throbber
The Organic Chemistry of
`Drug Design and Drug Action
`
`Richard B. Silverman
`Department of Chemistry
`Northwestern University
`Evanston, Illinois
`
`Ap
`
`Academic Press, Inc.
`Harcourt Brace Jovanovich,Publishers
`San Diego New York Boston London Sydney Tokyo Toronto
`
`MYLAN EXHIBIT - 1050
`Mylan Pharmaceuticals, Inc. v. Bausch Health Ireland, Ltd. - IPR2022-00722
`
`

`

`This book is printed on acid-free paper. 0
`
`Copyright © 1992 by ACADEMIC PRESS, INC.
`All Rights Reserved.
`No part of this publication may be reproduced or transmitted in any form or by any
`means, electronic or mechanical, including photocopy, recording, or any information
`storage and retrieval system, without permission in writing from the publisher.
`
`Academic Press, Inc.
`1250 Sixth Avenge, San Diego, California 92101-4311
`
`United Kingdom Edition published by
`Academic Press Limited
`24-28 Oval Road, London NW1 7DX
`
`Library of Congress Cataloging-in-Publication Data
`
`Silverman, Richard B.
`The organic chemistry of drug design and drug action / Richard B.
`Silverman.
`p. cm.
`Includes index.
`ISBN 0-12-643730-0 (hardcover)
`1. Pharmaceutical chemistry. 2. Bioorganic chemistry.
`I. Title.
`3. Molecular pharmacology. 4. Drugs--Design.
`[DNLM: 1. Chemistry, Organic. 2. Chemistry, Pharmaceutical.
`3. Drug Design. 4. Phannacokinetics. QV 744 S587o]
`RS403.S55 1992
`615'.19--dc20
`DNLM/DLC
`for Library of Congress
`
`PRINTED IN THE UNITED STATES OF AMERICA
`9 8 7 6 5 4 3 2 1
`HA
`92 93 94 95 96 97
`
`91-47041
`CIP
`
`

`

`To Mom and the memory of Dad,
`for their warmth, their humor, their ethics, their inspiration,
`but mostly for their genes.
`
`

`

`CHAPTER 3
`Receptors
`
`53
`
`54
`
`52
`I. Introduction
`II. Receptor Structure
`A. Historical
`53
`54
`B. What Is a Receptor?
`III. Drug—Receptor Interactions
`A. General Considerations
`54
`55
`B. Forces Involved in the Drug—Receptor Complex
`1. Covalent Bonds, 56 • 2. Ionic (or Electrostatic) Interactions, 56 • 3. Ion-Dipole
`and Dipole-Dipole Interactions, 56 • 4. Hydrogen Bonds, 57 • 5. Charge-Transfer
`Complexes, 60 • 6. Hydrophobic Interactions, 60 • 7. Van der Waals or London
`Dispersion Forces, 61 • 8. Conclusion, 62
`C. Ionization
`62
`D. Determination of Drug—Receptor Interactions
`71
`E. Drug—Receptor Theories
`1. Occupancy Theory, 71 • 2. Rate Theory, 72 • 3. Induced-Fit Theory, 72 •
`4. Macromolecular Perturbation Theory, 73 • 5. Activation—Aggregation Theory, 74
`74
`F. Topographical and Stereochemical Considerations
`1. Spatial Arrangement of Atoms, 75 • 2. Drug and Receptor Chirality, 76 •
`3. Geometric Isomers, 82 • 4. Conformational Isomers, 83 • 5. Ring Topology, 86
`G. Ion Channel Blockers
`87
`H. Example of Rational Drug Design of a Receptor Antagonist:
`88
`Cimetidine
`
`63
`
`95
`References
`General References
`
`97
`
`I. Introduction
`
`Up to this point in our discussion it appears that a drug is taken, and by some
`kind of magic it travels through the body and elicits a pharmaceutical effect.
`Pharmacokinetics (absorption, distribution, metabolism, and excretion) was
`mentioned in Chapter 2, but no discussion was presented regarding what
`produces the pharmaceutical effect. The site of drug action, which is ulti-
`mately responsible for the pharmaceutical effect, is called a receptor. The
`interaction of the drug with the receptor constitutes pharmacodynamics. In
`this chapter the emphasis is placed on pharmacodynamics of general noncata-
`lytic receptors, in Chapter 4 a special class of receptors that have catalytic
`
`52
`
`

