Cover Page

Methods and Principles in Medicinal Chemistry

Edited by R. Mannhold, H. Kubinyi, G. Folkers

Editorial Board:

H. Buschmann, H. Timmerman, H. van de Waterbeemd

Previous Volumes of this Series:

Pfannkuch, Friedlieb / Suter-Dick, Laura (Eds.)

Predictive Toxicology

From Vision to Reality

2014

ISBN: 978-3-527-33608-1

Vol. 64

Kirchmair, Johannes (Ed.)

Drug Metabolism Prediction

2014

ISBN: 978-3-527-33566-4

Vol. 63

Vela, José Miguel / Maldonado, Rafael / Hamon, Michel (Eds.)

In vivo Models for Drug Discovery

2014

ISBN: 978-3-527-33328-8

Vol. 62

Liras, Spiros / Bell, Andrew S. (Eds.)

Phosphodiesterases and Their Inhibitors

2014

ISBN: 978-3-527-33219-9

Vol. 61

Hanessian, Stephen (Ed.)

Natural Products in Medicinal Chemistry

2014

ISBN: 978-3-527-33218-2

Vol. 60

Lackey, Karen / Roth, Bruce (Eds.)

Medicinal Chemistry Approaches to Personalized Medicine

2013

ISBN: 978-3-527-33394-3

Vol. 59

Brown, Nathan (Ed.)

Scaffold Hopping in Medicinal Chemistry

2013

ISBN: 978-3-527-33364-6

Vol. 58

Hoffmann, Rémy / Gohier, Arnaud / Pospisil, Pavel (Eds.)

Data Mining in Drug Discovery

2013

ISBN: 978-3-527-32984-7

Vol. 57

Dömling, Alexander (Ed.)

Protein-Protein Interactions in Drug Discovery

2013

ISBN: 978-3-527-33107-9

Vol. 56

Kalgutkar, Amit S. / Dalvie, Deepak / Obach, R. Scott / Smith, Dennis A.

Reactive Drug Metabolites

2012

ISBN: 978-3-527-33085-0

Vol. 55

Edited by
György M. Keserü and David C. Swinney

Thermodynamics and Kinetics of Drug Binding

Wiley Logo

List of Contributors

Eleanor K. H. Allen

Stony Brook University

Department of Chemistry

Institute for Chemical Biology and Drug Discovery

John S. Toll Drive

Stony Brook, NY 11794

USA

Pelin Ayaz

Bayer Healthcare Pharmaceuticals

Lead Discovery Berlin

Müllerstrasße 178

13353 Berlin

Germany

Antoni R. Blaazer

VU University Amsterdam

Division of Medicinal Chemistry

Faculty of Sciences

Amsterdam Institute for Molecules, Medicines and Systems (AIMMS)

De Boelelaan 1083

1081 HV Amsterdam

The Netherlands

Pier Andrea Borea

University of Ferrara

Department of Medical Sciences

Pharmacology section, via Fossato di Mortara 17-19

44121 Ferrara

Italy

Andrew Chang

Stony Brook University

Department of Chemistry

Institute for Chemical Biology and Drug Discovery

John S. Toll Drive

Stony Brook, NY 11794

USA

Robert A. Copeland

Epizyme, Inc.

400 Technology Square

4th Floor

Cambridge, MA 02139

USA

Gareth Davies

Structure and Biophysics

Discovery Sciences

AstraZeneca, Mereside

Alderley Park

Macclesfield

Cheshire SK10 4TG

UK

Iwan J. P. de Esch

VU University Amsterdam

Division of Medicinal Chemistry

Faculty of Sciences

Amsterdam Institute for Molecules, Medicines and Systems (AIMMS)

