters that proper absorption of most drugs (with the possible exception of those that experience facilitated diffusion or active uptake) requires an adequate lipophilic-hydrophilic balance. They must display a certain minimal solubility in physiological media (only molecules in solution are transferred through biological barriers) but they must also possess some lipophilicity so that they can passively permeate to the bloodstream and then move forward to the remaining bodily compartments. Furthermore, lipophilic drugs tend to display a higher affinity for their pharmacological targets. On the one hand, the formation of the drug-target complex implies desolvation, and desolvation free energies, which are higher for polar drugs, oppose the binding event (Kolár et al. 2011). Lipophilic substituents are often used to exploit hydrophobic binding pockets in the target. The same requisites that a drug must generally fulfill to be absorbed, distributed, and interact with its molecular target conspire against excretion, as lipophilic drugs tend to be extensively reabsorbed at the gut and/or the renal tubules. Accordingly, biotransformation into polar metabolites is often required to minimize reabsorption and enhance drug excretion. Moreover, many metabolites are actively secreted into bile or urine by specialized transporters. Often (but far from always), drug metabolism produces pharmacologically inactive metabolites.

5.2 Metabolism: Biotransformation Reactions Definition Drug metabolism or biotransformation involves enzyme-mediated conversion of drugs into drug metabolites (the products of a biotransformation reaction). Drugs are usually metabolized sequentially. The parent (unchanged) drug (i.e., the pharmacologically active chemical entity that has been administered in a dosage form) is initially converted to a primary metabolite, which may be a substrate for a second biotransformation that produces a secondary (or sequential) metabolite (Smith 2008), and so on. Biotransformation reactions can be broadly classified into functionalization reactions (historically known as Phase I reactions), and conjugation or synthetic reactions (also known as Phase II reactions). Functionalization reactions involve the creation of a functional group or modification of an existing group. This type of biotransformation introduces a “chemical handle” or anchoring point in its substrate so that the resulting metabolite is more prone to experience a synthetic reaction (Testa and Krämer 2007; Talevi 2016). These include redox reactions and hydrolyses/hydrations (Testa and Krämer 2006). Synthetic reactions involve coupling the drug or one of its phase I metabolites with a diversity of (endogenous) moieties (e.g., glucuronic acid, glutathione, sulfate, phosphate, amino acids, acetyl, etc.), generally resulting in metabolites with a drastic increase in polarity, as well as a (moderately) higher molecular weight, which are not likely to be reabsorbed.

84

A. Talevi and C. L. Bellera

Sequential metabolism forms the basis for the common nomenclature of Phase I and Phase II metabolism, although a Phase II reaction may occur in a drug without a precedent Phase I reaction if the drug has a suitable anchoring site to which the corresponding moiety is transferred. In other words, conjugation reactions can produce first- and later-generation metabolites (Testa and Krämer 2008). Accordingly, Phase I and Phase II nomenclature may be misleading in the sense that it may be wrongly understood that Phase II biotransformations always occur to Phase I metabolites. Furthermore, a drug molecule often undergoes multiple successive Phase I reactions before a Phase II reaction. An analysis by Testa et al. (2012) of the biotransformation reactions to more than 1100 xenobiotics showed that first-­generation metabolites are formed mainly (almost 70%) by redox reactions, but about 22% are formed by conjugations. In the second generation, the contribution of redox reactions decreased to approximately 50%, whereas that of conjugations increased to 37%. In the third and later generations, both redox and conjugation reactions accounted for the same proportion (46%) of metabolites. The proportion of metabolites generated by hydrolysis did not vary significantly from one generation to another, remaining in the 8–12% range. The previous analysis demonstrates the inadequacy of the Phase I/Phase II classification, which assumes the metabolism of xenobiotics to begin with redox or hydrolysis reactions, followed by subsequent metabolic steps by conjugation. Other arguments against the traditional distinction between Phase I and Phase II reactions invoke the lack of a mechanistic basis for such classification (Josephy et al. 2005). The possible fate of xenobiotics after entering the body is illustrated in Fig. 5.1. The metabolism of acetaminophen (Fig. 5.2) is used as an example to provide the reader with a concrete idea of the complexity of the elimination processes and how Phase I and Phase II reactions may or may not be coupled. Drug metabolism involves chemical modifications of drugs. In addition to changes in molecular weight and polarity, the resulting metabolites are usually

Fig. 5.1  The fate of xenobiotics in the body may depend on different elimination pathways. Unchanged xenobiotic molecules may be directly excreted. Alternatively, they may be excreted as a Phase I or II metabolite. Synthetic reactions could occur directly on the parent compound, or may require a previous Phase I reaction to occur, which predisposes the drug to a conjugation. Often, all of these possibilities occur in parallel to different drug molecules and contribute to the overall elimination of the drug

5  Drug Metabolism

85

Fig. 5.2  Elimination pathways for acetaminophen/paracetamol. A small fraction of the dose is eliminated with no changes in urine. The drug undergoes direct conjugation with glucuronide and sulfate or sequential CYP450-mediated biotransformation to the highly toxic N-acetyl-p-­ benzoquinone imine (NAPQI) which is inactivated by conjugation with glutathione. If not inactivated, NAPQI reacts with hepatic proteins resulting in acute hepatic failure

86

A. Talevi and C. L. Bellera

inactivated from a pharmacological viewpoint. Occasionally, however, they can be pharmacologically active (and even more active than the parent compound) or, if excessively reactive from a chemical perspective, they may be toxic.

5.3 Functionalization (Phase I) Reactions Functionalization reactions involve the introduction of a “chemical handle” or anchoring point to their substrates, either by introducing a functional group, such as such as –OH,–NH2, or–SH, or by modifying an existing one. The enzymes that catalyze these reactions may be oxygenases/oxidases (such as cytochrome P450 (CYP450) isozymes, flavin-containing monoxygenases, monoamine oxidases, peroxidases, xanthine oxidases, alcohol oxidases, and others), reductases (such as aldo-­ keto reductases), or hydrolytic enzymes (e.g., esterases and amidases). Notably, some Phase I enzymes can catalyze both oxidation and reduction reactions (Testa and Krämer 2007). It has been estimated that redox reactions account for approximately 60% of all metabolic reactions in xenobiotics (Testa et al. 2012). Regarding the subcellular localization of Phase I reactions, a significant fraction of the corresponding enzymes are membrane proteins located in the smooth endoplasmic reticulum (e.g., CYP450, flavin-containing monooxygenases, some esterases, and epoxide hydrolases) (Cribb et al. 2005). The lumen of the endoplasmic reticulum has an oxidizing environment relative to the cytosol (the glutathione/glutathione disulfide ratio is approximately 3 compared to 100 in the cytoplasm). This balance appears to be crucial to the process of oxidative protein folding within the reticulum and may contribute to the generation of oxidative stress using oxygen as a terminal electron acceptor (Tu and Weissman 2004). It also creates an environment that favors the further oxidation of reactive intermediates that enter the endoplasmic reticulum lumen. To a lesser extent, some Phase I enzymes are also located in the mitochondria (e.g., monoamine oxidase) and the cytosol (e.g., xanthine oxidase or alcohol dehydrogenase). Important The liver is the organ that expresses the highest levels of metabolizing enzymes. However, some metabolic enzymes are also extensively expressed in other organs (and, in some specific cases, at higher levels than in the liver itself) (Jennen et al. 2010). In addition to the liver, other organs that significantly contribute to xenobiotic metabolism are the small intestine, lungs, and kidneys. Note that all of these are involved in matter exchange with the environment; accordingly, their involvement in the biotransformation of xenobiotics does not lack evolutionary logic. It is also important to note that the gut microbiota can also significantly contribute to drug metabolism and have an impact on drug elimination kinetics (Swanson 2015).

5  Drug Metabolism Table 5.1  Some examples of Phase I biotransformation reactions

Table 5.2  Examples of bioactivation of halogenated solvents and their associated toxicity

87 Reaction Hydroxylation Sulfooxidation Dehydrogenation Reduction Hydrolysis

Halogenated solvent Bromobenzene Vinyl chloride Carbon tetrachloride Chloroform

Examples Aromatic hydrocarbons (R-Ar) Disulfides (R-S-R) Alcohols (R-OH) Nitro compounds (R-NO2) Esters (R-COO-R’)

Toxicity Hepatic necrosis Liver cancer Hepatic necrosis, Renal necrosis Hepatic necrosis, Renal necrosis

Some examples of functionalization reactions can be found in Table 5.1. A key feature of Phase I reactions is that, while intending to contribute to xenobiotic detoxification, they may play a significant role in the etiology of some diseases and metabolic-mediated toxification processes by resulting in the formation of electrophilic intermediates capable of covalently modifying biological macromolecules (adduct formation), typically proteins or DNA, a process known as drug bioactivation (Tang and Lu 2010; Matias et al. 2014; Gan et al. 2016). Note, though, that mechanistic pathways of toxicity pathways for drugs that form adducts are not fully understood yet, and that adduct formation may coexist with adduct-­independent pathways. A relevant example of drug bioactivation will be discussed later when describing the molecular basis of paracetamol toxicity. Moreover, the moieties responsible for toxification after exposure to some halogenated solvents are presented in Table  5.2. After scrutinizing xenobiotic metabolites, Testa et  al. (2012) concluded that approximately 7% are toxic, mostly because of their strong electrophilicity. Toxic compounds usually arise from Csp2 and Csp3 oxidations (e.g., epoxide formation), N- and S-oxidation, and, above all, from the generation of quinones and analogs (quinonimines, quinonimides, and quinone-diimines), which account for approximately 40% of all toxic and/or reactive metabolites. The major enzyme systems involved in bioactivation are the CYP450 and peroxidase systems (Walsh and Miwa 2011). Furthermore, it is relatively frequent for Phase I metabolites to retain pharmacological activity (occasionally, a metabolite could display more potency than the intact drug) (Talevi 2016). Pharmacologically active metabolites are primarily first-­ generation biotransformation products. C-hydroxylation and hydrolysis play a predominant role in the generation of pharmacologically active metabolites (Testa et al. 2012). The pharmacological significance of alkyl and aryl hydroxylations is mainly due to the hydroxylated metabolite retaining the target affinity and thus the pharmacological activity of the parent compound. Note that Phase I metabolites are chemically similar to the parent compound, and a similar activity profile to that of the

88

A. Talevi and C. L. Bellera

intact drug might be attained if the modification is introduced in a non-­ pharmacophoric group or when it leads to optimization of binding to the molecular target (Fura 2006). Phase I reactions are extensively exploited in the drug discovery field to design prodrugs, soft drugs and codrugs.

Definition A prodrug is an intrinsically inactive (or scarcely active) compound that within the body is converted (enzymatically or chemically) to a pharmacologically active molecule (Ettmayer et al. 2004). Prodrugs may be advantageous due to improved absorption or safety, improved bioavailability at the site of action or to mask unpleasant organoleptic properties. There are many cases of unintended prodrugs that were not designed as such and were recognized as prodrugs only late in their development or even after approval. Many active Phase I metabolites have been marketed as distinct drugs and are used in specific therapeutic scenarios. For instance, benzodiazepines such as nordiazepam, temazepam, and oxazepam, are Phase I metabolites of diazepam. Diazepam is biotransformed into nordiazepam and temazepam by CYP2C19 and CYP3A4, respectively, which are hydroxylated and demethylated to oxazepam (Luk et  al. 2014). Oxazepam may be safer than other benzodiazepines in patients with impaired liver function or in the elderly, because it does not require hepatic oxidation and is directly metabolized by glucuronidation. Codrugs (or mutal prodrugs) comprise two active compounds that are coupled together and reciprocally act as promoieties for each other (Aljuffali et  al. 2016). Upon biotransformation, both active molecules are released simultaneously. For instance, the antibiotic codrug sultamicillin breaks into the β-lactam antibiotic ampicillin and the β-lactamase inhibitor penicillanic acid. Soft drugs (or antedrugs) are short-lived therapeutic agents that undergo predictable metabolism to inactive metabolites immediately after exerting their therapeutic effect (Rautio et al. 2018); their design should consider not only the molecular determinants of activity but also structural features that assure prompt deactivation. For example, clevidipine butyrate is an ultrashort-­ acting calcium-channel blocker that is rapidly hydrolyzed into an inactive metabolite after intravenous administration, with a half-life of approximately 1 min.

5  Drug Metabolism

89

Important Occasionally, Phase I metabolites retain the pharmacological activity of the parent compound. On some occasions, they may display higher activity than the parent compound. The bioactivation to toxic metabolites by Phase I reactions is also possible.