`

`II. Receptor Structure
`
`53
`
`properties, called enzymes, will be discussed, and in Chapter 6 another recep-
`tor, DNA, will be the topic of discussion. The drug—receptor properties de-
`scribed in this chapter also apply to drug-enzyme and drug-DNA complexes.
`
`II. Receptor Structure
`A. Historical
`
`In 1878 John N. Langley,' a physiology student at Cambridge University,
`while studying the mutually antagonistic action of the alkaloids atropine (3.1;
`now used as an antisecretory agent) and pilocarpine (3.2; used in the treat-
`ment of glaucoma, but causes sweating and salivation) on cat salivary flow,
`suggested that both of these chemicals interacted with some substance in the
`nerve endings of the gland cells. Langley, however, did not follow up this
`notion for over 25 years.
`
`, CH 3
`
`CH2OH
`
`0
`
`3.1
`
`C2Hf
`
`"CH
`
`CH3
`
`3.2
`
`Paul Ehrlich2 suggested his side chain theory in 1897. According to this
`hypothesis, cells have side chains attached to them that contain specific
`groups capable of combining with a particular group of a toxin. Ehrlich termed
`these side chains receptors. Another facet of this hypothesis was that when
`toxins combined with the side chains, excess side chains were produced and
`released into the bloodstream. In today's biochemical vernacular these excess
`side chains would be called antibodies, and they combine with toxins stoi-
`chiometrically.
`In 1905 and 1906 Langley' studied the antagonistic effects of curare (a
`generic term for a variety of South American quaternary alkaloid poisons that
`cause muscular paralysis) on nicotine stimulation of skeletal muscle. He con-
`cluded that there was a receptive substance that received the stimulus and, by
`transmitting it, caused muscle contraction. This was really the first time that
`attention was drawn to the two fundamental characteristics of a receptor,
`namely, a recognition capacity for specific ligands and an amplification com-
`ponent, the ability of the ligand—receptor complex to initiate a biological
`response.
`
`

`

`54
`
`B. What Is a Receptor?
`
`3. Receptors
`
`In general, receptors are integral proteins (i.e., polypeptide macromolecules)
`that are embedded in the phospholipid bilayer of cell membranes (see Fig.
`2.3). They, typically, function in the membrane environment; consequently,
`their properties and mechanisms of action depend on the phospholipid milieu.
`Vigorous treatment of cells with detergents is required to dissociate these
`proteins from the membrane. Once they become dissociated, however, they
`can lose their integrity. Since they generally exist in minute quantities and can
`be unstable, few receptors have been purified, and little structural information
`is known about them. Advances in molecular biology more recently have
`permitted the isolation, cloning, and sequencing of receptors,4 and this is
`leading to further approaches to molecular characterization of these proteins.
`However, these receptors, unlike many enzymes, are still typically character-
`ized in terms of their function rather than by their structural properties. The
`two functional components of receptors, the recognition component and the
`amplification component, may represent the same or different sites on the
`same protein. Various hypotheses regarding the mechanism by which drugs
`may initiate a biological response are discussed in Section III,E.
`
`III. Drug—Receptor Interactions
`A. General Considerations
`
`In order to appreciate mechanisms of drug action it is important to understand
`the forces of interaction that bind drugs to their receptors. Because of the low
`concentration of drugs and receptors in the bloodstream and other biological
`fluids, the law of mass action alone cannot account for the ability of small
`doses of structurally specific drugs to elicit a total response by combination
`with all, or practically all, of the appropriate receptors. One of my all-time
`favorite calculations, shown below, supports the notion that something more
`than mass action is required to get the desired drug—receptor interaction.5
`One mole of a drug contains 6.02 x 1023 molecules (Avogadro's number). If
`the molecular weight of an average drug is 200 g/mol, then 1 mg (often an
`effective dose) will contain 6.02 x 1023(10-3)/200 = 3 x 1018 molecules of
`drug. The human organism is composed of about 3 x 1013 cells. Therefore,
`each cell will be acted upon by 3 x 1018/3 x 1013 = 105 drug molecules. One
`erythrocyte cell contains about 1010 molecules. On the assumption that the
`same number of molecules is found in all cells, then for each drug molecule,
`there are 101°/105 = 105 molecules of the human body! With this ratio of
`human molecules to drug molecules, Le Chatelier would have a difficult time
`explaining how the drug could interact and form a stable complex with the
`desired receptor.
`
`