De Boelelaan 1083

1081 HV Amsterdam

The Netherlands

György G. Ferenczy

Hungarian Academy of Sciences

Research Centre for Natural Sciences

Medicinal Chemistry Research Group

Magyar tudósok körútja 2

1117 Budapest

Hungary

Ernesto Freire

The Johns Hopkins University

Department of Biology

3400 North Charles

Baltimore, MD 21218

USA

Jonathan C. Fuller

Molecular and Cellular Modeling Group

Heidelberg Institute for Theoretical Studies

Schloss-Wolfsbrunnenweg 35

69118 Heidelberg

Germany

Stefania Gessi

University of Ferrara

Department of Medical Sciences

Pharmacology section, via Fossato di Mortara 17-19

44121 Ferrara

Italy

Dong Guo

Leiden University

Department of Medicinal Chemistry

Einsteinweg 55

2333CC Leiden

The Netherlands

Laura H. Heitman

Leiden University

Department of Medicinal Chemistry

Einsteinweg 55

2333CC Leiden

The Netherlands

Geoffrey A. Holdgate

Structure and Biophysics

Discovery Sciences

AstraZeneca, Mereside

Alderley Park

Macclesfield

Cheshire SK10 4TG

UK

Walter Huber

F. Hoffmann-La Roche AG

Pharma Research and Early Development

Grenzacherstrasse 124

4070 Basel

Switzerland

Adriaan P. Ijzerman

Leiden University

Department of Medicinal Chemistry

Einsteinweg 55

2333CC Leiden

The Netherlands

Lyn H. Jones

Chemical Biology Group

WorldWide Medicinal Chemistry

Pfizer

610 Main Street

Cambridge, MA 02139

USA

Kanishk Kapilashrami

Stony Brook University

Department of Chemistry

Institute for Chemical Biology and Drug Discovery

John S. Toll Drive

Stony Brook, NY 11794

USA

György M. Keserü

Hungarian Academy of Sciences

Research Centre for Natural Sciences

Magyar Tudósok körútja 2

1117 Budapest

Hungary

Daria B. Kokh

Molecular and Cellular Modeling Group

Heidelberg Institute for Theoretical Studies

Schloss-Wolfsbrunnenweg 35

69118 Heidelberg

Germany

Andrew G. Leach

School of Pharmacy and Biomolecular Sciences

Liverpool John Moores University

Byrom Street

Liverpool

Merseyside L3 3AF

UK

Stefania Merighi

University of Ferrara

Department of Medical Sciences

Pharmacology section, via Fossato di Mortara 17-19

44121 Ferrara

Italy

Duncan C. Miller

Newcastle Cancer Centre

School of Chemistry

Newcastle University

Northern Institute for Cancer Research

Bedson Building

Newcastle upon Tyne NE1 7RU

UK

Anke Müller-Fahrnow

Bayer Healthcare Pharmaceuticals

Lead Discovery Berlin

Müllerstraße 178

13353 Berlin

Germany

Julia Romanowska

Molecular and Cellular Modeling Group

Heidelberg Institute for Theoretical Studies

Schloss-Wolfsbrunnenweg 35

69118 Heidelberg

Germany

Felix Schiele

Bayer Healthcare Pharmaceuticals

Lead Discovery Berlin

Müllerstraße 178

13353 Berlin

Germany

Andrew Scott

Molplex Ltd

BioHub @ Alderley Park

Macclesfield

Cheshire SK10 4TG

UK

David C. Swinney

Institute for Rare and Neglected Diseases Drug Discovery

897 Independence Ave

Suite 2C

Mountain View, CA 94043

USA

Peter J. Tonge

Stony Brook University

Department of Chemistry

Institute for Chemical Biology and Drug Discovery

John S. Toll Drive

Stony Brook, NY 11794

USA

Katia Varani

University of Ferrara

Department of Medical Sciences

Pharmacology section, via Fossato di Mortara 17-19

44121 Ferrara

Italy

Georges Vauquelin

Vrije Universiteit Brussel

Department of Molecular and Biochemical Pharmacology

Pleinlaan 2

1050 Brussels

Belgium

Rebecca C. Wade

Molecular and Cellular Modeling Group

Heidelberg Institute for Theoretical Studies

Schloss-Wolfsbrunnenweg 35

69118 Heidelberg

Germany

and

Heidelberg University

Center for Molecular Biology (ZMBH)

DKFZ-ZMBH Alliance and Interdisciplinary

Center for Scientific Computing (IWR)

Im Neuenheimer Feld 282

69120 Heidelberg

Germany

Michael J. Waring

Oncology Medicinal Chemistry

AstraZeneca

Mereside

Alderley Park

Macclesfield

Cheshire SK10 4TG

UK

Preface

In the realm of modern medicine, therapy has become molecular. Understanding and defining the requirements of how a molecular signal is transmitted to cellular chemistry is mainly based on the understanding of the thermodynamics, which governs the journey of the active compound and its interaction with a binding site. The whole field is defined by two remarkably simple, but remarkably true sentences:

  1. Corpora non agunt nisi liquida (Paracelsus)
  2. Corpora non agunt nisi fixata (Paul Ehrlich)

In between those two fundamental concepts, much of the content of the present volume, edited by György Keserü and David Swinney, is located. The thermodynamic perspective of drug action is complex, difficult to be accessed experimentally, and intellectually not easily managed. These are reasons why the whole topic has always been a little bit neglected under the shiny glaze of colorful animated ligands dancing with their receptors.