5.3.1 Cytochrome P450 CYP450 comprises a superfamily of hemoproteins (that is, they contain the organic cofactor heme, which is a prosthetic group crucial for their catalytic activity). This enzymatic system is arguably the most important Phase I xenobiotic metabolizing system, participating in the elimination of 70–85% of the known drugs. Furthermore, CYP450 is also involved in the metabolism of endogenous compounds, including biosynthesis, bioactivation, and, sometimes, the breakdown of steroids (sex hormones, neurosteroids, cholesterol, and vitamin D) and fatty acids (Gibbons 2002; Jones et al. 2014; Westphal et al. 2015). 57 individual P450s have been characterized in humans (Lewis 2004), and they are distributed across 18 families. In mammals, CYP450 members are located either in the endoplasmic reticulum or inner mitochondrial membrane. Mutations or other defects in genes encoding CYP450 members (especially those characterized by low redundancy and narrow substrate specificity) result in P450-mediated diseases, including those caused by aberrant steroidogenesis; defects in fatty acid, cholesterol, and bile acid pathways; and vitamin D and retinoid dysregulation (Nebert et al. 2013). Furthermore, CYP members from microorganisms provide the basis for the development of anti-infective therapies (e.g., antifungal and antiprotozoal drugs) (Lepesheva and Waterman 2011; Lepesheva et al. 2018). Important Monooxygenation is undoubtedly the most common reaction catalyzed by CYP450. It involves the insertion of one atom of oxygen into an organic substrate (RH), while the remaining oxygen atom from an oxygen molecule is reduced to water:

RH  O2  NADPH  H   ROH  H 2 O  NADP 

(5.1)

However, other less common reactions are catalyzed by CYP450, including reductions, ester cleavage, and ring expansions (Guengerich 2001). The liver, kidneys, intestine, lungs, and skin have the highest expression levels of CYP450. Since each CYP450 member has its own substrate specificity, it is not difficult to imagine that, jointly, this superfamily confers great versatility for the biotransformation of a wide range of xenobiotics. Furthermore, some of its members display wide substrate specificity (i.e., they are polyspecific): in particular, CYP3A4

90

A. Talevi and C. L. Bellera

Fig. 5.3  Comparison of the hepatic expression levels of individual P450 isoforms. (Humans, data extracted from Kwon 2002)

is the most promiscuous member of the superfamily, catalyzing the Phase I biotransformation of hundreds of known drugs (Bu 2006). Figure 5.3 displays the relative abundance of different CYP450 isozymes in the liver. Figure 5.4 shows the approximate percentage of drugs in the market whose metabolism is mediated by CYP450 members. Note that the rate at which a given member catalyzes the conversion of a substrate depends on the turnover number (the maximum number of chemical conversions of substrate molecules per second that a single catalytic site will execute), the number of available enzyme copies, and the affinity of the substrate for the enzyme. Consequently, there are some isozymes (for example, CYP2D6) which have a remarkable role in xenobiotic metabolism despite being expressed at comparatively low expression levels. Besides the monoxygenases, the CYP450 includes “auxiliary” enzymes that contribute with the electrons required to the redox reaction: NADPH CYP450 reductase and cytochrome b5, which in turn may be reduced by NADPH cytochrome reductase and NADH cytochrome b5 reductase (Gan et al. 2009). The most commonly proposed mechanism of catalysis is (Meunier et al. 2004; Guengerich 2007): (a) binding or the substrate in the vicinity of the heme group, which induces a change in the conformation of the active site, often displacing a

5  Drug Metabolism

91

Fig. 5.4  Estimated percent of marketed drugs metabolized different CYPs. (Data extracted from Kwon 2002)

water molecule from the distal axial coordination position of the heme iron; (b) substrate binding induces electron transfer from NADPH via CYP450 reductase or another associated reductase; (c) molecular oxygen binds to the resulting ferrous heme center at the distal axial coordination position, initially giving a dioxygen adduct; (d) a second electron is transferred from either cytochrome P450 reductase, cytochrome b5, or ferredoxins, reducing the Fe-O2 adduct to give a short-lived peroxo state; and (e) the peroxo group formed in step d is rapidly protonated twice, releasing one molecule of water and forming a highly reactive species (an iron(IV) oxo or ferryl species with an additional oxidizing equivalent, and the remaining oxygen atom is transferred to the substrate. After the product is released from the active site, the enzyme returns to its original state, with a water molecule occupying the distal coordination position of the iron nucleus (Fig. 5.5). The names of the CYP450 enzymes are given with a number-letter-number designation (for example, CYP3A4), with the CYP abbreviation indicating CYP450 membership, the first number indicating the P450 family based on 40% or greater sequence identity, the letter indicating the P450 subfamily based on 55% or greater sequence identity, and the final number representing the individual P450 within that family/subfamily. In humans, CYP3A4, CYP2D6, CYP2C9, CYP1A2, and CYP2C19 play major roles in drug metabolism, with some contributions from CYP2E1, CYP2A6, CYP2C8, and CYP2B6 (Furge and Guenguerich 2006). CYP450 stability of a drug is frequently studied in vitro using hepatocytes or liver microsomes, and by incubating the test drug with different recombinant

92

A. Talevi and C. L. Bellera

Fig. 5.5  Generalized catalytic cycle for CYPP450 reactions

enzymes expressing specific and pharmacokinetically relevant CYP enzymes (Butler and Riley 2022).

Definition Microsomes are vesicle-like artifacts that re-form from pieces of the endoplasmic reticulum when eukaryotic cells (most frequently hepatocytes) are broken up in the laboratory (Wilcock and Ward 2022). They are important tools in drug discovery, and they are widely available commercially. They can be concentrated and separated from other cellular debris using differential centrifugation. Unbroken cells, nuclei, and mitochondria sediment-out at 9000–10,000  g, whereas soluble enzymes and fragmented ER, which contains CYP450, remain in solution. At 100,000 × g, achieved by faster centrifuge rotation, the endoplasmic reticulum components sediment out and the soluble enzymes remain in the supernatant (called the cytosolic fraction). The S9 fraction can be obtained after the first centrifugation step at approximately 9000 × g. It contains cytosolic and microsomal enzymes. Incubation of test compounds with hepatic microsomal preparations is the primary method by which phase I biotransformations are determined. For microsomal stability determination, the compound is typically incubated for 30  min in approximately 1.0  mg/mL microsomal protein phosphate buffer pH 7.4 at 37 °C. The compound stability is then determined using an appropriate analytical method. However, it should be noted that microsomes cannot be used to study enzyme induction phenomena (see later in this chapter), since the latter requires functional whole-cell machinery.

93

5  Drug Metabolism

5.4 Synthetic (Phase II) Reactions In Phase II reactions, a chemical compound is conjugated with (generally polar) endogenous moieties (the endogenous conjugating moiety, sometimes abbreviated as the endocon), such as glucuronic acid, glutathione, sulfate, or acetate. A marked difference between Phase I and Phase II reactions is that the latter tend to produce pharmacologically inactive and chemically non-reactive products. Only approximately 10% of known toxic metabolites originate from conjugation reactions (Testa et al. 2012). Phase II metabolites are usually pharmacologically inactive (a typical counterexample is morphine 6-O-glucuronide). We will then say that, generally, Phase II reactions inactivate their substrates. Besides detoxifying reactive molecules, the addition of such polar groups through conjugation produces more hydrophilic and larger metabolites which are not able to easily diffuse across cell membranes, conditioning their reabsorption and distribution. Furthermore, many of the products from Phase II metabolism are weak acids, which are: a) nearly completely ionized at physiological pH and b) prone to interact with albumin, which also conspires against their extravasation (Table 5.3). As discussed in Chap. 6, some Phase II metabolites are so polar that they need the help of efflux transporters to leave the metabolizing cell. Conjugation reactions are catalyzed by a series of transferases. As mentioned previously, the substrate must have an appropriate “anchoring point” for conjugation to occur. Functional groups where conjugation reactions occur include carboxyl, hydroxyl, amino and sulfhydryl groups. An important aspect of Phase II reactions is that they require a co-substrate (often called a cofactor) to take place (the transferase has sometimes  been figuratively described as a “nuptial bed”). The co-substrate carries an endogenous conjugating moiety, with the chemical bond linking the cofactor and endocon being high-energy, such that the Gibbs energy released upon its cleavage drives the transfer of the endocon to the substrate (Testa and Krämer 2008). The molecular structures of the cofactors of the main Phase II reactions are shown in Fig. 5.6. Because there is a limited supply of

Table 5.3  Increase in molecular weight and pKa values of the resulting products from some common Phase II reactions. The capacity and affinity of the correspondent transferases is also listed Reaction Glucuronidation

Glycine conjugation Sulfation Glutathione conjugation Acetylation

Transferase UDP-­ Glucuronosyltransferases (UGTs) Glycine transferases

ΔMW pKa 176 3.0–3.5

Affinity low

57

3.5–4.0

intermediate intermediate

Sulfotransferases Glutathione S-transferases (GSTs) N-acetyl transferases

81 289

T confers protection effect against cisplatin-induced ototoxicity. Pharmacogenomics 16:323–332 Lavertu A, McInnes G, Daneshjou R et al (2018) Pharmacogenomics and big genomic data: from lab to clinic and back again. Hum Mol Genet 27:R72–R78

420

J. K. Roberts et al.

Leeder JS, Kearns GL (1997) Pharmacogenetics in pediatrics. Implications for practice. Pediatr Clin N Am 44:55–77 Lennard L, Van Loon JA, Lilleyman JS et al (1987) Thiopurine pharmacogenetics in leukemia: correlation of erythrocyte thiopurine methyltransferase activity and 6-thioguanine nucleotide concentrations. Clin Pharmacol Ther 41:18–25 Lennard L, Lilleyman JS, Van Loon J et al (1990) Genetic variation in response to 6-­mercaptopurine for childhood acute lymphoblastic leukaemia. Lancet 336:225–229 Liu SG, Gao C, Zhang RD et  al (2017) Polymorphisms in methotrexate transporters and their relationship to plasma methotrexate levels, toxicity of high-dose methotrexate, and outcome of pediatric acute lymphoblastic leukemia. Oncotarget 8:37761–37772 Maitland-van der Zee AH, Raaijmakers JA (2012) Variation at GLCCI1 and FCER2: one step closer to personalized asthma treatment. Pharmacogenomics 13:243–245 McLeod HL, Krynetski EY, Wilimas JA et al (1995) Higher activity of polymorphic thiopurine S-methyltransferase in erythrocytes from neonates compared to adults. Pharmacogenetics 5:281–286 Mlakar V, Huezo-Diaz Curtis P, Satyanarayana Uppugunduri CR et al (2016) Pharmacogenomics in pediatric oncology: review of gene-drug associations for clinical use. Int J Mol Sci 17:1502 Moriyama T, Nishii R, Lin TN et al (2017a) The effects of inherited NUDT15 polymorphisms on thiopurine active metabolites in Japanese children with acute lymphoblastic leukemia. Pharmacogenet Genomics 27:236–239 Moriyama T, Yang YL, Nishii R et al (2017b) Novel variants in NUDT15 and thiopurine intolerance in children with acute lymphoblastic leukemia from diverse ancestry. Blood 130:1209–1212 Mukattash TL, Nuseir KQ, Jarab AS et al (2014) Sources of information used when prescribing for children, a survey of hospital based pediatricians. Curr Clin Pharmacol 9:395–398 Murto K, Lamontagne C, McFaul C et al (2015) Celecoxib pharmacogenetics and pediatric adenotonsillectomy: a double-blinded randomized controlled study. Can J Anaesth 62:785–797 Neville KA, Becker ML, Goldman JL et al (2011) Developmental pharmacogenomics. Paediatr Anaesth 21:255–265 Palmaro A, Bissuel R, Renaud N et  al (2015) Off-label prescribing in pediatric outpatients. Pediatrics 135:49–58 PharmGKB (2023) PharmGKB FAQs. https://www.pharmgkb.org/page/faqs#what-­is-­the-­ difference-­between-­pharmacogenetics-­and-­pharmacogenomics. Accessed 18 Apr 2023 Pierre-François MJD, Gagné V, Brukner I et al (2022) Pharmacogenetic Expression of CYP2C19 in a Pediatric Population. J Pers Med 12:1383 Pussegoda K, Ross CJ, Visscher H et al (2013) Replication of TPMT and ABCC3 genetic variants highly associated with cisplatin-induced hearing loss in children. Clin Pharmacol Ther 94:243–251 Quiñones L, Roco A, Cayun JP et al (2017) Clinical applications of pharmacogenomics. Rev Med Chil 145:483–500 Ramos-Martin V, O’Connor O, Hope W (2015) Clinical pharmacology of antifungal agents in pediatrics: children are not small adults. Curr Opin Pharmacol 24:128–134 Ramsey LB, Brown JT, Vear SI et al (2020) Gene-based dose optimization in children. Annu Rev Pharmacol Toxicol 60:311–331 Ramsey LB, Prows CA, Chidambaran V et al (2023) Implementation of CYP2D6-guided opioid therapy at Cincinnati Children's Hospital Medical Center. Am J Health Syst Pharm 80:852–859 Relling MV, Gardner EE, Sandborn WJ et al (2011) Clinical Pharmacogenetics Implementation Consortium guidelines for thiopurine methyltransferase genotype and thiopurine dosing. Clin Pharmacol Ther 89:387–391 Roberts JK, Stockmann C, Constance JE et al (2014) Pharmacokinetics and pharmacodynamics of antibacterials, antifungals, and antivirals used most frequently in neonates and infants. Clin Pharmacokinet 53:581–610 Rodieux F, Daali Y, Rollason V et al (2023) Practice of CYP450 genotyping and phenotyping in children in a real-life setting. Front Pharmacol 14:1130100