`

`III. Drug—Receptor Interactions
`
`55
`
`The driving force for the drug—receptor interaction can be considered as a
`low-energy state of the drug—receptor complex [Eq. (3.1)], where kon is the
`rate constant for formation of the drug-receptor complex, which depends on
`the concentrations of the drug and the receptor, and /coif is the rate constant for
`breakdown of the complex, which depends on the concentration of the drug—
`receptor complex as well as other forces. The biological activity of a drug is
`related to its affinity for the receptor, which is measured by its KD , the
`dissociation constant at equilibrium [Eq. (3.2)]. Note that KD is a dissociation
`constant, so that the smaller the KD , the larger the concentration of the drug—
`receptor complex, and the greater is the affinity of the drug for the receptor.
`
`Drug + receptor
`
`Icon
`
`kofr
`
`drug—receptor complex
`
`KD
`
`[drug][receptor]
`[drug—receptor complex]
`
`(3.1)
`
`(3.2)
`
`B. Forces Involved in the Drug-Receptor Complex
`
`The forces involved in the drug—receptor complex are the same forces experi-
`enced by all interacting organic molecules and include covalent bonding, ionic
`(electrostatic) interactions, ion-dipole and dipole-dipole interactions, hydro-
`gen bonding, charge-transfer interactions, hydrophobic interactions, and van
`der Waals interactions. Weak interactions usually are possible only when
`molecular surfaces are close and complementary, that is, bond strength is
`distance dependent. The spontaneous formation of a bond between atoms
`occurs with a decrease in free energy, that is, AG is negative. The change in
`free energy is related to the binding equilibrium constant (Keq) by Eq. (3.3).
`Therefore, at physiological temperature (37°C) changes in free energy of —2 to
`—3 kcal/mol can have a major effect on the establishment of good secondary
`interactions. In fact, a decrease in AG° of —2.7 kcal/mol changes the binding
`equilibrium constant from 1 to 100. If the Keq were only 0.01 (i.e., 1% of the
`equilibrium mixture in the form of the drug—receptor complex), then a AG° of
`interaction of —5.45 kcal/mol would shift the binding equilibrium constant to
`100 (i.e., 99% in the form of the drug—receptor complex).
`(3.3)
`AG° —RT In Keq
`In general, the bonds formed between a drug and a receptor are weak
`noncovalent interactions; consequently, the effects produced are reversible.
`Because of this, a drug becomes inactive as soon as its concentration in the
`extracellular fluids decreases. Often it is desirable for the drug effect to last
`only a limited time so that the pharmacological action can be terminated. In
`the case of CNS stimulants and depressants, for example, a prolonged action
`
`

`

`56
`
`3. Receptors
`
`could be harmful. Sometimes, however, the effect produced by a drug should
`persist, and even be irreversible. For example, it is most desirable for a
`chemotherapeutic agent, a drug that acts selectively on a foreign organism or
`tumor cell, to form an irreversible complex with its receptor so that the drug
`can exert its toxic action for a prolonged period.6 In this case, a covalent bond
`would be desirable.
`In the following subsections the various types of possible drug—receptor
`interactions are discussed briefly. These interactions are applicable to all
`types of receptors, including enzymes and DNA, that are described in this
`book.
`
`1. Covalent Bonds
`
`The covalent bond is the strongest bond, generally worth anywhere from —40
`to —110 kcal/mol in stability. It is seldom formed by a drug—receptor interac-
`tion, except with enzymes and DNA. These bonds will be discussed further in
`Chapters 5 and 6.
`
`2. Ionic (or Electrostatic) Interactions
`
`For protein receptors at physiological pH (generally taken to mean pH 7.4),
`basic groups such as the amino side chains of arginine, lysine, and, to a much
`lesser extent, histidine are protonated and, therefore, provide a cationic envi-
`ronment. Acidic groups, such as the carboxylic acid side chains of aspartic
`acid and glutamic acid, are deprotonated to give anionic groups.
`Drug and receptor groups will be mutually attracted provided they have
`opposite charges. This ionic interaction can be effective at distances farther
`than those required for other types of interactions, and they can persist
`longer. A simple ionic interaction can provide a AG° = —5 kcal/mol which
`declines by the square of the distance between the charges. If this interaction
`is reinforced by other simultaneous interactions, the ionic interaction be-
`comes stronger (AG° = —10 kcal/mol) and persists longer. Acetylcholine is
`used as an example of a molecule that can undergo an ionic interaction (Fig.
`3.1).
`
`3. Ion—Dipole and Dipole—Dipole Interactions
`
`As a result of the greater electronegativity of atoms such as oxygen, nitrogen,
`sulfur, and halogens relative to that of carbon, C—X bonds in drugs and
`receptors, where X is an electronegative atom, will have an asymmetric distri-
`bution of electrons; this produces electronic dipoles. The dipoles in a drug
`molecule can be attracted by ions (ion—dipole interaction) or by other dipoles
`(dipole—dipole interaction) in the receptor, provided charges of opposite sign
`are properly aligned. Since the charge of a dipole is less than that of an ion, a
`
`