Switching from “maximizing” affinities in screening campaigns to “optimizing” it requires a deep understanding of the enthalpic and entropic interplay between ligand and receptor. And to make the scenario a little bit more complicated, ligand and its receptor are not alone! Their context provides all kinds of interferences, starting off with “water” and its delicate contribution to binding, going to the membrane, many receptors that are positioned in. Membranes may not only stabilize (or destabilize) conformations of the receptor protein, they also offer secondary binding sites, where ligands may be conformationally preselected to interact with their molecular target: not to talk about membrane traveling peptides in switch control of the receptor proteins or counterions and so on.

It is important to emphasize that this is only one side of the coin. The whole binding process has its kinetic perspective as well. How long, for instance, a drug molecule resides at the binding site is of utmost importance to know for translation into the clinics.

The rich collection of chapters presented in this book touches many of those problems and comes in two parts to cover thermodynamics in the first part and kinetics in the second part. It has the merit of doing this with the perspective of application because this is a “handbook series.” Hence, we learn in addition to some theoretical excursions a lot from case studies and very practical descriptions of how to approach reliable binding parameters experimentally, discern enthalpic and entropic parts, and transfer this knowledge into practical design by selecting a proper substituent located at the proper site of the ligand.

Not least because of this, the series editors are indebted to György Keserü and David Swinney as well as the chapter authors, who made it possible to cover this very essential issue.

We are as well very much indebted to Heike Nöthe, Waltraud Wüst, and Frank Weinreich, all at Wiley-VCH. Their support and ongoing engagement, not only for this book but also for the whole series Methods and Principles in Medicinal Chemistry, adds to the success of this excellent collection of monographs on various topics, all related to drug research.

December 2014

Gerd Folkers, Zürich

Hugo Kubinyi, Weisenheim am Sand

Raimund Mannhold, Düsseldorf

A Personal Foreword

There are many aspects of drug discovery that can be addressed to increase its lower than expected productivity. Understanding the thermodynamics and kinetics of drug action can provide opportunities to help identify effective new medicines and increase productivity. Drug action begins with an interaction of medicines with physiological proteins, known as drug targets. This interaction initiates a series of molecular events that must ultimately communicate a safe, therapeutically useful pharmacological response that corrects the pathophysiology. The molecular details of the response are, in part, dependent on the thermodynamics and binding kinetics.

Although Paul Erlich received the 1908 Nobel Prize for Physiology or Medicine for his contribution to immunology, one of the most impactful results of the father of chemotherapy is summarized in his famous maxim “Corpora non agunt nisi fixata,” which translated becomes – a substance is not (biologically) active unless it is “fixed” (bound to a biological macromolecule) in 1913. The formation of a ligand–macromolecule complex, often qualitatively described as the process of molecular recognition, is typically realized by specific interactions between the partners. Designing, understanding, and improving these interactions require quantitative measures that describe the energetics of complex formation. Binding thermodynamics that governs the process of molecular recognition has therefore a key role in characterizing and optimizing ligand–target interactions, and consequently, its exploitation might contribute to more efficient design of new medicines.

From a thermodynamic perspective, the main driving force of the formation of the ligand–target complex is the change in free energy of binding (ΔG). Since ΔG has two components, the binding enthalpy (ΔH) and the binding entropy (ΔS), one can improve ΔG both enthalpically and entropically. Recent efforts collected to the thermodynamic section of this book are trying to rationalize enthalpic and entropic contributions of ligand binding. Here, we first introduce the methodologies available for the evaluation of binding thermodynamics that include isothermal titration calorimetry, van't Hoff analysis, and computational approaches. The next chapter focuses on uncovering structure–thermodynamics relationship that is one of the most challenging parts of thermodynamics based on lead discovery and optimization. Finally, the authors coming from real-life drug discovery settings discuss the impact of binding thermodynamics studies on drug discovery programs.

Evaluation of binding thermodynamics contributes many aspects of drug discovery. Early-phase discovery programs might benefit identifying chemical starting points with enthalpy-driven binding. Fragment-based drug discovery is a typical example of this approach, demonstrating that the binding of most fragment hits is enthalpy driven. Later phase programs might utilize thermodynamic characterization when selecting compounds at milestones such as the identification of lead molecules, advanced leads, and development candidates. There is increasing evidence that binding thermodynamics influences not only the binding affinity but also selectivity, specificity, and drug-like properties. Considering all of these factors, we can conclude that thermodynamic characterization of discovery compounds might contribute to improving compound quality, and therefore could help making the preclinical phase of drug discovery more productive.