17  The Relationship Between Pharmacogenomics and Pharmacokinetics…

421

Rood JM, Engels MJ, Ciarkowski SL et al (2014) Variability in compounding of oral liquids for pediatric patients: a patient safety concern. J Am Pharm Assoc 54:383–389 Ross CJ, Katzov-Eckert H, Dube MP et al (2009) Genetic variants in TPMT and COMT are associated with hearing loss in children receiving cisplatin chemotherapy. Nat Genet 41:1345–1349 Sachs AN, Avant D, Lee CS et al (2012) Pediatric information in drug product labeling. JAMA 307:1914–1915 Schmiegelow K, Nielsen SN, Frandsen TL et  al (2014) Mercaptopurine/Methotrexate maintenance therapy of childhood acute lymphoblastic leukemia: clinical facts and fiction. J Pediatr Hematol Oncol 36:503–517 Sharma S, Ellis EC, Gramignoli R et al (2013) Hepatobiliary disposition of 17-OHPC and taurocholate in fetal human hepatocytes: a comparison with adult human hepatocytes. Drug Metab Dispos 41:296–304 Škorić B, Kuzmanović M, Jovanović M et al (2023) Methotrexate concentrations and associated variability factors in high dose therapy of children with acute lymphoblastic leukemia and non-­ Hodgkin lymphoma. Pediatr Hematol Oncol 40:446–457 Stevens JC, Marsh SA, Zaya MJ et  al (2008) Developmental changes in human liver CYP2D6 expression. Drug Metab Dispos 36:1587–1593 Stockmann C, Fassl B, Gaedigk R et  al (2013) Fluticasone propionate pharmacogenetics: CYP3A4*22 polymorphism and pediatric asthma control. J Pediatr 162:1222–1227 Stockmann C, Reilly CA, Fassl B et al (2015) Effect of CYP3A5*3 on asthma control among children treated with inhaled beclomethasone. J Allergy Clin Immunol 136:505–507 Teusink A, Vinks A, Zhang K et al (2016) Genotype-directed dosing leads to optimized voriconazole levels in pediatric patients receiving hematopoietic stem cell transplantation. Biol Blood Marrow Transplant 22:482–486 Theken KN, Lee CR, Gong L et al (2020) Clinical Pharmacogenetics Implementation Consortium Guideline (CPIC) for CYP2C9 and nonsteroidal anti-inflammatory drugs. Clin Pharmacol Ther 108:191–200 United States Food and Drug Administration (US FDA). New pediatric labeling information database. https://www.accessdata.fda.gov/scripts/sda/sdNavigation.cfm?sd=labelingdatabase. Accessed 1 May 2023 Vyhmeister K, Sierra CM (2023) Codeine prescribing practices before and after the 2017 FDA warning at an academic health system. J Am Pharm Assoc 63:S20–S24 Walsh TJ, Karlsson MO, Driscoll T et  al (2004) Pharmacokinetics and safety of intravenous voriconazole in children after single- or multiple-dose administration. Antimicrob Agents Chemother 48:2166–2172 Wehry AM, Ramsey L, Dulemba SE et al (2018) Pharmacogenomic testing in child and adolescent psychiatry: an evidence-based review. Curr Probl Pediatr Adolesc Health Care 48:40–49 Wright FA, Bebawy M, O'Brien TA (2015) An analysis of the therapeutic benefits of genotyping in pediatric hematopoietic stem cell transplantation. Future Oncol 11:833–851 Xie HG (2010) Personalized immunosuppressive therapy in pediatric heart transplantation: progress, pitfalls and promises. Pharmacol Ther 126:146–158 Xu H, Robinson GW, Huang J et al (2015) Common variants in ACYP2 influence susceptibility to cisplatin-induced hearing loss. Nat Genet 47:263–266 Zhao W, Elie V, Roussey G et al (2009) Population pharmacokinetics and pharmacogenetics of tacrolimus in de novo pediatric kidney transplant recipients. Clin Pharmacol Ther 86:609–618 Zhao W, Fakhoury M, Baudouin V et  al (2013) Population pharmacokinetics and pharmacogenetics of once daily prolonged-release formulation of tacrolimus in pediatric and adolescent kidney transplant recipients. Eur J Clin Pharmacol 69:189–195

Chapter 18

Bioavailability and Bioequivalence Z. Gulsen Oner and James E. Polli

Objectives After reading this chapter, a student will be able to: –– –– –– –– ––

Describe terminology related to therapeutic equivalence Describe bioavailability/bioequivalence studies through a drug product lifecycle Describe bioavailability/bioequivalence study types Describe a pharmacokinetic study design to assess in vivo bioequivalence Describe the biopharmaceutics classification system (BCS) and BCS-based biowaiver Overview A drug product may be placed on the market only after being reviewed and approved from the competent authority. In the United States, the Food and Drug Administration (FDA) ensures that safe and effective drugs are available by reviewing and approving New Drug Applications (NDAs), as well as Abbreviated New Drug Applications (ANDAs) or generic drug applications. Most types of NDAs typically include bioavailability (BA) and bioequivalence (BE) assessment. ANDAs typically include BE assessment. The FDA uses the Biologic License Application (BLA) pathway for innovator biological products, and abbreviated licensure pathway for biosimilar biological products [denoted 351(k) pathway]. Biological products demonstrate comparability or biosimilarity. Biologics and biosimilars are not discussed here.

Z. G. Oner Department of Pharmaceutical Sciences, University of Maryland, Baltimore, MD, USA Turkish Medicines and Medical Devices Agency, Ankara, Turkey J. E. Polli (*) Department of Pharmaceutical Sciences, University of Maryland, Baltimore, MD, USA e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_18

423

424

Z. G. Oner and J. E. Polli

18.1 Introduction: Why Bioavailability and Bioequivalence Studies? A drug product may be placed on the market only after being reviewed and approved from the competent authority. In the United States, the Food and Drug Administration (FDA) ensures that safe and effective drugs are available by reviewing and approving New Drug Applications (NDAs), as well as Abbreviated New Drug Applications (ANDAs) or generic drug applications. Most types of NDAs typically include bioavailability (BA) and bioequivalence (BE) assessment. ANDAs typically include BE assessment. The FDA uses the Biologic License Application (BLA) pathway for innovator biological products, and abbreviated licensure pathway for biosimilar biological products [denoted 351(k) pathway]. Biological products demonstrate comparability or biosimilarity. Biologics and biosimilars are not discussed here. For a drug product to exhibit a therapeutic effect, the active ingredient should be delivered to its site of action in an effective concentration and for a certain period of time. BA describes this phenomenon and emphasizes the rate and extent of absorption of the active ingredient or moiety to the site of action. The rate of absorption is often not simple to describe or measure, but it is sometimes desirable to increase or delay the absorption rate. The concentration of the active ingredient at the site of action is related to the concentration in the plasma, and therefore, it is desirable to measure the amount of drug absorbed. Drug products having the same amounts of active ingredient may exhibit different therapeutic responses due to different drug plasma levels. To receive approval as a generic drug, an applicant must demonstrate that their proposed drug product is bioequivalent to the innovator drug (i.e. to the reference listed drug). BE is the absence of a significant difference in the rate and extent to which the active ingredient in two dosage forms becomes available at the site of drug action. Therefore, a generic drug product that is determined to be bioequivalent is expected to have equal safety and efficacy compared to its brand name drug product. Definitions “Approved Drug Products with Therapeutic Equivalence Evaluations,” commonly known as the “Orange Book,” (U.S.  Food & Drug Administration 2017a) is a source for drug products approved by the FDA, therapeutic equivalence (TE) evaluations, brand-name drug data, and patent and exclusivity information. Code of Federal Regulations Title 21 (21 CFR) Part 320 defines the BA and BE requirements in US. Below are some TE related terms: Pharmaceutical equivalents: Drug products that contain the same active ingredient(s), dosage form, route of administration, and amount of active ingredient, while meeting the same standards (i.e. strength, quality, purity, and identity). They may differ in characteristics such as shape, scoring

18  Bioavailability and Bioequivalence

425

configuration, release mechanisms, packaging, excipients (including colors, flavors, preservatives), expiration time, and, within certain limits, labeling. Being pharmaceutical equivalents is a requirement to be therapeutic equivalents (e.g. to be a generic to the reference listed drug). Pharmaceutical alternatives: Drug products that contain the identical therapeutic moiety, or its precursor, but not necessarily in the same amount or dosage form or as the same salt or ester (21CFR314.3). One example is tetracycline hydrochloride 250  mg capsules vs. tetracycline phosphate complex 250 mg capsules. Another example is quinidine sulfate 200 mg tablets vs. quinidine sulfate 200  mg capsules (U.S.  Food & Drug Administration 2017a). Therapeutic equivalents: Pharmaceutical equivalents for which bioequivalence has been demonstrated, and they can be expected to have the same clinical effect and safety profile when administered to patients under the conditions specified in the labeling. FDA’s criteria for therapeutically equivalent drug products are: 1. Approved as safe and effective 2. Pharmaceutical equivalents in that they a) contain identical amounts of the same active drug ingredient in the same dosage form and route of administration, and b) meet compendial or other applicable standards of strength, quality, purity, and identity 3. Bioequivalent in that a) they do not present a known or potential bioequivalence problem, and they meet an acceptable in vitro standard, or b) if they do present such a known or potential problem, they are shown to meet an appropriate bioequivalence standard 4. Adequately labeled 5. Manufactured in compliance with Current Good Manufacturing Practice regulations Bioequivalent drug products: Pharmaceutical equivalent or pharmaceutical alternative products that display comparable bioavailability when studied under similar experimental conditions (U.S.  Food and Drug Administration 2017a). Generic drug: A copy of a brand-name drug that one company makes that was developed by another company (U.S. Food and Drug Administration 2017b). Therefore, it has the following properties compared to its innovator drug: • • • •

contains the same active ingredients (inactive ingredients may vary) is identical in strength, dosage form, and route of administration has the same use indications is bioequivalent

426

Z. G. Oner and J. E. Polli

• meets the same batch requirements for identity, strength, purity, and quality • is manufactured under the same strict standards of FDA’s good manufacturing practice regulations required for innovator products (U.S. Food and Drug Administration 2017c). The terms “generic drug” and “therapeutic equivalents” mean the same. “Generic drug” is usually used by health care professionals and patients, while “therapeutic equivalents” is more commonly used by regulatory scientists. Reference listed drug: The listed drug identified by FDA as the drug product upon which an ANDA applicant relies on for seeking approval of their generic drug. Reference standard: The drug product selected by FDA that an applicant seeking approval of an ANDA must use in conducting an in vivo bioequivalence study (21CFR314.3). This newer term generally is also the reference listed drug. Bioavailability: The rate and extent to which the active ingredient or active moiety is absorbed from a drug product and becomes available at the site of drug action. Note that this definition in connection with BE testing emphasizes rate and extent (e.g. Cmax and AUC). This term “bioavailability” can also refer to a single number (see absolute bioavailability and relative bioavailability below). Bioequivalence: The absence of a significant difference in the rate and extent to which the active ingredient/moiety in pharmaceutical equivalents/alternatives becomes available at the site of drug action when administered at the same molar dose under similar conditions in an appropriately designed study (21CFR314.3). Generic substitution: The dispensing to a patient of a product classified as therapeutically equivalent. Therapeutic substitution: The dispensing to a patient of not the prescribed drug product nor a therapeutic equivalent, but a drug product which is expected to yield the same therapeutic outcome. Therapeutic substitution generally occurs in closed health system (e.g. hospital) where health care professionals agree to the use of a more limited number of drugs than is commercially available, such as the use of one statin over all other statins. Therapeutic substitution has no conceptual overlap with therapeutic equivalents (i.e. generic drugs).

18  Bioavailability and Bioequivalence

427

18.2 Bioavailability/Bioequivalence Studies Through a Drug Product Lifecycle 18.2.1 Investigational New Drug Period and New Drug Development During a new drug’s early preclinical development, the sponsor’s primary goal is to determine if the product is safe for initial use in humans and if the compound exhibits pharmacological activity that justifies commercial development. These assessments are achieved by conducting animal pharmacology and toxicology studies (U.S.  Food and Drug Administration 2015a). These assessments contribute to an Investigational New Drug (IND) application. An IND is required to conduct human clinical trials on a new molecular entity. The IND period aim to lead to submission of an NDA. After approval to begin human clinical trials, the primary goal is to determine if the product is safe and effective, which is achieved by clinical efficacy, safety, and pharmacokinetic studies. BA studies are one type of pharmacokinetic study. NDAs often include efforts to measuring in vivo BA. Several reasons can cause BA to be difficult to measure. A typical NDA has 4–6 BE studies. BE documentation can be useful during the IND period to compare: 1. early and late clinical trial formulations; 2. formulations used in clinical trials and stability studies, if different; 3. clinical trial formulations and to-be-marketed drug products, if different; and 4. product strength equivalence or proportionality (U.S.  Food & Drug Administration 2014)

18.2.2 Generic Drug Development During development of generic drugs, the sponsor’s primary goal is to determine if the proposed product is bioequivalent to the innovator. However, clinical data to establish safety and efficacy is not required, since BE is taken as the major indicator of a proposed generic to be safe and effective, like the brand. BE is often in the form of plasma profiles of proposed generic and brand being similar, although in vitro or other studies can sometimes be applied in place of plasma profiles. Similarly, any person submitting an ANDA must include evidence demonstrating that the drug product is bioequivalent to the RLD (or reference listed drug), or information that would permit FDA to waive the submission of in vivo BE evidence.

428

Z. G. Oner and J. E. Polli

Important NDA must undergo clinical trials to demonstrate safety and efficacy. Generic products must demonstrate bioequivalence with the innovator/reference product.

18.2.3 The Scale-Up and Post-approval Changes (SUPAC) After a drug product is approved for marketing, the manufacturer may want or need certain changes in the finished product (e.g. manufacturing method or location; formulation composition) or in the active substance. BE (e.g. via human in vivo plasma profiles) is required for some manufacturing changes.

18.3 Bioavailability and Bioequivalence Studies FDA has issued NDA BA and BE Draft Guidance (U.S.  Food and Drug Administration 2014) and ANDA BE Draft Guidance (U.S.  Food and Drug Administration 2013), in which study protocol details are outlined. BA depends on several factors including route of administration, dosage form, physicochemical properties of the drug(s) and excipient(s), and biologic factors.

18.3.1 Absolute and Relative Bioavailability Studies BA may be absolute or relative: Absolute bioavailability is the ratio of the BA of a given dosage form to that of the intravenous (IV) administration (100%) (e.g. oral solution vs. IV). F=

AUCA / dose A AUCIV / dose IV

(18.1)

where AUCA and dose A are the area under the concentration time curve and dose for the test drug product, respectively; and AUCIV and dose IV refer to the area under the concentration time curve and dose for the intravenous reference drug product. For example, F  =  40% if AUCA is 40% that of AUCIV, after giving the same doses. More generally, relative bioavailability is the ratio of the BA of a given dosage form to that of a form administered by the same or another non-intravenous route (e.g. tablet vs. oral solution).

18  Bioavailability and Bioequivalence

F=

429

AUCA / dose A AUCB / dose B

(18.2)

where AUCA and dose A are the area under the concentration time curve and dose for the test drug product, respectively; and AUCB and dose B refer to the area under the concentration time curve and dose for the reference drug product. Absolute BA if often determined for a drug during the IND phase of drug development. One reason why absolute BA may not be measure is lack of availability of an IV formulation. An IV formulation requires effort to formulation, including sufficient drug solubility in such a formulation. Another reason is concern about drug safety after IV administration.

18.3.2 Mass Balance Studies In the mass-balance study, the drug is administered with a radioactive label in a metabolically stable position on the drug structure. The systemic exposure of drug and its metabolites, including drug and metabolites in urine and feces, is obtained and compared to the amount of total radioactive drug administered. The absolute BA can be measured through this approach, which involves identify how much drug was excreted relative to how much drug was administered.