`

`III. Drug—Receptor Interactions
`
`57
`
`0
`-o
`+
`II
`CH3COCH2NMe3
`
`Figure 3.1. Example of a simple ionic interaction. The wavy line represents the receptor
`surface.
`
`dipole-dipole interaction is weaker than an ion—dipole interaction. In Fig. 3.2
`acetylcholine is used to demonstrate these interactions, which can provide a
`AG° of —1 to —7 kcal/mol.
`
`4. Hydrogen Bonds
`Hydrogen bonds are a type of dipole—dipole interaction formed between the
`proton of a group X—H, where X is an electronegative atom, and other
`electronegative atoms (Y) containing a pair of nonbonded electrons. The only
`significant hydrogen bonds occur in molecules where X and Y are N, 0, or F.
`X removes electron density from the hydrogen so it has a partial positive
`charge, which is strongly attracted to nonbonded electrons of Y. The interac-
`tion is denoted as a dotted line,
`to indicate that a covalent
`bond between X and H still exists, but that an interaction between H and Y
`also occurs. When X and Y are equivalent in electronegativity and degree of
`ionization, the proton can be shared equally between the two groups, that is,
`
`The hydrogen bond is unique to hydrogen because it is the only atom that
`can carry a positive charge at physiological pH while remaining covalently
`bonded in a molecule, and hydrogen also is small enough to allow close
`approach of a second electronegative atom. The strength of the hydrogen
`bond is related to the Hammett a- constants.'
`There are intramolecular and intermolecular hydrogen bonds; the former
`are stronger (see Fig. 3.3). Hydrogen bonding can be quite important for
`biological activity. For example, methyl salicylate (3.3), an active ingredient
`
`ion-dipole
`
`O8
`
`NH3
`
`+
`I I
`CH3COCH2NMe3
`8+
`45OH
`
`8+
`
`0
`5
`
`5+
`
`dipole-dipole
`
`Figure 3.2. Examples of ion—dipole and dipole—dipole interactions. The wavy line repre-
`sents the receptor surface.
`
`

`

`58
`
`3. Receptors
`
`intramolecular
`
`intermolecular
`
`Figure 3.3. Examples of hydrogen bonds. The wavy lines represents the receptor surface.
`
`in many muscle pain remedies and at least one antiseptic, is a weak antibacte-
`rial agent. The corresponding para isomer, methyl p-hydroxybenzoate (3.4),
`however, is considerably more active as an antibacterial agent and is used as a
`food preservative. It is believed that the antibacterial activity of 3.4 is derived
`from the phenolic hydroxyl group. In 3.3 this group is masked by intramolecu-
`lar hydrogen bonding'
`
`O
`
`OCH3
`3.3
`
`CH3O
`
`\ /
`
`O
`
`3.4
`
`Hydrogen bonds are essential in maintaining the structural integrity of «-
`helix and /3-sheet conformations of peptides and proteins (3.5)1 and the double
`helix of DNA (3.6).2 As discussed in Chapter 6, many antitumor agents act by
`intercalation into the DNA base pairs or by alkylation of the DNA bases,
`thereby preventing hydrogen bonding. This disrupts the double helix and
`destroys the DNA.
`Another instance where hydrogen bonding is suggested to be important
`arises when the potency of various oxygen-containing drugs becomes reduced
`by substitution of a sulfur atom for the oxygen atom in the drug. Sulfur, which
`is very poor at hydrogen bonding relative to oxygen, presumably cannot
`interact with the receptor group that hydrogen bonds to the oxygen, and
`drug—receptor complex stability becomes diminished.
`The AG° for hydrogen bonding can be between —1 and —7 kcal/mol but
`usually is in the range of —3 to —5 kcal/mol.
`
`From B. Alberts, D. Bray, J. Lewis, M. Raff, K. Roberts, and J. D. Watson, "Molecular
`Biology of the Cell," 2nd Ed., pp. 110 and 109, respectively. Garland Publishing, New York,
`1989, with permission. Copyright © 1989 Garland Publishing.
`2 From B. Alberts, D. Bray, J. Lewis, M. Raff, K. Roberts, and J. D. Watson, "Molecular
`Biology of the Cell," 2nd Ed., p. 99. Garland Publishing, New York, 1989, with permission.
`Copyright © 1989 Garland Publishing.
`
`