The importance of kinetics to a response has long been recognized. The concept of binding kinetics dates back to work in the 1960s by William Paton, one of the pioneers of pharmacology. In one paper, Paton postulated a rate theory, which uses the interaction of a drug with its receptor to explain drug action, potency, and speed of offset. Recent retrospective analyses have proposed that a drug's dissociation rate from the receptor, koff, also known as residence time, 1/koff, is associated with the evolution of optimal efficacy, safety, and drug use within therapeutic classes. A greater understanding of binding kinetics may create opportunities for more efficient optimization of molecules into medicines.

To evaluate and exploit potential opportunities, a number of questions have to be addressed. Of primary importance to medicinal chemists is the understanding of structure–kinetic relationships (SKR) and how binding kinetics translates to clinical utility. This will be enabled by reliable assays and systematic analysis of SKRs. They will help address questions of can binding kinetics be optimized prospectively? And, can we predict how kinetics will translate to clinical responses. To date, there are few reports of systematic analysis of SKR to inform design principles, and there is uncertainty how to realize the full value of binding kinetics.

The study and use of binding kinetics is currently getting more attention as evidenced by its inclusion in this book series. The understanding of binding kinetics, the opportunities, and the value are evolving. Binding kinetics has the potential to impact many aspects of drug discovery, pharmacology, and medicine. First, the increased awareness of the role of time-dependent processes and dynamics will inform experimental design and interpretation. Medical researchers from all disciplines will be empowered by thinking in terms of kinetics in addition to equilibrium thermodynamics. Second, we think that further understanding of the molecular features governing association and dissociation of a drug with its target will facilitate rational drug design and understanding the molecular mechanisms of drug action. It is clear that the equilibrium dissociation constant can be influenced by both kon and koff. Third, understanding binding kinetics has the potential to better inform clinical pharmacology and understanding and optimizing PK/PD relationships. And last, binding kinetic has the potential to increase productivity by contributing to an optimal therapeutic index. A better understanding of how to early predict and optimize binding kinetics to provide an optimal therapeutic index should help decrease attrition in clinical studies. For example, medicines that have the potential for mechanism-based toxicity may benefit from fast off rates, whereas medicines without the potential for mechanism-based toxicity may benefit for very slow off rates that create irreversible/insurmountable pharmacological behavior.

Addressing both the thermodynamic and kinetic aspects of ligand binding provides opportunities for medicinal chemistry, computational chemistry, computational biology, structural chemistry and biology, analytical chemistry, and pharmacology. Clarity on first principles, methods of analysis, medicinal chemistry design, and translation to clinical pharmacology are all important. To this end, leaders in the study of binding thermodynamics and kinetics have contributed chapters that describe their current understandings. It is clear from the breadth of examples that binding thermodynamics and kinetics are an important features of drug action, and that there are many opportunities to further understand and use them in drug discovery. The challenge is to prospectively apply the knowledge to maximize the value of the opportunities.

We would like to acknowledge to all contributing authors for sharing their knowledge and perspective on the thermodynamic and kinetic aspects of ligand binding, we thank the series editors Raimund Mannhold, Hugo Kubinyi, and Gerd Folkers for the opportunity addressing the topic, and Frank Weinreich, Gregor Cicchetti, and Waltraud Wuest at Wiley-VCH for their support and commitment.

December 2014

György M. Keserű

Hungary

David C. Swinney

USA

Section I

Thermodynamics

1
The Binding Thermodynamics of Drug Candidates

Ernesto Freire

1.1 Affinity Optimization

The affinity optimization of drug candidates is a major goal in drug development. Most often, the starting points for optimization are compounds or fragments identified in screening campaigns. For full-size compounds, the top hits usually have binding affinities in the mid-micromolar range, while for fragments, hits with affinities as weak as millimolar are not uncommon. In both cases, the binding affinity needs to be improved by 5 orders of magnitude or more for a hit to become a reliable drug candidate. Five orders of magnitude improvement in affinity is equivalent to an additional binding energy of −7.0 kcal mol−1G = −RT ln(1/Kd)); that is, essentially doubling the binding energy of the starting compound. Performing this task in an efficient way and simultaneously improving or maintaining drug-like properties is not an easy task and, arguably, can be facilitated by an in-depth knowledge of the binding thermodynamics of a compound.