18.3.3 Bioequivalence Studies A BE study is basically a comparative BA study designed to establish equivalence between test and reference products. Although BA and BE are related, BE comparisons normally rely on (1) a criterion, (2) a confidence interval for the criterion, and (3) a predetermined BE limit. BE comparisons could also be used in certain pharmaceutical product line extensions, such as additional strengths (e.g. introduce a 75 mg strength, between the existing 50 mg and 100 mg), new dosage forms [e.g. change from immediate-release (IR) to extended-release], and new routes of administration (e.g. oral to topical). Approaches to determine BE generally follow approaches similar to those used for BA. The order of general preference to assess relative BA or BE is: human pharmacokinetic (PK) studies, in vitro tests predictive of human in vivo BA (in vitro-in vivo correlation), pharmacodynamic (PD) studies, studies with clinical benefit endpoints, and other in vitro studies.

430

Z. G. Oner and J. E. Polli

18.3.3.1 General Pharmacokinetic Study Design The design of a BA/BE study depends on the objectives of the study, the ability to analyze the drug in biological fluids, the pharmacodynamics of the drug substance, the route of drug administration, and the nature of the drug and drug product. General requirements for the design and conduct of BA or BE studies are summarized below. Studies are usually evaluated by a single dose, two period, two treatment, two sequence, open label, randomized crossover design, comparing equal doses of the test and reference products in fasted, adult, healthy volunteers. For example, a BE study may involve n  =  24 subjects to compare generic to brand. There are two treatments (i.e. generic and brand). There are two sequences: TR and RT. TR denotes test (or generic) is tested first in a subject, followed by reference (or brand) in that same subject. RT denotes brand (or reference) is tested first. Each subject is randomized to be in sequence TR or sequence RT. In an open label study, subjects (and immediate research staff) are not blinded but aware of what treatments are administered in each period. The study is two period, since treatments are tested over two occasions (i.e. once for generic and once for brand). This design is a non-­ replicate design, as each test and reference products are only evaluated in one period. By virtue of being a single dose study, each period only involves drug administration once, rather than, for example, every 12 h for the entire period. A single dose period can vary from several hours to several days, depending on how long it is necessary to measure drug levels in the blood to capture at least 80–90% of the area under the concentration time curve, compared to if samples were collect for an infinite length of time. Long half-life drugs require longer sampling periods. The study should be designed in such a way that the formulation effect can be distinguished from other effects. There are two ways of designing the study: crossover and parallel. A crossover design is a study where each subject receives the test drug product and reference drug product separately, following a washout period between treatments. Compared to parallel design, the crossover design provides an advantage because it reduces variability caused by subject-specific factors, thereby increasing the ability to discern differences due to formulation. An adequate washout period should be scheduled between the two periods to separate each treatment before administration of the second dose (e.g. ≥5 half-lives of the moieties to be measured). In a parallel design, subjects are divided randomly into groups, and each group receives only one treatment, and is less powerful than crossover. The most common reason to employ a parallel design is long drug half-life, such that a crossover study with a washout period would be too long in duration to successfully complete (e.g. excessive number of subject dropouts due to loss to follow up). Single dose PK studies recommended over steady-state (multiple dose) PK studies. Single dose PK studies are generally more sensitive than steady-state studies in assessing rate of release of the drug substance from the drug product into the systemic circulation. Steady-state dose study is less sensitive in detecting differences in Cmax, since PK profiles from multiple dosing typically reflect not only the most recent dose, but many of the most recent doses.

18  Bioavailability and Bioequivalence

431

Meanwhile, steady-state PK studies involve subjects, usually patients, taking the drugs many times (e.g. every 12 h) over the study period. A main reason to use multiple dose design is that the drug is too toxic for healthy volunteers, such that the BE study is only conducted in patients who can be expected to benefit from drug, which generally necessitates multiple doses of the drug. Another reason for multiple dose is the ability to measure drug level after multiple dose, but not after single dose. The BA or BE study should generally be conducted under fasting conditions. After an overnight fast of at least 10 h, a pre-dose blood sample is collected. Then, the test or reference products is administered with about 8 ounces (240 milliliters) of water. No food is allowed for at least 4 h post-dose. Blood samples are collected periodically after dosing, according to the protocol. PK sampling typically involves about 12–18 times points. Food effect studies are usually conducted for new drugs and drug products during the IND period to assess the effects of food on the rate and extent of absorption of a drug when the drug product is administered shortly after a meal (under fed conditions), as compared to administration under fasting conditions. Fed studies are typically conducted using meals that provide the greatest effects on gastrointestinal (GI) physiology and systemic drug availability (e.g. high fat meal). Non-replicate crossover study designs are generally recommended for BA and BE studies of IR and modified-release dosage forms. Non-replicate means drug is only taken once (i.e. for only one period). Some BE studies involve replicate design, where at least one product is given twice (i.e. single doses are given in one period, and then repeated in another period). Replicate crossover designs are used when evaluating BE of highly variable drugs or BE of narrow therapeutic index drugs. Replicate crossover designs are used to allow estimation of (1) with in-subject variance for the reference product, or for both the test and reference products, and (2) the subject by formulation interaction variance component. This design accounts for the inter-occasion variability that may confound the interpretation of a BE study as compared to a non-replicate crossover approach. Bioanalytical methods used for BA and BE studies should be accurate, precise, specific, sensitive, and reproducible, along with other properties. 18.3.3.2 Reference and Test Product When in vivo BE testing is required, the FDA generally designates a single reference standard that ANDA applicants must use in in  vivo BE testing. All generic versions must establish BE to that standard. Generally, FDA selects the RLD as the reference standard. However, in some instances (e.g. where the RLD has been withdrawn from sale and an ANDA is selected as the reference standard), the RLD and the reference standard may be different (21CFR314.3). The assayed drug content of the test product batch should not differ from the reference product by more than ±5%.

432

Z. G. Oner and J. E. Polli

18.3.3.3 Subjects In general, BA and BE studies should be conducted in healthy volunteers if the product can be safely administered to this population. Subjects should be 18 years of age or older, capable of giving informed consent, and together, represent the entire general population. Both male and female subjects should be enrolled in BA and BE studies unless there is a specific reason to exclude one sex. For example, oral contraceptives are only evaluated in female subjects because the indication is specific to females. In addition, the total number of subjects in a study should be sufficient to provide adequate statistical power. The study should have 80% or 90% power to conclude BE between these two formulations. According to EMA’s BE guideline (European Medicines Agency 2010), at least 12 volunteers must be enrolled, where they should have a body mass index between 18.5 and 30  kg/m2. Standardization of diet, fluid intake, exercise and posture is recommended. 18.3.3.4 Sample Collection and Sampling Times The active ingredient that is released from the dosage form or its active moiety should be measured in biological fluids in BA/BE studies. When appropriate, active metabolites should be collected. Under normal circumstances, blood, rather than urine or tissue, should be used. In most cases, drug or metabolites are measured in serum or plasma. However, in certain cases, such as when an assay of sufficient sensitivity cannot be developed for plasma, whole blood may be more appropriate for analysis. Twelve to eighteen samples, including a pre-dose sample, should be collected per subject per dose. This sampling should continue for at least three or more terminal elimination half-lives of the drug to capture 80–90% of the relevant AUC. For multiple dose studies, sampling should occur across the dose interval and include the beginning and the end of the interval. Three or more samples should be obtained during the terminal log-linear phase to obtain an accurate estimate of λz (i.e. elimination rate constant) from linear regression. 18.3.3.5 Subjects with Pre-dose Plasma Concentrations A sufficiently long washout should result in zero or negligible drug in plasma for the second period, at t = 0. If the pre-dose concentration is ≤5% of Cmax value in that subject, the subject’s data without any adjustments can be included in all PK measurements and calculations. However, if the pre-dose value is >5% of Cmax, the subject should be dropped from all PK evaluations. The subject data should be reported and the subject should be included in safety evaluations.

18  Bioavailability and Bioequivalence

433

18.3.3.6 Data Deletion Because of Vomiting Occasionally, a subject vomits during a study. Data from subjects who experience emesis during the course of a study for IR products should be deleted from statistical analysis if vomiting occurs at or before 2 times median Tmax. For modified-­ release products, subjects who experience emesis at any time during the labeled dosing interval should not be included in PK analysis. 18.3.3.7 Evaluation and Methods to Document BA/BE: Use of Plasma Drug Concentration BA and BE frequently rely on two main PK measures: area under the concentration time curve (AUC) to assess the extent of systemic exposure and peak or maximum concentration (Cmax) to assess the rate of systemic absorption. AUC(0-t) denotes the area under the concentration time curve from zero to the last measurable time point. AUC(0-∞) denotes the area under the concentration time curve extrapolated to infinity. If drug concentration is in terms of μg/ml, the units of AUC are μg/ml × hour and generally represents the extent of drug exposure in the body. Tmax is the time to reach peak concentration and is also reported but is generally not used in decision-making since it has limitations due to sampling scheme and statistical properties. Figure 18.1 illustrates the plasma drug concentrations over time following administration of a single oral dose drug. Once the minimum effective concentration (MEC) is reached, the pharmacological response begins. The concentration of the drug at the peak is Cmax, and the time required to reach Cmax is Tmax. At Cmax, in general, the rate of drug absorption is equal to the rate of drug elimination. In general, time that precedes Cmax (i.e. to the left of Cmax) represents the majority of the absorption phase. Meanwhile, time after Cmax largely represents the elimination phase. Within the therapeutic range, the desired pharmacological response is obtained (i.e. maximum efficacy and minimum toxicity). AUC expresses the total amount of drug exposure in the systemic circulation after drug administration. The duration of action is the time period that the drug concentration is greater than the MEC level. Minimum toxic concentration (or maximum therapeutic Fig. 18.1 Plasma concentration-time curve

MTC Therapeutic window

Drug concentration

Cmax

AUC MEC

Duration of action

Tmax

Time

434

Z. G. Oner and J. E. Polli

concentration – MTC) is the drug concentration at which the drug becomes toxic. Therefore, the concentration range between MEC and MTC, the therapeutic window, gives the optimal drug efficacy while limiting toxicity. For a drug with linear pharmacokinetics, the AUC and Cmax values increase or decrease proportionally with the dose. For a drug with non-linear kinetics, the changes in AUC and Cmax values are not proportional to dose and testing should be conducted at several different doses. BE assessment is typically made based on Cmax and AUC. However, it is possible for two different profiles to have the same Cmax and AUC, but the profiles to be meaningfully different (e.g. differ in Tmax). Hence, it is possible that Cmax and AUC do not capture all important attributes of profile shape. Differences in the shape of the systemic concentration profile between the test and reference products could imply that the test product may not produce the same clinical response as the reference product. In such cases, additional data analysis (e.g. partial AUCS), exposure-­ response evaluation, or clinical studies may be recommended to evaluate the BE of the two products. For example, the sleep medicine zolpidem extended-release tablet requires partial AUC0–1.5 and AUC 1.5-t for evaluation of BE, under fasting conditions. These two metrics aim to reflect early drug exposure for sleep onset (i.e. 0–1.5 h), and later drug exposure for sleep maintenance (i.e. 1.5 h to the last measurable time point) (U.S. Food & Drug Administration 2011). 18.3.3.8 Evaluation and Methods to Document BA/BE: Use of Urinary Drug Excretion

Fig. 18.2 Cumulative urinary excretion-time curve

Cumulative amount excreted

Urinary drug excretion studies are generally not encouraged. If serial measurements of the drug or its metabolites in plasma, serum, or blood cannot be accomplished, measurement of urinary excretion can be used to demonstrate BE. The concentration time profile of the parent drug is more sensitive to changes in formulation performance than a metabolite. Cumulative urinary excretion (Ae) and its maximal rate (Rmax) are employed as extent and rate metrics. Ae(0-t) is cumulative urinary excretion of unchanged drug (mg). Rmax is maximal rate of urinary excretion (mg/h). Figure 18.2 illustrates urinary drug concentrations over time.

Time

18  Bioavailability and Bioequivalence

435

18.3.4 Statistical Analysis and Acceptance Limits to Establish Bioequivalence For AUC and Cmax the 90% confidence interval for the ratio of the test and reference products should be contained within the acceptance interval of 80.00–125.00%. A statistical evaluation of Tmax is not required. However, if rapid release is claimed to be clinically relevant and of importance for onset of action or is related to adverse events, there should be no apparent difference in median Tmax and its variability between test and reference product. For the demonstration of complete drug absorption, data from absolute bioavailability or mass-balance studies could be used.

18.4 Bioequivalence Studies 18.4.1 The Basis for Determining Bioequivalence BE is established if the in vivo BA of a test drug product does not differ significantly in rate and extent of drug absorption from that of the reference when administered under the same molar dose of the active moiety under similar experimental conditions, either as single or multiple dose. If two formulations show similar concentration time profiles, they should exhibit similar therapeutic effects. The objective of the BE study is to measure and compare the formulation performance between two (or more) pharmaceutically equivalent drug products. A drug product may be placed on the market in the US after being approved from the FDA. The FDA uses the NDA pathway for innovators and the ANDA pathway for generics under the Federal Food, Drug, and Cosmetic Act. Table 18.1 summarizes the requirements imposed by the FDA for NDA and ANDA. Generics are generally not required to replicate the extensive clinical trials that have already been used in the development of the innovator, brand-name drug, since the safety and efficacy of the brand has already been well established. Instead, generic drug Table 18.1  NDA and ANDA requirements

NDA Chemistry Manufacturing Controls Labelling Testing Animal studies Clinical studies Bioavailability

ANDA Chemistry Manufacturing Controls Labelling Testing Bioequivalence

436

Z. G. Oner and J. E. Polli

products have to show they are pharmaceutical equivalent to and bioequivalent to innovator drug products, along with other manufacturing and labeling requirements. There are several common scenarios for BE testing. In general, the BE testing occurs in the context of change, where the reference is the current or pre-change formulation, and test is the post-change formulation. Innovator companies and generic companies both employ BE testing by virtue of various formulation or manufacturing changes. Scale-up and post-approval changes (SUPAC) is one context for BE testing. For example, perhaps soon after NDA approval, an innovator company may have the need to manufacture the product at a larger manufacturing scale or in a new location, such that the new “post-change” product needs to bioequivalent to the currently manufactured (i.e. currently approved) product. Generics also perform BE testing due to SUPAC. Generics also typically obtain ANDA approval through BE testing. Also, 505(b)(2) applications often rely on BE testing (e.g. approval of a new extended-release formulation based on BE to the approved IR formulation). Important Bioequivalence studies are performed during the development of generic products, which must prove bioequivalent to the innovator. When post-­ approval manufacturing changes are introduced, the reference will be the pre-­ change formulation.