`

`0 sheet
`
`NN
`
`1141
`
`nO
`
`2.
`
`CO
`
`a
`
`tT
`O
`a .
`
`A
`
`iv
`D
`3
`
`.y4
`
`3
`fa
`
`CD
`
`a helix
`
`bond
`
`right-handed
`helix
`
`3.5
`
`sugar phosphate
`backbone
`
`1.6 nm
`
`5'
`
`/
`
`V
`
`A
`
`base •
`
`1 hei.cal !urn -- 3-4 nm •
`
`3.6
`
`hydrogen
`bond
`
`

`

`60
`
`5. Charge-Transfer Complexes
`
`3. Receptors
`
`When a molecule (or group) that is a good electron donor comes into contact
`with a molecule (or group) that is a good electron acceptor, the donor may
`transfer some of its charge to the acceptor. This forms a charge-transfer
`complex, which, in effect, is a molecular dipole—dipole interaction. The po-
`tential energy of this interaction is proportional to the difference between the
`ionization potential of the donor and the electron affinity of the acceptor.
`Electron donors contain 77-electrons, for example, alkenes, alkynes, and
`aromatic moieties with electron-donating substituents, or groups that have a
`pair of nonbonded electrons, such as oxygen, nitrogen, and sulfur moieties.
`Acceptor groups contain electron-deficient 7F orbitals, for example, alkenes,
`alkynes, and aromatic moieties having electron-withdrawing substituents, or
`weakly acidic protons. There are groups on receptors that can act as electron
`donors, such as the aromatic ring of tyrosine or the carboxylate group of
`aspartate, as electron acceptors, such as cysteine, and electron donors and
`acceptors, such as histidine, tryptophan, and asparagine.
`Charge-transfer interactions are believed to provide the energy for interca-
`lation of certain planar aromatic antimalarial drugs, such as chloroquine (3.7),
`into parasitic DNA (see Chapter 6). The fungicide, chlorothalonil, is shown in
`Fig. 3.4 as a hypothetical example for a charge-transfer interaction with a
`tyrosine.
`
`Cl
`
`NH-CH(CH2)3N(C2115)2
`CH3
`
`3.7
`The AG° for charge-transfer interactions also can range from —1 to —7
`kcal/mol.
`
`6. Hydrophobic Interactions
`In the presence of a nonpolar molecule or region of a molecule, the surround-
`ing water molecules orient themselves and, therefore, are in a higher energy
`
`CN
`
`CI
`
`CI
`
`CI
`
`CN
`
`OH
`
`Figure 3.4. Example of a charge-transfer interaction. The wavy line is the receptor surface.
`
`