Affinity optimization is not a simple task because it needs to adhere to constraints that maintain or improve the drug-like character of the compound. A common framework is given by the Lipinski rules of five [1, 2], which limit the molecular weight and the number and type of functionalities that are present in the final compound. For screening hits that already have molecular weights around 500, improving the affinity to the required drug levels essentially means doubling the ligand efficiency (LE = ΔG/(number of heavy atoms)). For fragments (MW ∼ 200), it means that the chemical functionalities that are added to grow the compound must have a better LE than the starting fragment. Furthermore and in addition to binding affinity, other binding-related properties like selectivity or susceptibility to drug-resistant mutations need to be addressed.

Recently, researchers have become aware of the tendency for new drug candidates to be excessively hydrophobic, to exhibit low solubility and poor permeability, and correspondingly exhibit poor drug quality [3]. In order to identify high quality compounds at an early stage or to improve the quality of existing leads, different metrics have been proposed. It has been realized that high quality compounds are those characterized by high potency and simultaneously low hydrophobicity [4]. In fact, if a plot is made of the logarithm of the potency of the compounds versus their ClogP (Figure 1.1) the high quality compounds cluster in the upper left corner. Those compounds are said to have a high lipophilic efficiency (LipE defined as pKd–ClogP; for any given series pIC50 or pKi can also be used in the analysis) [4]. From a fundamental standpoint, an important issue is to assess whether LipE and similar metrics have a solid thermodynamic foundation and how they can be implemented in a prospective way. This is the main topic of this chapter.

nfgz001

Figure 1.1 The logarithm of the binding affinity as a function of ClogP for a series of protease inhibitors analogs. The solid lines represent lines of constant LipE (indicated by the numbers). Compounds with the higher LipE arguably display the best drug-like properties.

1.2 The Binding Affinity

The binding affinity is dictated by the Gibbs energy of binding (ΔG = −RTlnKa or ΔG = −RTln(1/Kd)), which in turn is the sum of the binding enthalpy (ΔH) and the binding entropy contribution (−TΔS), as shown in Figure 1.2. The bar graph in Figure 1.2, is called the thermodynamic signature [5, 6], and provides an instantaneous visual representation of the magnitude of the different interactions that contribute to binding. The thermodynamic signature can be measured by isothermal titration calorimetry (ITC) at any given temperature. Of all the techniques available to measure binding, ITC is the only one capable of measuring not only binding affinities but also the thermodynamic parameters that determine the binding energy. Since the enthalpy and entropy changes originate from different types of interactions, having access to those quantities provide an indication of the forces that drive the binding of a compound, and simultaneously delineate paths for optimization. In the past, ITC has been primarily used retrospectively rather than prospectively as a guiding tool for lead optimization. This situation is changing due to two factors: improved understanding of the relationships between thermodynamic forces (enthalpy, entropy, and heat capacity) and chemical structure, and a new generation of instruments with reduced sample requirements, better sensitivity, and much faster throughputs [7, 8].

nfgz002

Figure 1.2 The thermodynamic signature is a representation of the binding thermodynamic parameters that permit a rapid assessment of the forces that determine the binding of a compound.

A nanomolar affinity corresponds to a Gibbs energy close to −12.5 kcal mol−1, while a picomolar affinity corresponds to a Gibbs energy close to −16.5 kcal mol−1. In a typical scenario, a common design goal is engineering a compound with a binding affinity on the order of 0.1 nM, which is equivalent to a Gibbs energy of −14 kcal mol−1. If the starting compound identified in a screen has a 10 μM affinity (equivalent to −6.8 kcal mol−1 as in the example in Figure 1.2) its affinity optimization will require an additional −7 kcal mol−1. This additional −7 kcal mol−1 of binding affinity can be achieved by any possible enthalpy and entropy combinations that add up to the required amount. While in a test tube, the precise enthalpy/entropy balance may be irrelevant, in real life it is very important as the enthalpy and entropy changes originate from different types of interactions and, consequently, compounds with different thermodynamic signatures will have otherwise different properties. From the point of view of drug development, the particular way by which the additional −7 kcal mol−1 are achieved will determine to a large extent the drug quality of the compound.