18.4.2 Methods for Determining Bioequivalence According to FDA’s draft guidance for industry, “Bioavailability and Bioequivalence Studies Submitted in NDAs or INDs – General Considerations,” in vivo and in vitro methods can be used to measure BA and establish BE. The following approaches are used to determine BE. 18.4.2.1 Pharmacokinetic (PK) Studies BE frequently relies on pharmacokinetic endpoints such as Cmax and AUC, as explained above. 18.4.2.2 In Vitro Studies: Biopharmaceutics Classification System (BCS) and BCS-Based Biowaiver 21 CFR Part 320 addresses BA and BE requirements. One may request FDA to waive the requirement for the submission with evidence measuring the in vivo BA or demonstrating the in vivo BE of the drug product that is subject of a NDA, ANDA or supplemental application. Some eligibility criteria;

437

18  Bioavailability and Bioequivalence Table 18.2 Biopharmaceutics Classification System

BCS Class 1 2 3 4

Solubility High Low High Low

Permeability High High Low Low

1. The in vivo BA or BE of the drug product may be self-evident. Example include parenteral solutions, ophthalmic or otic solutions; some inhalation products; and topical (skin) solutions, oral solutions, and nasal solutions. 2. BA may be measured or BE may be demonstrated by in vitro in lieu of in vivo data, if the drug product is • in a different strength to another product with demonstrated BE • shown to meet an in vitro test that has been correlated with in vivo data • a reformulated product that is identical, except for a different color, flavor, or preservative that will not affect the BA 3. Protection of the public health (21CFR320.22) 4. A solid oral dosage form that meets the BCS criteria Biopharmaceutics Classification System The BCS is a scientific framework for classifying drug substances based on their aqueous solubility and intestinal permeability. According to the BCS, drug substances are classified as follows (Table 18.2). When combined with in vitro dissolution characteristics of the drug product, the BCS takes into account three major factors: solubility, intestinal permeability, and dissolution. Definitions A drug substance is considered highly soluble when the highest strength is soluble in 250 mL or less of aqueous media over the pH range of 1–6.8. A drug substance is considered to be highly permeable when the extent of absorption in humans is determined to be 85 percent or more of an administered dose based on a mass balance determination (along with evidence showing stability of the drug in the GI tract) or in comparison to an intravenous reference dose. An IR drug product is considered rapidly dissolving when 85 percent or more of the labeled amount of the drug substance dissolves within 30 minutes, using United States Pharmacopeia (USP) Apparatus I (basket method) at 100 rpm or Apparatus II (paddle method) at 50 rpm (or at 75 rpm when appropriately justified) in a volume of 500 mL or less in each of the following media:

438

Z. G. Oner and J. E. Polli

1. 0.1 N HCl or Simulated Gastric Fluid USP without enzymes; 2. a pH 4.5 buffer; and 3. a pH 6.8 buffer or Simulated Intestinal Fluid USP without enzymes. An IR product is considered very rapidly dissolving when 85 percent or more of the labeled amount of the drug substance dissolves within 15 minutes using the above mentioned conditions. The FDA is revising its guidance for industry on the waiver of in vivo bioavailability and bioequivalence studies for IR solid oral-dosage forms based on BCS (U.S. Food & Drug Administration 2015b). These waivers are intended for: a) subsequent in vivo BA or BE studies of formulations after the initial establishment of the in vivo BA of IR dosage forms during the IND period; and b) in vivo BE studies of IR dosage forms in ANDAs. This guidance is applicable for BA/BE waivers (biowaivers) based on BCS; for BCS class 1 and class 3 IR solid oral dosage forms. For BCS class 1 drug products, the following should be demonstrated to obtain a biowaiver: • • • •

the drug substance is highly soluble the drug substance is highly permeable the drug product (test and reference) is rapidly dissolving (85% within 30 min) the product does not contain any excipients that will affect the rate or extent of absorption of the drug

For BCS class 3 drug products, the following should be demonstrated to obtain a biowaiver: • the drug substance is highly soluble • the drug product (test and reference) is very rapidly dissolving (>85% within 15 min) • the test product formulation is qualitatively the same and quantitatively very similar, e.g., falls within SUPAC IR level 1 and 2 changes, in composition to the reference The concept is applicable to IR, solid pharmaceutical products for oral administration and systemic action having the same pharmaceutical form. BCS-based biowaivers are not applicable for narrow therapeutic range drugs and products designed to be absorbed in the oral cavity. The scientific basis of BCS considers how drug is absorbed. Observed in vivo differences in the rate and extent of absorption of a drug from two pharmaceutically equivalent solid oral products may be due to differences in drug dissolution in vivo. However, when the in vivo dissolution of an IR solid oral dosage form is rapid or very rapid in relation to gastric emptying and the drug has high solubility, the rate

18  Bioavailability and Bioequivalence

439

and extent of drug absorption is unlikely to be dependent on drug dissolution and/or GI transit time. Under such circumstances, demonstration of in vivo BA or BE may not be necessary for drug products containing class 1 and class 3 drug substances, as long as the inactive ingredients used in the dosage form do not significantly affect absorption of the active ingredients. Important Biowaivers are applicable to IR, solid pharmaceutical products for oral administration and systemic action having the same dosage form. BCS-based biowaivers are not applicable for narrow therapeutic range drugs and products designed to be absorbed in the oral cavity.

18.4.2.3 Other In Vitro Studies Dissolution/Drug-Release Testing In vitro dissolution studies has several uses, including contributing to waiver of in vivo BE for lower strengths of an IR solid oral dosage form. Most products have several strengths (e.g. 10 mg, 25 mg, 75 mg, and 100 mg tablets). In vivo BE is typically demonstrated using the highest strength (e.g. 100 mg tablet), and waivers of in vivo studies may be granted on lower strengths (e.g. 10 mg, 25 mg, and 75 mg tablets), based on in vitro dissolution testing. However, two requirements are that the multiple strengths of IR products show linear kinetics, and that the lower strengths are proportionately similar in composition to the highest strength product (e.g. the 50  mg tablet uses half of the same powder as the 100  mg tablet). The approval of the lower strengths is based on dissolution profile comparisons between these lower strengths and the strength of the batch used in the BE study (i.e. typically the highest strength). The dissolution profile of the lower strength must be similar to the dissolution profile of the highest strength. For IR formulations, comparison at 15 min is essential. Where more than 85% of the drug is dissolved within 15 min, dissolution profiles may be accepted as similar. That is, comparison is not necessary for very rapidly dissolving drug products (i.e. when over 85% dissolved in 15 min). When less than 85% is dissolved at 15 min, further mathematical evaluation should be used to compare profiles, often using the f2 similarity factor. An f2 value >50 suggests that two dissolution profiles are similar (U.S. Food & Drug Administration 1997). 18.4.2.4 Pharmacodynamic Studies Pharmacodynamics (PD) concerns drug action or drug effect (e.g. reduction in blood pressure). PD studies should consider the relationship of drug concentrations to drug effect. PD studies are not recommended for orally administered drug

440

Z. G. Oner and J. E. Polli

products when the drug is absorbed into systemic circulation; rather, a PK approach can be used to assess systemic exposure and evaluate BA or BE. PK endpoints are preferred when they are the most accurate, sensitive, and reproducible approach. However, in instances where a PK endpoint is not possible, a well-justified PD endpoint can be used to demonstrate BA or BE.

18.4.3 Statistical Approaches for Bioequivalence The only applied method to test BE is average bioequivalence (ABE), which focuses on the comparison of population averages response of a test and reference product, not on the variances. Meanwhile, population and individual BE approaches are alternative potential approaches to compare both averages and variances of the two (or more) formulations. The population BE approach assesses total variability of the measure in the population. The individual BE approach assesses within-subject variability for the test and reference products, as well as the subject-by-formulation interaction (U.S. Food & Drug Administration 2001).

18.4.4 Supra-Bioavailability If the test product shows larger bioavailability than reference (i.e. more than the upper limit of 125.00%), than the reference product, the test product cannot be considered as therapeutic equivalent. The new product may be considered as a new medicinal product.

18.5 Conclusion BE studies are frequently used though the pharmaceutical industry. BE has acceptable approaches to determine the safety and efficacy of drug products, which accelerates the approval process and reduce the cost of product development.

References 21CFR314.3. https://www.ecfr.gov/cgi-­bin/text-­idx?SID=d1173eb1e0b5c9151b05905ffe183d e3&mc=true&node=se21.5.314_13&rgn=div8. Accessed 28 Aug 2017 21CFR320.22. https://www.ecfr.gov/cgi-­bin/text-­idx?SID=d1173eb1e0b5c9151b05905ffe183d e3&mc=true&node=se21.5.320_122&rgn=div8. Accessed 28 Aug 2017

18  Bioavailability and Bioequivalence

441

European Medicines Agency (2010) Guideline on the investigation of bioequivalence. http://www. ema.europa.eu/docs/en_GB/document_library/Scientific_guideline/2010/01/WC500070039. pdf. Accessed 28 Aug 2017 U.S. Food and Drug Administration (1997) Guidance for industry, dissolution testing of immediate release solid oral dosage forms. https://www.fda.gov/downloads/drugs/guidances/ucm070237. pdf. Accessed 28 Aug 2017 U.S.  Food and Drug Administration (2001) Guidance for industry, statistical approaches to establishing bioequivalence. https://www.fda.gov/downloads/Drugs/ GuidanceComplianceRegulatoryInformation/Guidances/UCM070244.pdf. Accessed 28 Aug 2017 U.S. Food and Drug Administration (2011) Guidance on Zolpidem. https://www.fda.gov/downloads/Drugs/GuidanceComplianceRegulatoryInformation/Guidances/UCM175029.pdf. Accessed 28 Aug 2017 U.S.  Food and Drug Administration (2013) Guidance for industry, bioequivalence studies with pharmacokinetic endpoints for drugs. Submitted under an ANDA. https://www.fda.gov/ downloads/Drugs/GuidanceComplianceRegulatoryInformation/Guidances/UCM377465.pdf. Accessed 28 Aug 2017 U.S. Food and Drug Administration (2014) Guidance for industry, bioavailability and bioequivalence studies. Submitted in NDAs or INDs  – General Considerations. https://www.fda.gov/ downloads/Drugs/GuidanceComplianceRegulatoryInformation/Guidances/UCM389370.pdf. Accessed 28 Aug 2017 U.S.  Food and Drug Administration (2015a) Drug development and review definitions. https:// www.fda.gov/drugs/developmentapprovalprocess/howdrugsaredevelopedandapproved/ approvalapplications/investigationalnewdrugindapplication/ucm176522.htm. Accessed 28 Aug 2017 U.S. Food and Drug Administration (2015b) Guidance for industry, Waiver of in vivo bioavailability and bioequivalence studies for immediate-release solid oral dosage forms based on a biopharmaceutics classification system. https://www.fda.gov/downloads/Drugs/ GuidanceComplianceRegulatoryInformation/Guidances/UCM070246.pdf. Accessed 28 Aug 2017 U.S. Food and Drug Administration (2017a) Approved drug products with therapeutic equivalence evaluations, 37th edn. https://www.fda.gov/downloads/Drugs/DevelopmentApprovalProcess/ UCM071436.pdf. Accessed 28 Aug 2017 U.S.  Food and Drug Administration (2017b) What is the approval process for generic drugs? https://www.fda.gov/Drugs/ResourcesForYou/Consumers/BuyingUsingMedicineSafely/ UnderstandingGenericDrugs/ucm506040.htm. Accessed 28 Aug 2017 U.S. Food and Drug Administration (2017c) Generic drugs: questions & answers. https://www. fda.gov/drugs/resourcesforyou/consumers/questionsanswers/ucm100100.htm. Accessed 28 Aug 2017

Chapter 19

Drug Transporters Alan Talevi and Carolina Leticia Bellera

19.1 Introduction Membrane transporters play a central role in many essential body functions, including traffic and compartmentalization of physiological compounds, absorption of physiological compounds (and, occasionally, structurally related xenobiotics, including drugs), and protection from potentially toxic exogenous compounds. The latter function may be performed by restricting their absorption and/or distribution, and/or by promoting their elimination (either with no chemical modifications or as metabolites). The impact of drug biotransformation enzymes on pharmacokinetics and clinical outcomes is better understood than that of drug transporters, possibly because of the earlier discovery of metabolizing enzymes, although this has changed considerably in recent years. If we go back for a moment to Fig. 6.5, in Chap. 6, we can observe that approximately three quarters of the drugs are predominantly ionized (either as cations, anions, or zwitterions) at pH 7. This suggests that the movement of a significant proportion of the known therapeutic agents across biological membranes would be greatly restricted if not for the presence of transporters. Moreover, without transporters most of the metabolites of xenobiotic compounds, which are highly polar molecules, would be indefinitely confined to the metabolizing cell.

A. Talevi (*) · C. L. Bellera Biopharmacy/Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), CCT La Plata, La Plata, Argentina e-mail: [email protected] © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_19

443

444

A. Talevi and C. L. Bellera

Important Although ubiquitously expressed throughout the body, the highest levels of drug transporters are found in organs and tissues with barrier and/or elimination functions, including the gut, liver, kidney, blood-brain barrier, and placenta. They may  also play a significant role in the development of drug resistance. Transporters are involved in different functions: a) helping their substrates move across diffusional barriers (typically cell or organelle membranes), which they may not be able to (efficiently) cross otherwise. This is usual for polar (including charged) substrates; b) moving their substrates against an electrochemical gradient, which might be useful for building up reserves of a given chemical substance in a certain cell or cell compartment; and c) acting in cooperation with metabolizing enzymes, by regulating the flux of xenobiotics to which these enzymes are exposed. As we will see later, more than one of the preceding general functions might be facilitated by the same transporter, depending on the substrate under consideration and the transporter locations in the body.