`

`III. Drug—Receptor Interactions
`
`,6\7
`C761/ 4 16,7
`VA767
`7676''c7/447A v,S AAv
`'S,ii
`e`e''>'34 \ 7 7 e.667
`ov,, /
`,, ,,e1v4,
`\7
`.6, AV:n:7
`V:A ////A7A7
`CZAG2FA
`
`i,
`
`/
`
`,
`
`/
`
`61
`
`G. p
`
`:/;43/:///:N/I'''7°//;:/:,:‘,1: ;
`/.-
`v,727:1:6,1:Y. 4,
`/ADA767,4C7
`4'7,6,yAV ,q zs <
`zs,
`D
`D .A
`
`,
`
`b.
`
`Figure 3.5. Formation of hydrophobic interactions. (Reprinted with permission of John
`Wiley & Sons, Inc. from Korolkovas, A. 1970. "Essentials of Molecular Pharmacology," p.
`172. Wiley, New York. Copyright © 1970. John Wiley & Sons, Inc. and by permission of
`Kopple, K. D. 1966."Peptides and Amino Acids." Addison-Wesley, Reading, Massachusetts.)
`
`state than when only other water molecules are around. When two nonpolar
`groups, such as a lipophilic group on a drug and a nonpolar receptor group,
`each surrounded by ordered water molecules, approach each other, these
`water molecules become disordered in an attempt to associate with each
`other. This increase in entropy, therefore, results in a decrease in the free
`energy that stabilizes the drug—receptor complex. This stabilization is known
`as a hydrophobic interaction (see Fig. 3.5). Consequently, this is not an at-
`tractive force of two nonpolar groups "dissolving" in one another but, rather,
`is the decreased free energy of the nonpolar group because of the increased
`entropy of the surrounding water molecules. Jencks9 has suggested that hy-
`drophobic forces may be the most important single factor responsible for
`noncovalent intermolecular interactions in aqueous solution. Hildebrand,19 on
`the other hand, is convinced that hydrophobic effects do not exist. Every
`methylene-methylene interaction (which actually may be a van der Waals
`interaction; see Section III,B,7) liberates 0.7 kcal/mol of free energy. In Fig.
`3.6 the topical anesthetic butamben is depicted in a hypothetical hydrophobic
`interaction with an isoleucine group.
`
`7. Van der Waals or London Dispersion Forces
`
`Atoms in nonpolar molecules may have a temporary nonsymmetrical distribu-
`tion of electron density which results in the generation of a temporary dipole.
`
`NH2
`
`\
`
`/
`
`Figure 3.6. Example of hydrophobic interactions. The wavy line represents the receptor
`surface.
`
`

`

`62
`
`3. Receptors
`
`As atoms from different molecules (such as a drug and a receptor) approach
`each other, the temporary dipoles of one molecule induce opposite dipoles in
`the approaching molecule. Consequently, an intermolecular attraction,
`known as van der Waals forces, results. These weak universal forces only
`become significant when there is a close surface contact of the atoms; how-
`ever, when there is molecular complementarity, numerous atomic interac-
`tions (each contributing about —0.5 kcal/mol to the AG') result, which can add
`up to a significant overall drug—receptor binding component.
`
`8. Conclusion
`
`Since noncovalent interactions are generally weak, cooperativity by several
`types of interactions is critical. To a first approximation, enthalpy terms will
`be additive. Once the first interaction has taken place, translational entropy is
`lost. This results in a much lower entropy loss in the formation of the second
`interaction. The effect of this cooperativity is that several rather weak interac-
`tions may combine to produce a strong interaction. Since several different
`types of interactions are involved, selectivity in drug—receptor interactions
`can result. In Fig. 3.7 the local anesthetic dibucaine is used as an example to
`show the variety of interactions that are possible.
`
`C. Ionization
`
`At physiological pH (pH 7.4), even mildly acidic groups, such as carboxylic
`acid groups, will be essentially completely in the carboxylate anionic form;
`phenolic hydroxyl groups may be partially ionized. Likewise, basic groups,
`such as amines, will be partially or completely protonated to give the cationic
`form. The ionization state of a drug will have a profound effect not only on its
`drug—receptor interaction, but also on its partition coefficient (log P; see
`Section II,E,2,b of Chapter 2).
`
`/ydrophobic
`
`hydrogen bond
`
`charge transfer
`
`1-1""
`
`\
`
`ionic or ion-dipole
`
`:N
`\
`
`CH3CH2CH2CH2O
`
`H
`I+
`N
`NCH2CH2—IrCH2CH3
`CH2CH3
`
`hydrophobic
`
`dipole-dipole
`
`hydrophobic
`
`Figure 3.7. Examples of potential multiple drug—receptor interactions. The van der Waals
`interactions are excluded.
`
`