1.3 The Enthalpy Change

Drug molecules are composed of polar and nonpolar (carbon) atoms and they contribute very differently to the enthalpy change. In binding, two processes occur simultaneously: desolvation and the formation of drug/protein interactions. The desolvation of polar groups is highly unfavorable whereas the desolvation of nonpolar groups is favorable. There is a penalty of 6.2 kcal mol−1 for desolvating a hydroxyl group, for example, and a gain of −0.76 kcal mol−1 for desolvating a methyl group. The penalty of 6.2 kcal mol−1 is equivalent to more than 4 orders of magnitude in binding affinity, indicating that unless polar groups establish very strong interactions with the protein they are going to contribute unfavorably or very little to binding affinity. Furthermore, the desolvation penalty of polar groups is of an enthalpic origin. In fact, the desolvation enthalpies of polar groups are on the order of 5–9 kcal mol−1, which is about 1 order of magnitude higher than that of nonpolar groups. Unfavorable binding enthalpies are usually associated with polar groups that become desolvated and do not establish strong interactions with the protein. Table 1.1 summarizes enthalpies for the desolvation (transfer from aqueous solution to gas phase) [9] of common chemical functionalities used in lead optimization.

Table 1.1 Desolvation free energies and enthalpies for chemical functionalities commonly used in lead optimizationa

Group ΔG desolvation (kcal mol−1) ΔH desolvation (kcal mol−1)
NH2 5.8 7.9
NH 6.0 9.4
N 5.9 9.3
Naromatic 4.1 4.9
NO2 4.8 4.7
O 3.8 5.2
OH 6.2 8.7
CO 5.5 5.5
COO 4.9 5.4
COOH 7.7 8.4
CH3 −0.76 0.57
CH2 −0.18 0.77

Values from Cabani et al. [9].

a Enthalpies associated with the transfer of different chemical functionalities from aqueous solution to the gas phase.

The thermodynamic signature provides a rapid evaluation of the enthalpic contribution to binding. If the enthalpy is unfavorable, the first task is to identify the origin of this behavior and, if necessary, localize the problematic polar groups and eliminate them (see, for example, [10, 11]). If high-resolution structural information is not available, conventional structure activity relationships based upon the thermodynamic signature of compounds can provide the required information to identify unwanted polar groups.

Polar groups that establish strong hydrogen bonds with the protein, usually contribute favorably on the order of −4 to −5 kcal mol−1 to the binding enthalpy [12]. Figure 1.3 shows the thermodynamic signature for two pairs of compounds that vary by a single functionality. In the top panel, the change of a thioether to a sulfonyl results in a strong hydrogen bond and an enthalpic gain of 3.9 kcal mol−1. In the bottom panel, the change of a methyl group to a hydroxyl group results in an enthalpic gain of 4.4 kcal mol−1. It must be noted, however, that in both cases the enthalpic gains are accompanied by compensating entropic losses. In the first case, the entropic loss is larger than the enthalpic gain resulting in a small drop in binding affinity. In the second case, the entropic loss is smaller than the enthalpic gain resulting in a twofold gain in binding affinity. These results provide the rationale for the common observation in lead optimization that the introduction of polar groups that establish strong hydrogen bonds often results in no binding-affinity gains. These results also indicate that in order to obtain binding-affinity gains with polar groups, it is necessary to overcome the ubiquitous phenomenon of enthalpy/entropy compensation [12].

nfgz003

Figure 1.3 The thermodynamic signatures for two pairs of protease inhibitors that vary by the addition of a single polar group that establish a strong hydrogen bond with the target. The formation of the hydrogen bond results in an enthalpic gain of −4 to −5 kcal mol−1. The enthalpic gain is compensated to different extents by opposite entropy loses [12].

Contrary to polar groups, the introduction of nonpolar groups usually results in small enthalpy and entropy gains that bring about moderate improvements in binding affinity. The cumulative improvements may result in high affinity but also highly hydrophobic compounds. Figure 1.4 shows a typical situation in which a methyl group is added to a compound. In this case, favorable contributions of 0.8 and 0.5 kcal from the enthalpy and entropy changes result in a binding-affinity gain of 8.7-fold. In our work, we have observed situations like this many times and provide a rationale to the practice of using (or abusing) hydrophobic groups to improve binding affinity. The introduction of nonpolar functionalities is devoid of the large compensatory enthalpy/entropy changes that greatly difficult lead optimization with polar groups.

nfgz004

Figure 1.4 The thermodynamic signature for a pair of protease inhibitors that vary by the addition of a single methyl group. A well-packed methyl group is usually associated with small enthalpy and entropy gains resulting in a binding affinity gain [13].