The term transporter refers to a diversity of membrane proteins with diverse functions, structures, and subcellular locations. As seen in Chap. 3, they can mediate either facilitated diffusion or active transport. Facilitated diffusion implies the movement of the drug across the membrane down an electrochemical gradient. In contrast, active transport relies on energy coupling, which directly or indirectly implies ATP consumption; transporters responsible for active transport frequently move their substrates uphill. The most commonly invoked model to explain substrate transport by transporters is the alternating-access model. It proposes a substrate-binding site that can alternatively access either the extracellular (or extra-organellar) or intracellular (or intra-organellar) side of the membrane, resulting in outward and inward facing conformations, respectively (Forrest and Rudnick 2009; Rees et  al. 2009). After substrate(s) binding (coupled with energy consumption mechanisms in the case of active transport), conformational changes are induced, resulting in the exposure of this binding site to the other side of the membrane, producing substrate translocation. The relative binding affinities of the two conformations for the substrate largely determine the net direction of transport (e.g., the outward-facing conformation of an importer is expected to have a higher affinity for the substrate than the inward-­ facing conformation, whereas the opposite holds for exporters). Transporters can be classified or grouped according to different criteria. For instance, they may be grouped based on the direction of the flow of their substrates, substrate specificity, or sequence homology/evolutionary relationships.

19  Drug Transporters

445

Definition Taking the flow direction of their substrates as classification criterion, carriers may be classified as efflux transporters (which export their substrates out of a cell or, in some cases, an organelle), influx or uptake transporters (which import their substrates to the cell or organelle), and bidirectional transporters (which, as the name suggests, display bidirectional fluxes, e.g., the hepatic glucose transporter GLUT2 facilitates bidirectional glucose transport across the hepatocyte plasma membrane under insulin regulation, see Navale and Paranjape 2016).

The two most important transporter superfamilies related to drug transport are ATP-binding cassette (ABC) and solute carrier (SLC) transporters. In the case of SLCs, the term ‘superfamily’ may be misleading because it generally implies a common ancestor as well as detectably similar sequences and structures. However, phylogenetic analysis of human SLC suggests that it consists of 15 related families and 32 additional unlinked families (Schlessinger et al. 2010). Figure 19.1 shows how the alternating access model may be adapted to ABC transporters or SLC transporters that facilitate diffusion, such as symporters or antiporters. The existence of drug transporters gives rise to the possibility of drug-drug and food-drug interactions involving transporter inhibition and induction, as well as the existence of genetic polymorphisms that may result in inter-individual variability in drug sensitivity. We will also see that some drug transporters are responsible for multidrug resistance  issues. The potential involvement of some important drug transporters in drug-drug interactions is reflected in the different guidelines of regulatory organizations (US Food and Drug Administration 2017a, b; Cole et al. 2020). In addition to discussing the general features of the members of the ABC and SLC superfamilies that are more directly involved in drug transport, this chapter will discuss the coordination of drug transporters and metabolizing enzymes. Many SLC transporters are valuable pharmacological targets, including dopamine,

Fig. 19.1  Schemes showing how the alternating access model adapts to different transportation mechanisms. ABC transporters (left) and SLC transporters (right)

446

A. Talevi and C. L. Bellera

serotonin, and monoamine transporters. However, the focus of this chapter is the pharmacokinetic role of transporters, and SLC transporters of pharmacodynamic relevance will not be discussed.

19.2 ATP-Binding Cassette (ABC) Transporters Transporters from the ABC superfamily have a characteristic (minimal) modular architecture, which consists of four domains: two transmembrane domains (TMDs) that are embedded in the membrane bilayer and two ABCs (or nucleotide binding domains, NBDs) located in the cytoplasm (Rees et al. 2009; ter Beek et al. 2014). At the sequence level, the ABC transporter superfamily is identified by the highly conserved motifs present in the ABCs (which give name to the superfamily). Conversely, the sequences and architectures of TMDs, where the substrate binding sites are located, are variable, which implies variable substrate-specificity and reflects the chemical diversity of the substrates that can be translocated by the ABC superfamily. NBDs do not always function in association with TMDs and are involved in various functions that do not occur at the membrane (Ter Beek et al. 2014). Nevertheless, the name “ABC transporter” is only used when the NBDs are associated with TMDs that participate in substrate translocation. ATP binding to ABCs and its subsequent hydrolysis is coupled with the conformational states of the TMDs; accordingly, ABC transporters produce primary active transport. Important According to our current knowledge, ABC transporters in mammals (and, more generally, in eukaryotes) act exclusively as exporters, that is, they remove their substrates from the cell (or organelle) to the outside of the cell (or organelle). In prokaryotes, they can act as exporters or importers. The subcellular location of the transporter (e.g., apical or basolateral membranes of polarized cells) will determine the direction of the substrate flux. For instance, an ABC transporter expressed in the apical membrane of an enterocyte will export its substrates to the gut lumen, whereas another ABC transporter located at the basolateral membrane will export its substrates in the opposite direction, facilitating access to the capillary blood. In the first case, systemic bioavailability of the drug would be limited by the transporter. In the second, systemic bioavailability would be favored.

Approximately 50 ABC transporters, which participate in cholesterol and lipid transport, antigen presentation, mitochondrial iron homeostasis, and other fundamental processes, have been described in humans. Accordingly, mutations or dysfunction in these proteins have been associated with a variety of disorders, including

19  Drug Transporters

447

cystic fibrosis, diabetes, hypercholesterolemia, and Alzheimer’s disease (Rees et al. 2009; Pereira et al. 2018). In relation to drug pharmacokinetics, the most relevant transporters are p-­glycoprotein (Pgp) (the product of the ABCB1 gene, also known as multidrug resistance protein 1 (MDR1)), nine members of the multidrug resistance-associated protein (MRP) family (also known as MRPS or ABCCs), including MRP1 to MRP9, and the breast cancer resistance protein (BCRP) (also known as ABCG2). Important The nomenclature of ABC transporters that have an impact on drug pharmacokinetics (MDR1, MRPs, BCRP) reflects the fact that, initially, high expression levels of these transporters were identified in cancerous cells and were associated with cross-resistance phenomena  (in which a patient is being treated with one drug and the tumor acquires a type of drug resistance that enables it to avoid being killed by a variety of drugs). Later, it was observed that ABC transporters are ubiquitously expressed and play a pivotal role as detoxifying systems, acting in coordination with metabolizing enzymes, and that they also have a role in the traffic of endogenous susbtances. Their involvement in multi-drug resistance phenomena, as well as their ability to protect the body from a wide diversity of structurally and functionally unrelated xenobiotics (drugs included), emerge from their individual and collective polyspecificity (or wide substrate specificity). For example, Pgp alone can transport diverse substrates, such as aldosterone, amprenavir, bilirubin, cyclosporine, digoxin, doxorubicin, erythromycin, etoposide, itraconazole, paclitaxel, and verapamil, among many others (Kim 2002). The ability of a given transporter to recognize a variety of substrates arises from the presence of multiple and flexible binding sites (Safa 2004; Ferreira et al. 2013). Remarkably, the set of substrates of Pgp and CYP3A4 (the most promiscuous CYP450 member) partially overlap (Wacher et  al. 1995; Katoh et  al. 2001; Zhou 2008), and the tissue distributions of these key transporter and enzyme  are similar (Wacher et  al. 1995). They also share some common inducers, possibly being one of  the best-known examples of coordinated enzyme and transporter regulation (Salphaty and Benet 1998; Matheny et al. 2004; Huwyler et al. 2006; Annaert et al. 2007). Despite their wide substrate specificity, specific ABC transporters prefer certain types of substrates. The general features of the substrates for each transporter are listed in Table 19.1.

In addition to their well-known role as major determinants of multidrug resistance in cancer (Pérez-Tomás 2006; Sun et al. 2012), they have also been linked to refractoriness in other disorders, such as epilepsy (Feldmann and Koepp 2016). Several therapeutic strategies have been proposed to prevent or cope with constitutive or acquired cross-resistance linked to ABC transporters (Couyoupetrou et al.

448

A. Talevi and C. L. Bellera

Table 19.1  Some (very) general features of the known substrates of pharmacokinetically relevant ABC transporters. Note that, in some way, the universe of substrates of Pgp and MRPs seem to be complementary. As will be discussed later in the chapter, the specific substrate preferences of each transporter coupled with their cellular location in different organs will contribute to their coordinated function as well as their articulation with metabolizing enzymes Transporter/s Substrates characteristics Pgp Relatively hydrophobic drugs. Cationic molecules. Many substrates possess planar aromatic rings and positively-charged tertiary N atoms (Sarkadi et al. 2006; Subramanian et al. 2016). Strong overlap with BCRP. MRPs Anionic Organic anions and drug conjugated to glutathione, glucuronate and sulfate. MRP1 to MRP3 can transport unconjugated neutral organic drugs by (co) transporting them with free glutathione. MRP1 can confer resistance to arsenite and MRP2 to cisplatin. MRP4 overexpression is associated with high-level resistance to the nucleoside analogues (Borst et al. 2000; Homolya et al. 2003). Comparatively low overlap with Pgp. BCRP Large, hydrophobic substrates, both cations and anions. Sulfate and glucuronic conjugates, such as estrone-3-sulfate and 17β-estradiol 17-(β-d-glucuronide). In general, sulfated conjugates seem to be better BCRP substrates than glutathione and glucuronide conjugates. In addition, it transports phosphorylated nucleosides and nucleotides. More than 200 verified substrates (Sarkadi et al. 2004; Mao and Unadkat 2015).

2017). The first explored approach was the co-administration of chemosensitizers capable of reversing the transporter effect. However, there are currently no approved drugs available for clinical use to reverse multi-drug resistance by inhibiting ABC transporters (Nanayakkara et al. 2018). Although some Pgp inhibitors have reached clinical trials, discouraging results have been obtained, mainly due to toxicity issues and limited efficacy (Lhommé et al. 2008; Tiwari et al. 2011; Choi and Yu 2014). The reader should consider the physiological role of ABC transporters as a general detoxification mechanism and their involvement in the traffic of physiological compounds, which may discourage the use of add-on inhibitors in the context of longterm therapies. However, the study of next-generation, highly selective inhibitors of ABC transporters continue to be an active area of research. Some studies have shifted their focus to therapeutic agents that target the signaling cascade that regulates efflux transporter expression. Another possibly safer approach involves the use of Trojan horse stratagems to deliver therapeutic levels of ABC transporters to the targeted tissue. By encapsulating ABC transporter substrates in particulate delivery systems (primarily pharmaceutical nanocarriers), recognition and translocation by efflux pumps can be avoided or minimized (Potschka and Luna-Munguia 2014; Blanco et al. 2015). Alternatively, the design of novel therapeutics that are not recognized by the transporters (disregarding their substrates early in the drug discovery and development process) may be possible, thus considering ABC transporters as anti-targets.

19  Drug Transporters

449

Example An 11-year and, so far, healthy patient was admitted to the emergency room with simple partial motor seizures accompanied by fever and headache. The seizures were immediately terminated with intravenous diazepam. The next day, the child again experienced a simple partial motor seizure, which was unresponsive to diazepam, lorazepam, and phenytoin. The episode evolved to generalized convulsive status epilepticus (a medical emergency associated with high mortality and morbidity, characterized by continuous seizures lasting more than 30 min, or two or more seizures without full recovery of consciousness between any of them), which the physicians were not able to control despite sequential administration of lorazepam, fentanyl, and phenobarbital. The boy became comatose. After trying other unsuccessful therapeutic options, verapamil was administered on day 37 based on the evidence that it has anticonvulsant activity in an animal model and the fact that the boy had supraventricular tachycardia. The patient regained consciousness, was able to breathe spontaneously, and status epilepticus promptly disappeared. The dramatic response to verapamil may be explained by its direct anticonvulsant action, and also because this drug is a known P-gp inhibitor that may have acted in the cerebrovascular endothelium and in the epileptic focus by facilitating the brain penetration of the antiseizure drugs that the patient was simultaneously receiving (at that time, valproic acid and phenobarbital). Remarkably, phenobarbital has been shown to act as a Pgp substrate (Luna-­ Tortós et al. 2008). Example based on a case report by Iannetti et al. (2005).

19.3 Solute Carriers (SLC) SLC transporters are the other most important family of drug-transport proteins, but their role in pharmacokinetics has possibly been overlooked. Currently, the Gene Nomenclature Committee of the Human Genome Organization classifies approximately 400 human SLC transporters into over 60 families based on their sequences, number of transmembrane α-helices, and biological functions (Schlessinger et al. 2010, 2013) (note, however, that their organization is highly dynamic). SLC function as facilitative transporters (allowing solutes to flow downhill with their electrochemical gradients) or as secondary or tertiary active transporters (allowing their substrates to flow uphill against their electrochemical gradient by coupling to the transport of a second solute, including co-transporters and exchangers). Despite their weak sequence similarities, solute carriers can share similar structural features (Taft 2009). Most members of this superfamily are located in the cell membrane, although some are embedded in organelle membranes. They are typically composed of one large domain, consisting of 10–14 transmembrane α-helices. The nomenclature for these transporters is, in general, similar to that already discussed for CYP450

450

A. Talevi and C. L. Bellera

members in Chap. 5. A protein from this superfamily is generically denoted SLCnXm, where SLC is the root name that specifies membership to the superfamily, n is an integer representing a family, X is a letter denoting a subfamily, and m is an integer that specifies an individual family member. Members of the the SLC1 (glutamate and neutral amino acid transporter), SLC6 (Na+− and Cl−-dependent neurotransmitter transporters), SLC17 (vesicular glutamate transporters), SLC18 (vesicular amine transporters), and SLC32 (vesicular inhibitory amino acid transporters) regulate the concentration of neuro-transmitters, thereby regulating neurosignaling pathways (Gether et al. 2006); therefore, several members of these families are exploited as drug targets. Members of several SLC families, such as SLCO (organic anion transporting polypeptide superfamily, formerly SLC21), SLC15 (oligopeptide transporters), SLC22 (organic cation/anion/zwitterion transporters), and SLC47 (multidrug and toxin extrusion -MATE- transporters), are abundant in the liver, kidney, and blood-­ brain-­ barrier and thus impact drug distribution, metabolism, and excretion (Schlessinger et al. 2013; Hagenbuch and Stieger 2013; Koepsell 2013). Some members of the SLC are particularly relevant in terms of drug absorption and/or disposition. SLC22 members are characterized by a wide substrate specificity and wide tissue distribution. OAT1 is a dicarboxylate exchange (sodium dependent) protein responsible for basolateral uptake of organic anions into the tubular epithelial cells in the kidney (Taft 2009). It mediates the urinary excretion of a wide array of substrates, including antibiotics, antiviral nucleoside analogs and nonsteroidal anti-inflammatory medications. More than 100 drugs and toxins have been found to interact with OAT1. OAT3 is a basolateral transporter similarly to OAT1. There is some degree of overlap in the substrate specificity between them  both; OAT3 transports sulfate- and glucuronide steroid conjugates. OAT1 and OAT3 are inhibited by probenecid. OATs function as antiporters. Other members of SLC22 specialize in the translocation of cationic drugs, mainly the organic cation transporters (OCT1/2) that function as uniporters and the organic cation and carnitine transporters (OCTN1/2) (Russel 2010). OCT1 is strongly expressed in the liver. OCT2 is highly expressed in the kidneys. Both mediate the cellular uptake of small monovalent organic cations by facilitating diffusion down the electrochemical gradient. Other cation transporters belonging to the SLC47 family have also recently been identified. They are proton-cation antiporters of the multidrug and toxic compound extrusion (MATE) family. MATE1 and MATE2 are efflux transporters fueled by an inwardly directed proton gradient and (partially) responsible for the final step in the excretion of metabolic waste and xenobiotic organic cations in the kidney and liver. The SLCO family consists of organic anion transporting polypeptides (OATPs) that are involved in the cellular uptake of bulky (MW  >  450  Da) and relatively hydrophobic organic anions. In addition, the substrate specificity of OATPs covers a wide range of amphipathic compounds, including bile salts, steroid conjugates, thyroid hormones, and several drugs, such as enalapril and digoxin. OATPs are widely expressed in different physiological barriers, including the gut, renal tubules, blood-brain barrier, and placenta.

19  Drug Transporters

451

Peptide transporters PEPT1 and PEPT2 are responsible for the absorption and conservation of the nitrogen from dietary protein (dipeptides, tripeptides) but they also have a key role in the absorption and disposition of peptidomimetic/peptide-­ like therapeutics, including various prodrugs functionalized with amino acids, such as valaciclovir and valganciclovir.

19.4 Cooperation Is Key ABC and SLC transporters are associated with drug absorption and disposition, and operate in a highly concerted manner with biotransformation enzymes. For several years, a persistent and often ignored question was how the xenobiotic molecules gained access to the metabolism sites within the metabolizing cell (Döring and Petzinger 2014) and, above all, how the resulting polar metabolites left such cells to be ultimately excreted (note that before undergoing metabolism, lipophilic drugs may enter the metabolizing cell by passive diffusion through the cell membrane; however, it is highly unlikely that very hydrophilic metabolites, e.g., drug conjugates, can exit the cell passively). Accumulation of drug metabolites within the metabolizing cell was not an option either, as high concentrations of the biotransformation products within the metabolizing cell would most likely disrupt cell homeostasis and become toxic. Furthermore, Phase I and Phase II metabolites appear in the urine and/or feces, which implies that they do not accumulate in metabolizing cells. Therefore, there must be a facilitative mechanism for exporting drug metabolites. It was not until the first ABC efflux pump, Pgp, was discovered in the late 1970s by Ling and Baker and cloned in the mid-1980s (Ling and Baker 1978; Riordan et  al. 1985) that an answer to such enduring query began being delineated. The term “Phase III” was coined by Ishikawa (1992) and commented later by Zimniak et al. (1993) to describe the next sequential step for drug metabolites awaiting elimination and having undergone Phase I and Phase II metabolism. In his original article, Ishikawa distinguished between the elimination of intact (non-metabolized) drugs such as vinblastine and doxorubicin, and the elimination of glutathione conjugates by a hypothetical “GS-X pump” localized at the canalicular membrane of the hepatocytes. Currently, such export pumps are identified as different members of the ABC superfamily, particularly MRPs. Ishikawa et al. did not consider carrier-mediated uptake from the blood or gut lumen into metabolizing cells. Such process was later called “Phase 0 transport” (Petzinger and Geyer 2006). In the liver and kidney, Phase 0 transport takes place at the basolateral blood-facing membrane, which is separated from the apical (canalicular or tubular) membrane. The sequence of the transport phases (Phase 0-Phase III) allows for vectorial transcellular xenobiotic transfer (Döring and Petzinger 2014). Cytosolic transport (which comprises intracellular delivery between metabolic sites and membrane transport sites) is a largely ignored part of the whole process. This may occur through several processes, including free diffusion,

452

A. Talevi and C. L. Bellera

transport by interaction with cytoplasmic membranes or proteins, cytoplasmic flow, and vesicular transport (Zucker et al. 1996; Petzinger and Geyer 2006; Boyer 2013). Important There are three general mechanisms of interplay between transporters and metabolic enzymes. In the first case, an efflux transporter may (down) regulate the rate at which a given xenobiotic is taken up by a cell, so that the enzymatic machinery is not saturated by transient exposure to high xenobiotic levels (Fagiolino 2017). This type of transporter/enzyme cooperation may have a significant impact on drug absorption in the gut. Imagine, for instance, the cooperation of Pgp and CYP3A4 (we have already highlighted that these two systems have a strongly overlapping substrate specificity, and that they are colocalized). Pgp is located in the apical membrane of enterocytes, whereas CYP3A4 is expressed in the endoplasmic reticulum of these cells (Fig. 19.2). It should be noted that a) Pgp removes lipophilic drugs from the cell, and b) the drug concentration gradient across the intestinal epithelia can be quite high, especially when high doses of a drug are orally administered. Lipophilic drugs tend to be highly permeable; when administered at high doses, it is highly probable that enterocyte-­ metabolizing enzymes are overwhelmed by the high drug influx rate (the number of drug molecules entering the enterocyte per unit of time exceeds the number of available enzyme copies; accordingly, a fraction of the drug molecules will evade the intestinal first-pass effect). This situation is illustrated in Fig. 19.3. The efflux transporter in the apical membrane limits the number of xenobiotic molecules entering the cell per unit of time so that the system does not become saturated (the same principle impacts the number of molecules that reach the liver per time unit, with the concurrent optimization of the hepatic first-pass effect). A second method of cooperation between transporters and enzymes can be observed in eliminating organs, especially for drugs with high extraction ratios. Passive permeability can be rate-limiting for (hepatic) clearance if intrinsic metabolism is fast. Therefore, uptake transporters in the sinusoidal membrane operate by improving drug influx into the cell so that the high levels of enzymatic systems within the liver can be better exploited (Annaert et al. 2007; Di et al. 2012) (see Fig. 19.4). Finally, we must consider the contribution of efflux transporters to eliminate (potentially toxic) waste products (Annaert et al. 2007). In particular, it should be noted that Phase II metabolites, owing to their high polarity, cannot readily diffuse through cell membranes. If not with the help of transporters, and even considering their favorable concentration gradient, they would find it difficult to leave hepatocytes for subsequent biliary excretion (Fig. 19.4), or to reach the cytosol of the tubular cells (Fig. 19.5) and move from the tubular cells to the tubular lumen, as in tubular secretion.

Fig. 19.2  Scheme showing the expression and location of different transporters in the gut wall

Fig. 19.3  Scheme illustrating how the efflux transporters cooperate with biotransformation enzymes by limiting the flux of the drug to the metabolizing organs. Adapted from Fagiolino (2017)

454

A. Talevi and C. L. Bellera

Fig. 19.4  Scheme showing the expression of different transporter systems in hepatocyte sinusoidal and canalicular membranes

Fig. 19.5  Scheme showing the expression of different transporter systems in the tubular epithelium

19  Drug Transporters

455

References Annaert P, Swift D, Lee JK et  al (2007) Drug transport in the liver. In: You G, Morris ME (eds) . Wiley Blanco E, Shen H, Ferrari M (2015) Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat Biotechnol 33:941–951 Borst P, Evers R, Kool M et  al (2000) A family of drug transporters: the multidrug resistance-­ associated proteins. J Natl Cancer Inst 92:1295–1302 Boyer JL (2013) Bile formation and secretion. Compr Physiol 3:1035–1078 Choi YH, Yu AM (2014) ABC transporters in multidrug resistance and pharmacokinetics, and strategies for drug development. Curr Pharm Des 20:793–807 Cole S, Kerwash E, Andersson A (2020) A summary of the current drug interaction guidance from the European medicines agency and considerations of future updates. Drug Metab Pharmacokinet 35:2–11 Couyoupetrou M, Gantner ME, Di Ianni ME et  al (2017) Computer-aided recognition of ABC transporters substrates and its application to the development of new drugs for refractory epilepsy. Mini Rev Med Chem 17:205–215 Di L, Keefer C, Scott DO et al (2012) Mechanistic insights from comparing intrinsic clearance values between human liver microsomes and hepatocytes to guide drug design. Eur J Med Chem 57:441–448 Döring B, Petzinger E (2014) Phase 0 and phase III transport in various organs: combined concept of phases in xenobiotic transport and metabolism. Drug Metab Rev 46:261–282 Fagiolino P (2017) In Farmacocinética y biofarmacia. Parte I: principios fundamentales. UdelaR-FQ; FUNDAQUIM, Montevideo Feldmann M, Koepp M (2016) ABC transporters and drug resistance in patients with epilepsy. Curr Pharm Des 22:5793–5807 Ferreira RJ, Ferreira MJ, dos Santos DJ (2013) Molecular docking characterizes substrate-­ binding sites and efflux modulation mechanisms within P-glycoprotein. J Chem Inf Model 53:1747–1760 Forrest LR, Rudnick G (2009) The rocking bundle: a mechanism for ion-coupled solute flux by symmetrical transporters. Physiology (Bethesda) 24:377–386 Gether U, Andersen PH, Larsson OM et al (2006) Neurotransmitter transporters: molecular function of important drug targets. Trends Pharmacol Sci 27:375–383 Hagenbuch B, Stieger B (2013) The SLCO (former SLC21) superfamily of transporters. Mol Asp Med 34:396–412 Homolya L, Váradi A, Sarkadi B (2003) Multidrug resistance-associated proteins: export pumps for conjugates with glutathione, glucuronate or sulfate. Biofactors 17:103–114 Huwyler J, Wright MB, Gutmann H et  al (2006) Induction of cytochrome P450 3A4 and P-glycoprotein by the isoxazolyl-penicillin antibiotic flucloxacillin. Curr Drg Metab 7:119–126 Iannetti P, Spalice A, Parisi P (2005) Calcium-channel blocker verapamil administration in prolonged and refractory status epilepticus. Epilepsia 46:967–969 Ishikawa T (1992) The ATP-dependent glutathione S-conjugate export pump. Trends Biochem Sci 17:463–468 Katoh M, Nakajima M, Yamazaki H et al (2001) Inhibitory effects of CYP3A4 substrates and their metabolites on P-glycoprotein-mediated transport. Eur J Pharm Sci 12:505–513 Kim RB (2002) Drugs as p-glycoprotein substrates, inhibitors, and inducers. Drug Metab Rev 34:47–54 Koepsell H (2013) The SLC22 family with transporters of organic cations, anions and zwitterions. Mol Asp Med 34:413–435 Lhommé C, Joly F, Walker JL et  al (2008) (PSC 833) combined with paclitaxel and carboplatin compared with paclitaxel and carboplatin alone in patients with stage IV or suboptimally debulked stage III epithelial ovarian cancer or primary peritoneal cancer. J Clin Oncol 26:2674–2682

456

A. Talevi and C. L. Bellera

Ling V, Baker RM (1978) Dominance of colchicine resistance in hybrid CHO cells. Somatic Cell Genet 4:193–200 Luna-Tortós C, Fedrowitz M, Löscher W (2008) Several major antiepileptic drugs are substrates for human P-glycoprotein. Neuropharmcology 55:1364–1375 Mao Q, Unadkat JD (2015) Role of the breast cancer resistance protein (BCRP/ABCG2) in drug transport–an update. AAPS J 17:65–82 Matheny CJ, Ali RY, Yang X et al (2004) Effect of prototypical inducing agents on P-glycoprotein and CYP3A expression in mouse tissues. Drug Metab Dispos 32:1008–1014 Nanayakkara AK, Follit CA, Chen G et al (2018) Targeted inhibitors of P-glycoprotein increase chemotherapeutic-induced mortality of multidrug resistant tumor cells. Sci Rep 8:967 Navale AM, Paranjape AN (2016) Glucose transporters: physiological and pathological roles. Biophys Rev 8:5–9 Pereira CD, Martins F, Wiltfang J et al (2018) ABC transporters are key players in Alzheimer’s disease. J Alzheimer Dis 61:463–485 Pérez-Tomás R (2006) Multidrug resistance: retrospect and prospects in anti-cancer drug treatment. Curr Med Chem 13:1859–1876 Petzinger E, Geyer J (2006) Drug transporters in pharmacokinetics. Naunyn Schmiedeberg’s Arch Pharmacol 372:465–475 Potschka H, Luna-Munguia H (2014) CNS transporters and drug delivery in epilepsy. Curr Pharm Des 20:1534–1542 Rees DC, Johnson E, Lewinson O (2009) ABC transporters: the power to change. Nat Rev Mol Cell Biol 10:218–227 Riordan JR, Deuchars K, Kartner N et  al (1985) Amplification of P-glycoprotein genes in multidrug-­resistant mammalian cell lines. Nature 316:817–819 Russel FGM (2010) Transporters: importance in drug absorption, distribution, and removal. In: Rodrigues AD, Peter RM (eds) Pang KS. Springer, New York Safa AR (2004) Identification and characterization of the binding sites of P-glycoprotein for multidrug resistance-related drugs and modulators. Curr Med Chem Anticancer Agents 4:1–17 Salphaty L, Benet LZ (1998) Modulation of P-glycoprotein expression by cytochrome P450 3A inducers in male and female rat livers. Biochem Pharmacol 55:387–395 Sarkadi B, Ozvegy-Laczka C, Német K et al (2004) BCG2 – a transporter for all seasons. FEBS Lett 567:116–120 Sarkadi B, Homolya L, Szakács G et al (2006) Human multidrug resistance ABCB and ABCG transporters: participation in a chemoimmunity defense system. Physiol Rev 86:1179–1236 Schlessinger A, Matsson P, Shima JE et al (2010) Comparison of human solute carriers. Protein Sci 19:412–428 Schlessinger A, Khuri N, Giacomini KM et al (2013) Molecular modeling and ligand docking for solute carrier (SLC) transporters. Curr Top Med Chem 13:843–856 Subramanian N, Schumann-Gillett A, Mark A et  al (2016) Understanding the accumulation of P-glycoprotein substrates within cells: The effect of cholesterol on membrane partitioning. Biochim Biophys Acta Biomembr 1858:776–782 Sun YL, Patel A, Kumar P et al (2012) Role of ABC transporters in cancer chemotherapy. Chin J Cancer 31:51–57 Taft DR (2009) Drug excretion. In: Hacker M, Bachmann K, Messer W (eds) . Academic, Burlington Ter Beek J, Guskov A, Slotboom DJ (2014) Structural diversity of ABC transporters. J Gen Physiol 143:419–435 Tiwari AK, Sodani K, Dai CL et al (2011) Revisiting the ABCs of multidrug resistance in cancer chemotherapy. Curr Pharm Biotechnol 12:570–594 US Food and Drug Administration (2017a) In vitro metabolism and transporter-mediated drug-­ drug interaction studies. Guidance for industry US Food and Drug Administration (2017b) Clinical drug interaction studies – study design, data analysis, and clinical implications. Guidance for industry