`

`III. Drug—Receptor Interactions
`
`63
`
`The importance of ionization was recognized in 1924 when Steam and
`Stearn11 suggested that the antibacterial activity of stabilized triphenyl-
`methane cationic dyes was related to an interaction of the cation with some
`anionic group in the bacterium. Increasing the pH of the medium also in-
`creased the antibacterial effect, presumably by increasing the ionization of the
`receptors in the bacterium, Albert and co-workers12 made the first rigorous
`proof that a correlation between ionization and biological activity existed. A
`series of 101 aminoacridines, including the antibacterial drug, 9-aminoacridine
`or aminacrine (3.8), all having a variety of pKa values, was tested against 22
`species of bacteria. A direct correlation was observed between ionization
`(formation of the cation) of the aminoacridines and antibacterial activity.
`However, at lower pH values, protons can compete with these cations for the
`receptor, and antibacterial activity is diminished. When this was realized,
`Albert13 notes, the Australian Army during World War II was advised to
`pretreat wounds with sodium bicarbonate to neutralize any acidity prior to
`treatment with aminacrine. This, apparently, was quite effective in increasing
`the potency of the drug. The mechanism of action of aminoacridines is dis-
`cussed in Chapter 6.
`
`NH2
`
`3.8
`
`The great majority of alkaloids which act as neuroleptics, local anesthetics,
`and barbiturates have pKa values between 6 and 8; consequently both neutral
`and cationic forms are present at physiological pH.13 This may allow them to
`penetrate membranes in the neutral form and exert their biological action in
`the ionic form. Antihistamines and antidepressants tend to have pKa values of
`about 9. The uricosuric (increases urinary excretion of uric acid) drug phenyl-
`butazone [3.9, R = (CH2)3CH31 has a pKa of 4.5 and is active as the anion (the
`OH proton is acidic). However, since the pH of urine is 4.8 or higher, subopti-
`mal concentrations of the anion were found in the urinary system. Sulfinpyra-
`zone (3.9, R = CH2CH2SOPh) has a lower pKa of 2.8 and is about 20 times
`more potent than phenylbutazone; the anionic form blocks reabsorption of
`uric acid by renal tubule cells.14
`
`HO
`
`Ph
`N.
`
`3.9
`
`-Ph
`
`O
`
`

`

`64
`
`3. Receptors
`
`The antimalarial drug pyrimethamine (3.10) has a pKa of 7.2 and is best
`absorbed from solutions of sufficient alkalinity that it has a high proportion of
`molecules in the neutral form (to cross membranes). Its mode of action, the
`inhibition of the parasitic enzyme dihydrofolate reductase, however, requires
`that it be in the protonated cationic form.
`
`CI
`
`NH2
`
`C2H5
`
`N
`
`NH2
`
`3.10
`
`Similarly, there are drugs such as the anti-inflammatory agent indomethacin
`(2.18) and the antibacterial agent sulfamethoxazole (3.11) whose pharmaco-
`kinetics (migration to site of action) depend on their nonionized form, but
`whose pharmacodynamics (interaction with the receptor) depend on the an-
`ionic form (carboxylate and sulfonamido ions, respectively). In a cell-free
`system the antibacterial activity of 3.11 and other sulfonamides was directly
`proportional to the degree of ionization, but in intact cells, where the drug
`must cross a membrane to get to the site of action, the antibacterial activity
`also was dependent on lipophilicity (the neutral form).15
`
`NH2
`
`/ \
`
`CH3
`
`3.11
`Up to this point only the ionization of the drug has been considered. As
`indicated in Section III,B,2, there are a variety of acidic and basic groups on
`receptors. Anionic groups in DNA include phosphoric acid groups (pKa 1.5 or
`6.5) and purines and pyrimidines (pKa —9); anionic groups in proteins are
`carboxylic acids (aspartic and glutamic acids; pKa 3.5-5), phenols (tyrosine;
`pKa 9.5-11), sulfhydryls (cysteine; pKa 8.5), and hydroxyls (serine and
`threonine; pKa —13.5). Cationic groups in DNA include amines (adenine and
`cytidine; pKa 3.5-4) and in proteins include imidazole (histidine, pKa 6.5-7),
`amino (lysine, pKa —10), and guanidino (arginine, pKa —13) groups. There-
`fore, the structure and function of a receptor can be strongly dependent on the
`pH of the medium, especially if an in vitro assay is being used. The pKa values
`of various groups embedded in a receptor, however, can be quite variable,
`and will depend on the microenvironment. If a carboxyl group is in a nonpolar
`region, its pKa will be raised because the anionic form is destabilized. Gluta-
`mate-35 in lysozyme and the lysozyme—glycolchitin complex has a pKa of 6.5
`and 8.2, respectively.16 If the carboxylate forms a salt bridge, it will be stabi-
`
`