1.4 The Entropy Change

Two terms are of major concern from an engineering point of view: the solvation/desolvation entropy and the conformational entropy. The desolvation of both polar and nonpolar groups is favorable to binding [9]. On the other hand, the ordering or structuring of side chains or backbone in the protein or the drug molecule contributes unfavorably to the binding entropy. In most situations, the desolvation entropy dominates and the observed binding entropy is favorable. The exceptions are those cases in which the binding reaction is coupled to large structuring processes like the refolding of disordered domains in proteins [14]. This situation is illustrated in Figure 1.5 for the binding of the inhibitor NBD-556

nfgz005

Figure 1.5 The thermodynamic signature of NBD-556a HIV-1 gp120/CD4 inhibitor. The binding of this inhibitor is coupled to a large refolding (structuring) of gp120 and consequently the thermodynamic signature is dominated by a large unfavorable binding entropy and an equally large favorable enthalpy attributed to the refolding of the protein [14].

to gp120, the envelope glycoprotein of HIV-1 [14]. In this case, the large negative enthalpy and large unfavorable entropy are associated with the structuring of regions of gp120 that are intrinsically disordered in the unliganded state.

Because nonpolar groups exhibit a favorable entropy of desolvation and a very small enthalpic penalty that can be easily overcome by van der Waals interactions with the target, they have been a favorite tool for optimization [6, 15, 16], and consequently drug candidates have become more hydrophobic in recent years. The binding of these compounds is dominated by large favorable entropies and enthalpies that are often unfavorable or only slightly favorable. In fact, it was noticed earlier for HIV-1 protease inhibitors as well as statins [6, 15, 16] that enthalpically optimized compounds appeared only after the first-in-class drugs had already been in the market for a while. This observation is a testimony to the difficulties in enthalpic optimization.

1.5 Engineering Binding Contributions

It is obvious that the main complication during optimization arises from the introduction (location and type) of polar groups. There are different situations in which polar groups can be found. (i) Polar groups can be introduced as solubilizers of hydrophobic compounds. To be effective solubilizers, they should remain exposed to the solvent and not pay the desolvation penalty. (ii) Polar groups that become desolvated but do not establish strong interactions with the protein. These groups can be identified by their unfavorable contribution to the binding enthalpy; they should be eliminated as they only pay the desolvation penalty without contributing favorably to binding. (iii) Polar groups that establish strong hydrogen bonds but do not contribute to binding affinity. These groups are characterized by strong favorable contributions to the binding enthalpy (−4 to −5 kcal mol−1, Figure 1.3) but equally large entropic compensation that negates any gain in affinity. Even though these groups do not contribute to affinity, they do contribute to selectivity and should be preserved [13]. (iv) Polar groups that establish strong hydrogen bonds and contribute favorably to the binding affinity. These groups are the most important ones as they contribute both to binding affinity and selectivity.

As mentioned earlier, the enthalpy gain associated with strong hydrogen bonds can be opposed by a large entropy loss resulting in no gain or even a loss in binding affinity. Often, the origin of the entropy loss is due to: (i) the structuring effect triggered by the newly formed hydrogen bond (loss in conformational entropy); and/or (ii) losses in desolvation entropy if the new hydrogen bond forces some groups to be more exposed to water.

Improving binding affinity with hydrogen bonds (i.e., polar groups) is difficult because it requires overcoming large unfavorable entropic effects. Hydrogen bonds should be directed to structured regions of the protein in order to minimize structuring effects and their associated compensating entropy changes. If the crystallographic structure of the target protein is known, structured regions can be identified either computationally or by examining the B-factors in the pdb structure file [12]. On the other hand, unfavorable entropy changes due to diminished desolvation can be overcome by modifying the size/geometry of the group or its stereochemistry [17].

The effects on binding affinity of polar groups that establish strong hydrogen bonds are always the difference between a large favorable enthalpy and a large unfavorable entropy changes. For example, a strong hydrogen bond that improves binding affinity by 1 order of magnitude (−1.4 kcal mol−1) will often contribute a favorable enthalpy of −4.5 kcal mol−1 and be opposed by an unfavorable entropy contribution of 3.1 kcal mol−1. On the other hand, hydrophobic functionalities are usually characterized by small favorable enthalpy and entropy changes. Thermodynamically, it is not surprising that hydrophobic groups represent the fastest way of optimizing affinity. This example illustrates the qualitative difference of improving affinity by nonpolar and polar functionalities. Enthalpy, being more difficult to optimize, has appeared only as the dominant driving force in second generation or “best in class” drugs [6]. Obviously, problems associated with highly hydrophobic compounds like solubility, bioavailability, and selectivity, to name a few, have been recognized and provide the rationale and the incentive for the development of enthalpically optimized compounds.

1.6 Lipophilic Efficiency and Binding Enthalpy

Figure 1.6

nfgz006

Figure 1.6 Correlation of binding enthalpy with LipE for FDA approved HIV-1 protease inhibitors (a), statins (b), and a series of HIV-1 protease analogs belonging to the same chemical scaffold (c).

Acknowledgments

This work was supported by grants from the National Institutes of Health (GM56550 and GM57144) and the National Science Foundation (MCB0641252).

References

  1. 1. Lipinski, C.A. (2000) Drug-like properties and the causes of poor solubility and poor permeability. Journal of Pharmacological and Toxicological Methods, 44, 235–249.
  2. 2. Lipinski, C.A. (2003) in Molecular Informatics Confronting Complexity. Proceedings of the Beilstein Institute Workshop, May 13–16, 2002 (eds C. Kettner and M.G. Hicks), Logos Verlag Berlin, Bozen, pp. 59–78.
  3. 3. Leeson, P.D. and Springthorpe, B. (2007) The influence of drug-like concepts on decision-making in medicinal chemistry. Nature Reviews Drug Discovery, 6, 881–890.
  4. 4. Ryckmansa, T., Edwards, M.P., Hornea, V.A., Correia, A.M., Owena, D.R., Thompson, L.R., Trana, I., Tuttc, M.F., and Young, T. (2009) Rapid assessment of a novel series of selective CB2 agonists using parallel synthesis protocols: a Lipophilic Efficiency (LipE) analysis. Bioorganic and Medicinal Chemistry, 19, 4406–4409.
  5. 5. Velazquez-Campoy, A., Kiso, Y., and Freire, E. (2001) The binding energetics of first and second generation HIV-1 protease inhibitors: implications for drug design. Archives of Biochemistry and Biophysics, 390, 169–175.
  6. 6. Freire, E. (2008) Do enthalpy and entropy distinguish first in class from best in class? Drug Discovery Today, 13, 869–874.
  7. 7. Velazquez Campoy, A. and Freire, E. (2005) ITC in the post-genomic era…? Priceless. Biophysical Chemistry, 115, 115–124.
  8. 8. Velazquez-Campoy, A. and Freire, E. (2006) Isothermal titration calorimetry to determine association constants for high-affinity ligands. Nature Protocols, 1, 186–191.
  9. 9. Cabani, S., Gianni, P., Mollica, V., and Lepori, L. (1981) Group contributions to the thermodynamic properties of non-ionic organic solutes in dilute aqueous solution. Journal of Solution Chemistry, 10, 563–595.
  10. 10. Sarver, R.W., Peevers, J., Cody, W.L., Ciske, F.L., Dyer, J., Emerson, S.D., Hagadorn, J.C., Holsworth, D.D., Jalaie, M., Kaufman, M., Mastronardi, M., McConnell, P., Powell, N.A., Quin, J. III, Van Huis, C.A., Zhang, E., and Mochalkin, I. (2007) Binding thermodynamics of substituted diaminopyrimidine renin inhibitors. Analytical Biochemistry, 360, 30–40.
  11. 11. Sarver, R.W., Bills, E., Bolton, G., Bratton, L.D., Caspers, N.L., Dunbar, J.B., Harris, M.S., Hutchings, R.H., Kennedy, R.M., Larsen, S.D., Pavlovsky, A., Pfefferkorn, J.A., and Bainbridge, G. (2008) Thermodynamic and structure guided design of statin based inhibitors of 3-hydroxy-3-methylglutaryl coenzyme A reductase. Journal of Medicinal Chemistry, 51, 3804–3813.
  12. 12. Lafont, V., Armstrong, A.A., Ohtaka, H., Kiso, Y., Mario Amzel, L., and Freire, E. (2007) Compensating enthalpic and entropic changes hinder binding affinity optimization. Chemical Biology and Drug Design, 69, 413–422.
  13. 13. Kawasaki, Y. and Freire, E. (2011) Finding a better path to drug selectivity. Drug Discovery Today, 16, 985–990.
  14. 14. Schon, A., Madani, N., Klein, J., Hubicki, A., Ng, D., Yang, X., Smith, A.B. III, Sodroski, J., and Freire, E. (2006) Binding thermodynamics of a small-molecular-weight CD4 mimetic to HIV-1 gp120. Biochemistry, 45, 10973–10980.
  15. 15. Ohtaka, H. and Freire, E. (2005) Adaptive Inhibitors of the HIV-1 protease. Progress in Biophysics and Molecular Biology, 88, 193–208.
  16. 16. Carbonell, T. and Freire, E. (2005) Binding thermodynamics of statins to HMG-CoA reductase. Biochemistry, 44, 11741–11748.
  17. 17. Freire, E. (2009) A thermodynamic approach to the affinity optimization of drug candidates. Chemical Biology and Drug Design, 74, 468–472.