19  Drug Transporters

457

Wacher VJ, Wu CY, Benet LZ (1995) Overlapping substrate specificities and tissue distribution of cytochrome P450 3A and P-glycoprotein: Implications for drug delivery and activity in cancer chemotherapy. Mol Carcinog 13:129–134 Zhou SF (2008) Drugs behave as substrates, inhibitors and inducers of human cytochrome P450 3A4. Curr Drug Metab 9:310–322 Zimniak P, Awasthi S, Awasthi YC (1993) Phase III detoxification system. Trends Biochem Sci 18:164–166 Zucker SD, Goessling W, Gollan JL (1996) Intracellular transport of small hydrophobic compounds by the hepatocyte. Semin Liver Dis 16:159–167

Further Reading Transporter science is an expanding field, and this Chapter in only intended as an introduction to the topic. The reader may find deeper insight into some excellent volumes that specifically deal with the subject, such as the ones edited by You and Morris (Drug Transporters: Molecular Characterization and Role in Drug Disposition, Second Edition, Wiley, 2014); Pang, Rodrigues and Peter (Enzyme- and Transporter-based Drug Drug Interactions. Progress and Future Challenges, Springer, 2010); or Ecker and Chiba (Transporters as drug carriers. Structure, Function, Substrates, Wiley-VCH, 2009)

Chapter 20

Blood Flow Distribution and Membrane Transporters as Determinant Factors of Tissue Drug Concentration Pietro Fagiolino, Alan Talevi, Marta Vázquez, and Manuel Ibarra

20.1 Basic Principles of Drug Exchange Between Blood and Tissues 20.1.1 Tissue Uptake of Solutes The different tissues of the human body do not gather around the blood as if they could take up the molecules dissolved therein in the required amount. In contrast, blood enters each of them in different amounts, according to their requirements for such solutes. In other words, extravascular tissues do not take up solutes from the blood but the blood delivers them on demand. There are types of tissues present in almost all organs, whereas some tissues are present in specific organs; however, all of them exhibit a practically constant distance between their cells and the capillary endothelium. Even in tissues with low capillary density, such as skeletal muscle, a convoluted three-dimensional geometry with extensive intercapillary branching approaches the tissue cells to the bloodstream (Poole et al. 2005). Consequently, in normal physiological conditions there are no tissues deeper than others in strictly spatial terms. A completely different scenario may occur when non-physiologic, abnormal irrigation of tissues is observed. Tumor cells often exhibit a high metabolic demand P. Fagiolino (*) · M. Vázquez · M. Ibarra Pharmaceutical Sciences Department, Faculty of Chemistry, Universidad de la República, Montevideo, Uruguay e-mail: [email protected] A. Talevi Laboratory of Bioactive Compound Research and Development (LIDeB), Faculty of Exact Sciences, University of La Plata (UNLP), La Plata, Argentina Argentinean National Council of Scientific and Technical Research (CONICET), La Plata, Argentina © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 A. Talevi, P. A. Quiroga (eds.), ADME Processes in Pharmaceutical Sciences, https://doi.org/10.1007/978-3-031-50419-8_20

459

460

P. Fagiolino et al.

and require access to the circulation system to thrive. For that purpose, they release pro-angiogenic signals that induce deregulated formation of new blood vessels with irregular architecture (e.g., inadequate division, markedly heterogeneous caliber size, chaotic blood flow, disrupted endothelial junctions) (Lugano et  al. 2020; Sriraman et al. 2014). Particularly in the case of solid tumors, deep tissues (in spatial terms), may be observed, as evidenced by tumor areas of persisting or intermittent hypoxia (Kimura et  al. 1996). Enhanced permeability and lack of proper lymphatic drainage due to uneven lymphatic vessel distribution can result in increased interstitial fluid pressure, which poses a substantial diffusional barrier. These same factors that limit oxygen supply to tumor cells can hamper drug delivery. In general, though, the concept of “deep tissues” has been used in the context of pharmacokinetic modelling to describe poorly perfused tissues (those that present low metabolic demand o for which their physiological role does not require an intensive mass exchange with the blood). August Krogh’s centennial hypothesis of capillary recruitment posed that capillaries in skeletal muscle could assume opened or closed states depending on the energetic demand of the tissue. The Krogh cylinder geometry has however been proven inadequate for modelling oxygen and nutrients transport towards the extravascular space (Poole et al. 2020). Blood flows through all capillaries, even when the energy demand of the tissue is low (Poole et al. 2013). As the demand increases, as occurs in skeletal muscle when it exercises, the flow rate increases not because of a capillary recruitment (Krogh’s hypothesis) but because of an increase in the exchange area along the capillary endothelium due to the increase in red blood cell and plasma flow in already flowing capillaries (longitudinal recruitment). The movement of molecules is key to understanding mass transfer throughout the body. Two main mechanisms of mass transfer complement each other: diffusion and convection. The former involves the molecules themselves moving through the solvent, mainly water in living beings, or any other non-aqueous barriers that stand in their way, whereas the second involves both the solute and the solvent moving jointly. Diffusion becomes relevant over short distances, while convective motion is extremely efficient for traveling long distances (Jacob et al. 2016; Hladky and Barrand 2014). The life of multicellular organisms would be unfeasible if convective movement did not occur. The convective movement of the bloodstream provides the main driving force for the disposition of xenobiotics in the body. Convection and diffusion also operate in the extravascular space of the tissues. Once the molecules arrive in the vicinity of the capillary endothelium, diffusion allows them to cross the blood/tissue interface. There is an important difference between convective movements that operate in intravascular and extravascular spaces. While extravascular convection only affects the tissue itself, intravascular convection affects all tissues in the body. To better understand what the latter implies, it is convenient to take note of some of the most significant volumes of the circulatory system. For a 70 kg man, the stroke volume is approximately 80 mL, this is one third of the blood volume housed in all the capillaries of the body (240 mL). If each tissue received exactly the same fraction of the stroke volume that corresponds to its relative size, the blood volume of all capillary beds would be similar. In other words, each capillary in the

20  Blood Flow Distribution and Membrane Transporters as Determinant Factors…

461

body would be irrigated by the same flow rate and, therefore, their respective endothelium would be equally stressed. Body water is distributed in approximately the same ratio: 4.5 L in the blood, 9 L in the interstitial spaces, and 27 L within the cells of the tissues. That is, one third of the water moves outside the cells of the tissues (blood versus interstitial space), just as one third of the blood water circulates through the capillaries during each systole. After the systole period, two thirds of the blood remain in the capillary beds, as well as two thirds of the extracellular tissue water remains in the interstitial space. However, these volumes suffered the onslaught of the convective movement, transferring such kinetic force towards the respective interfaces: blood/tissue (capillary endothelium) and extra/intracellular compartment of tissues (cell membrane). Not all tissues receive a fraction (σ) of cardiac output similar to the fraction (ω) of body water they contain; the fraction of cardiac output that reaches them is directly related to their energy demand (oxygen plus fuel). Therefore, some tissues are supplied by a high blood flow rate (highly perfused), whereas others are not (poorly perfused). This leads to different stresses at the different interfaces (capillary endothelium and cell membrane) of tissues. The impact caused by fluid flowing in contact with solid surfaces has been extensively studied at the level of the so-­ called boundary layer (Welty et al. 1969; Rocha-Uribe and Soto-Armenta 2021); in this case, the boundary layer of the capillary endothelium. Definition In fluid mechanics, a boundary layer is the thin layer of fluid in the immediate vicinity of a bounding surface. The fluid’s interaction with this surface induces a no-slip boundary condition (zero velocity at the wall). The flow velocity monotonically increases as we move away from surface, until it reaches the bulk flow velocity. At the laminar flow boundary, a very smooth flow can be observed, while the turbulent flow boundary layer exhibits swirls or “eddies.”

Under a laminar flow regime, fluids flowing through pipelines develop, adjacent to the cylindrical wall, a narrow zone where the axial motion is less than that in the center of the cylinder. Very narrow at the beginning and then it widens as it progresses throughout the pipeline (Fig. 20.1). The increased pressure against the endothelial wall at the beginning of the capillary generates the driving force for solute transfer towards the extravascular space. As the boundary layer widens, the transfer wanes. A higher fraction of cardiac output delivered to the tissue causes lengthening of the narrow boundary layer along the capillary (longitudinal recruitment) thus causing greater uptake of solutes by the tissue. The boundary layer of constant thickness shown in Fig. 20.1 (right side) represents a simplification of the system in order to handle the kinetics of such a complex biological phenomenon. Because the solutes contained in the vascular compartment shorten their distance to the exit interface, the transfer rate increases (Fagiolino et  al. 2003; Fagiolino 2004; Fagiolino and Vázquez 2022). The boundary layer concept is now becoming very useful in explaining the uptake of solutes by different types of tissues (Nishihara and Ackerman

462 Fig. 20.1  Blood flowing through the capillary according to the laminar flow model. The higher cardiac output fraction, the longer the boundary layer developed at the capillary wall (left panel) is. In the right panel this is represented as a narrower boundary layer along the entire capillary wall

P. Fagiolino et al.





2007; Lichtenberg et  al. 2017). In addition, some investigations have shown the impact of a pulsatile flow on the uptake of oxygen (Moschandreou et al. 2010). It is interesting to note in Fig.  20.1 that the blood volume in highly perfused capillaries increases, which generates extra pressure at the boundary layer of the capillary endothelium (Pries et al. 2000). Some researchers suggest that this high pressure would cause the permeation of water even in barriers as restrictive as the blood-brain barrier (Hladky and Barrand 2014), thus leading to an increase in solute concentration at the boundary layer and, therefore, a greater diffusion towards the tissue. It is thus not surprising that highly perfused tissues are the largest producers of lymph. Aquaporins 1 and 4 (AQP1 and AQP4), members of a family of membrane water channels, are expressed in the central nervous system, and it is possible that these proteins contribute to water transport across the blood-brain barrier (Bonomini and Rezzani 2010). Narrowing of the capillary endothelium boundary layer accelerates mass exchange in both directions (blood-to-tissue and tissue-to-blood). However, not only the rate of exchange increases, but there is also a net transfer of solutes to more perfused tissues at the expense of less perfused ones. Since blood flowing through both highly and poorly perfused capillaries has the same concentration of solutes (arterial concentration), greater transfer to the former leads to less transfer to the latter. In fact, an asymmetric tissue/blood balance of free solute is established at both tissue/blood interface: A water/water partition coefficient greater than 1 at the interface that receives a higher fraction of cardiac output and less than 1 at the interface that receives a lower fraction (Fagiolino and Vázquez 2022). Important A higher fraction of the cardiac output delivered to a given tissue tends to increase the rate at which solutes are extravasated due to narrowing of the capillary endothelial boundary layer. Besides, the increased pressure on the capillary wall causes water to leak into the extravascular space, thus increasing the concentration of solute within the boundary layer. The opposite happens in the boundary layer of tissues that receive a lower fraction of the cardiac output than their water content.

20  Blood Flow Distribution and Membrane Transporters as Determinant Factors…

463

HP: highly perfused PP: poorly perfused HP HP V HP

A HP

V HP

V

A

V

V PP

A PP

V PP

A

PP PP

Fig. 20.2  Tissue and blood (artery [A] and vein [V]) free drug concentrations after reaching steady state from chronic administration. Highly or poorly perfused tissues according to the fraction of cardiac output received in relation to their respective aqueous contents. Hatched blocks represent binding to tissue constituents (e.g., macromolecules) in both blood and tissues

Figure 20.2 illustrates the distribution of molecules caused by the asymmetric distribution of the cardiac output between two tissues. The different thicknesses of the boundary layers of the capillary endothelium of both tissues are highlighted. The image represents steady-state concentrations in highly and poorly perfused tissues and in the blood of arteries and veins after chronic drug administration. The black striped blocks represent non-aqueous components (lipids, proteins, etc.) into which the drug is distributed or bound. Equal colors and intensities refer to the same free concentrations. The veins that come out of each tissue have been not colored because they will be the subject of analysis later. Another important detail revealed in Fig. 20.2 is the competition between permeation into the tissue and circulation throughout the bloodstream for estimating tissue uptake of solutes. A classical equation (the Renkin-Crone equation) allows the calculation of the extraction rate (E) of solutes from the blood (Crone 1963).



 P S E   1  e Q    (20.1)

Where P•S is the permeability coefficient (cm.s−1) multiplied by the endothelial surface area (cm2) and Q is the tissue blood flow (cm3.s−1). Thus, it is expected that for Q> > P•S the extraction will be zero, while for Q