`

`III. Drug—Receptor Interactions
`
`65
`
`lized and its pKa will be lower. Likewise, an amino group buried in a nonpolar
`microenvironment will have a lower pKa because protonation will be disfa-
`vored; the s-amino group of the active site lysine residue in acetoacetate
`decarboxylase has a pKa of 5.9.17 If the ammonium group forms a salt bridge,
`it will be stabilized, deprotonation will be inhibited, and the pKa will be raised.
`Now that the importance of drug—receptor interactions has been empha-
`sized, we turn our attention to the principal method for the determination of
`these interactions.
`
`D. Determination of Drug-Receptor Interactions
`
`Hormones and neurotransmitters are important natural compounds that are
`responsible for the regulation of a myriad of physiological functions. These
`molecules interact with a specific receptor in a tissue and elicit a specific
`characteristic response. For example, the activation of a muscle by the cen-
`tral nervous system is mediated by release of the neurotransmitter acetylcho-
`line (ACh; the molecule in Figs. 3.1 and 3.2). If a plot is made of the logarithm
`of the concentration of the acetylcholine added to a muscle tissue preparation
`versus the percentage of total muscle contraction, the graph shown in Fig. 3.8
`may result. This is known as a dose—response or concentration—response
`curve. The low concentration part of the curve results from too few neuro-
`transmitter molecules available for collision with the receptor. As the concen-
`tration increases, it reaches a point where a linear relationship is observed
`between the logarithm of the neurotransmitter concentration and the biologi-
`cal response. As most of the receptors become occupied, the probability of a
`drug and receptor molecule interacting diminishes, and the curve deviates
`
`i
`i
`I
`I
`I
`I
`8 7 6 5 4
`10 9
`-log [ACh] M
`
`100%-
`
`50 -
`
`0-
`
`% Muscle Contraction
`
`Figure 3.8. Effect of increasing the concentration of a neurotransmitter on muscle contrac-
`tion.
`
`

`

`3. Receptors
`
`100%
`
`50
`
`0
`
`66
`
`% Muscle Contraction
`
`10 9 8 7 6 5
`
`4
`
`-log [X] M
`
`Figure 3.9. Dose—response curve for an agonist.
`
`from linearity (the high concentration end). Dose—response curves are a
`means of measuring drug—receptor interactions and are the standard method
`for comparing the potencies of various compounds that interact with a partic-
`ular receptor. Any measure of a response can be plotted on the ordinate, such
`as LD50 , ED50 , or percentage of a physiological effect.
`If another compound (X) is added in increasing amounts to the same tissue
`preparation and the curve shown in Fig. 3.9 results, the compound, which
`produces the same maximal response as the neurotransmitter, is called an
`agonist. A second compound (Y) added to the tissue preparation shows no
`response at all (Fig. 3.10A); however, if it is added to the neurotransmitter,
`the effect of the neurotransmitter is blocked until a higher concentration of the
`neurotransmitter is added (Fig. 3.10B). Compound Y is called a competitive
`antagonist. There are two general types of antagonists, competitive antago-
`nists and noncompetitive antagonists. The former, which is the larger cate-
`gory, is one in which the degree of antagonism is dependent on the relative
`concentrations of the agonist and the antagonist; both bind to the same site on
`the receptor, or, at least, the antagonist directly interferes with the binding of
`the agonist. The degree of blocking of a noncompetitive antagonist (Y') is
`independent of the amount of agonist present; two different binding sites may
`be involved (Fig. 3.10C). Only competitive antagonists will be discussed fur-
`ther in this text.
`If a compound Z is added to the tissue preparation and some response is
`elicited, but not a full response, regardless of how high the concentration of Z
`used, then Z is called a partial agonist (see Fig. 3.11A). A partial

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket