throbber
+
`
`Chem. Rev. 1996, 96, 3147- 3176
`
`Bioisosterism: A Rational Approach in Drug Design
`
`George A. Patani and Edmond J. LaVoie*
`
`+
`
`3147
`
`Department of Pharmaceutical Chemistry, College of Pharmacy, Rutgers, The State University of New Jersey, Piscataway, New Jersey 08855-0789
`
`Received May 15, 1996(Revised Manuscript Received July 25, 1996)
`
`3147
`3149
`3149
`3149
`3150
`
`3154
`
`Contents
`I.
`Introduction
`II. Classical Bioisosteres
`A. Monovalent Atoms or Groups
`1. Fluorine vs Hydrogen Replacements
`2.
`Interchange of Hydroxyl and Amino
`Groups
`Interchange of Hydroxyl and Thiol Groups 3151
`3.
`4. Fluorine and Hydroxyl, Amino, or Methyl
`3152
`Groups as Replacements for Hydrogen
`(Grimm’s Hydride Displacement Law)
`5. Monovalent Substitutions Involving
`Chloro, Bromo, Thiol, and Hydroxyl
`Groups (Erlenmeyer’s Broadened
`Classification of Grimm’s Displacement
`Law)
`B. Divalent Isosteres
`1. Divalent Replacements Involving Double
`Bonds
`2. Divalent Replacements Involving Two
`Single Bonds
`C. Trivalent Atoms or Groups
`D. Tetrasubstituted Atoms
`E. Ring Equivalents
`1. Divalent Ring Equivalents
`2. Trivalent Ring Equivalents
`III. Nonclassical Bioisosteres
`A. Cyclic vs Noncyclic Nonclassical Bioisosteric
`Replacements
`B. Nonclassical Bioisosteric Replacements of
`Functional Groups
`1. Hydroxyl Group Bioisosteres
`2. Carbonyl Group Bioisosteres
`3. Carboxylate Group Bioisosteres
`4. Amide Group Bioisosteres
`5. Thiourea Bioisosteres
`6. Halogen Bioisosteres
`IV. Conclusion
`V. Acknowledgments
`VI. References
`
`3155
`3155
`
`3155
`
`3156
`3157
`3158
`3158
`3159
`3160
`3160
`
`3165
`
`3165
`3166
`3168
`3170
`3171
`3172
`3172
`3172
`3172
`
`George Patani graduated with a B.Pharm. in 1992 from the College of
`Pharmaceutical Sciences, Mangalore University at Manipal, India.
`In 1996,
`he received his M.S. in Pharmaceutical Science at Rutgers University
`under the direction of Professor Edmond J. LaVoie. He is presently
`pursuing graduate studies in pharmaceutics. His current research interests
`are focused on drug design and controlled drug delivery.
`
`Edmond J. LaVoie received his B.S. in Chemistry from Fordham University
`in 1971 and his Ph.D. in Medicinal Chemistry from S.U.N.Y. at Buffalo
`under the direction of Dr. Wayne K. Anderson. After postdoctoral study
`with Dr. S. Morris Kupchan at the University of Virginia, he joined the
`American Health Foundation in Valhalla, NY.
`In 1988, he was appointed
`Professor of Medicinal Chemistry in the College of Pharmacy at Rutgers
`University. His current research interests are in the design and synthesis
`of cancer chemotherapeutics and in the elucidation of mechanism(s) of
`carcinogenesis.
`for the rational modification of lead compounds into
`safer and more clinically effective agents. The con-
`cept of bioisosterism is often considered to be qualita-
`tive and intuitive.1
`The prevalence of the use of bioisosteric replace-
`ments in drug design need not be emphasized. This
`topic has been reviewed in previous years.2-5 The
`objective of this review is to provide an overview of
`bioisosteres that incorporates sufficient detail to
`enable the reader to understand the concepts being
`delineated. While a few popular examples of the
`successful use of bioisosteres have been included, the
`
`I. Introduction
`Years of cumulative research can result in the
`development of a clinically useful drug, providing
`either a cure for a particular disease or symptomatic
`relief from a physiological disorder. A lead compound
`with a desired pharmacological activity may have
`associated with it undesirable side effects, charac-
`teristics that limit its bioavailability, or structural
`features which adversely influence its metabolism
`and excretion from the body. Bioisosterism repre-
`sents one approach used by the medicinal chemist
`
`S0009-2665(95)00066-5 CCC: $25 00
`
`© 1996 American Chemical Society
`
`LUPIN EX 1037
`
`Page 1 of 30
`
`

`
`+
`
`3148 Chemical Reviews, 1996, Vol. 96, No. 8
`
`present review is focused primarily upon specific
`examples from current literature. The emphasis in
`this review was to outline bioisosteric replacements
`which have been used to advance drug development.
`No attempt was made to be exhaustive or to illustrate
`all of the specific analogues represented within a
`single study.
`The ability of a group of bioisosteres to elicit similar
`biological activity has been attributed to common
`physicochemical properties. In this review an at-
`tempt has been made to quantitate,
`in specific
`instances, physicochemical effects such as electro-
`negativity, steric size, and lipophilicity and to cor-
`relate these values to the observed biological activity.
`Thus, an additional objective of this review was to
`demonstrate the opportunities that one has in em-
`ploying bioisosteres to gain more specific insight into
`the quantitative structure-activity relationships
`(QSAR) associated with a specific class of drugs.
`While in some instances such associations were
`detailed by the authors of these literature examples,
`others were developed on the basis of evident cor-
`relations. To further explain and rationalize the
`biological activity observed with nonclassical bioiso-
`steric groups, the observed biological activity has also
`been correlated with some substituent constants
`commonly employed in QSAR studies. These obser-
`vations are consistent with the fact that bioisosteric
`replacements often provide the foundation for the
`development of QSAR in drug design.4,6 Recent
`advances in molecular biology, such as cloning of the
`various receptor subtypes, have enabled a clearer
`definition of the pharmacophoric sites. Bioisosteric
`replacements of functional groups based on this
`understanding of the pharmacophore and the phys-
`icochemical properties of the bioisosteres have en-
`hanced the potential for the successful development
`of new clinical agents.
`The bioisosteric rationale for the modification of
`lead compounds is traced back to the observation by
`Langmuir in 1919 regarding the similarities of vari-
`ous physicochemical properties of atoms, groups,
`radicals, and molecules.7 Langmuir compared the
`physical properties of various molecules such as N2
`- and NCO- and found
`and CO, N2O and CO2, and N3
`them to be similar. On the basis of these similarities
`he identified 21 groups of isosteres. Some of these
`groups are listed in Table 1. He further deduced from
`the octet theory that the number and arrangement
`of electrons in these molecules are the same. Thus,
`isosteres were initially defined as those compounds
`or groups of atoms that have the same number and
`
`Table 1. Groups of Isosteres as Identified by
`Langmuir
`groups
`1
`2
`3
`4
`V
`8
`9
`10
`V
`20
`21
`
`isosteres
`H-, He, Li+
`O2-, F-, Ne, Na+, Mg2+, Al3+
`S2-, Cl-, Ar, K+, Ca2+
`Cu2-, Zn2+
`V
`N2, CO, CN-
`+
`CH4, NH4
`-, CNO-
`CO2, N2O, N3
`V
`2-MnO4-, CrO4
`
`
`3-SeO42-, AsO4
`
`
`
`+
`
`Patani and LaVoie
`
`arrangement of electrons. He further defined other
`relationships in a similar manner. Argon was viewed
`as an isostere of K+ ion and methane as an isostere
`+ ion. He deduced, therefore, that K+ ions and
`of NH4
`+ ions must be similar because argon and meth-
`NH4
`ane are very similar in physical properties. The
`biological similarity of molecules such as CO2 and
`N2O was later coincidentally acknowledged as both
`compounds were capable of acting as reversible
`anesthetics to the slime mold Physarum polyceph-
`alum.8
`A further extension to this concept of isosteres
`came about in 1925 with Grimm’s Hydride Displace-
`ment Law.9,10 This law states: “Atoms anywhere up
`to four places in the periodic system before an inert
`gas change their properties by uniting with one to
`four hydrogen atoms, in such a manner that the
`resulting combinations behave like pseudoatoms,
`which are similar to elements in the groups one to
`four places respectively, to their right.” Each vertical
`column as illustrated in Table 2, according to Grimm,
`would represent a group of isosteres.
`Table 2. Grimm’s Hydride Displacement Law
`C
`N
`O
`F
`Ne
`CH
`NH
`OH
`FH
`CH2
`NH2
`OH2
`CH3
`NH3
`CH4
`
`Na
`-
`+
`FH2
`+
`OH3
`+
`NH4
`
`4
`N+
`P+
`S+
`As+
`Sb+
`
`Erlenmeyer11 further broadened Grimm’s clas-
`sification and redefined isosteres as atoms, ions, and
`molecules in which the peripheral layers of electrons
`can be considered identical (Table 3).
`Table 3. Isosteres Based on the Number of
`Peripheral Electrons
`no. of peripheral electrons
`8
`5
`6
`7
`ClH
`P
`S
`Cl
`BrH
`As
`Se
`Br
`IH
`Sb
`Te
`I
`SH2
`PH
`SH
`PH2
`PH3
`The widespread application of the concept of iso-
`sterism to modify biological activity has given rise
`to the term bioisosterism. As initially defined by
`Friedman,2 bioisosteres were to include all atoms and
`molecules which fit the broadest definition for iso-
`steres and have a similar type of biological activity,
`which may even be antagonistic. More recently this
`definition has been broadened by Burger as “Com-
`pounds or groups that possess near-equal molecular
`shapes and volumes, approximately the same distri-
`bution of electrons, and which exhibit similar physi-
`cal properties...”.5 The critical component for bio-
`isosterism is that bioisosteres affect the same phar-
`macological target as agonists or antagonists and,
`thereby, have biological properties which are related
`to each other.
`Bioisosteres have been classified as either classical
`or nonclassical.12 Grimm’s Hydride Displacement
`Law and Erlenmeyer’s definition of isosteres outline
`a series of replacements which have been referred
`to as classical bioisosteres. Classical bioisosteres
`have been traditionally divided into several distinct
`categories:
`(A) monovalent atoms or groups; (B)
`
`Page 2 of 30
`
`

`
`+
`
`+
`
`Bioisosterism: A Rational Approach in Drug Design
`
`divalent atoms or groups; (C) trivalent atoms or
`groups; (D) tetrasubstituted atoms; and (E) ring
`equivalents.
`Nonclassical isosteres do not obey the steric and
`electronic definition of classical isosteres. A second
`notable characteristic of nonclassical bioisosteres is
`that they do not have the same number of atoms as
`the substituent or moiety for which they are used as
`a replacement. Nonclassical bioisosteres can be
`further divided into groups:
`(A) rings vs noncyclic
`structures; and (B) exchangeable groups.
`This approach to classifying bioisosteres will be
`used to review literature examples of those bioiso-
`steric replacements that have provided useful infor-
`mation on the structure-activity relationships as-
`sociated with various pharmacologically active com-
`pounds.
`
`II. Classical Bioisosteres
`
`A. Monovalent Atoms or Groups
`Similarities in certain physicochemical properties
`have enabled investigators to successfully exploit
`several monovalent bioisosteres. These can be di-
`vided into the following groups:
`(1) fluorine vs
`hydrogen replacements; (2) amino-hydroxyl inter-
`changes; (3) thiol-hydroxyl interchanges; (4) fluorine,
`hydroxyl, amino, and methyl group interchanges
`(Grimm’s Hydride Displacement Law); (5) chloro,
`bromo, thiol, and hydroxyl group interchanges (Er-
`lenmeyer’s Broadened Classification of Grimm’s Dis-
`placement Law).
`
`1. Fluorine vs Hydrogen Replacements
`The substitution of hydrogen by fluorine is one of
`the more commonly employed monovalent isosteric
`replacements. Steric parameters for hydrogen and
`fluorine are similar, their van der Waal’s radii being
`1.2 and 1.35 Å, respectively.13 Thus, the difference
`in the electronic effects (fluorine being the most
`electronegative element in the periodic table) is often
`the basis for the major differences in the pharmaco-
`logical properties of agents where fluorine has been
`substituted for hydrogen. Due to its electronegativ-
`ity, fluorine exerts strong field and inductive effects
`on the adjacent carbon atom. Fluorine substitution,
`in general, exerts a diminished electron-withdrawing
`effect at distal sites. However, fluorine can donate
`a lone pair of electrons by resonance. This is com-
`monly referred to as its mesomeric effect. The
`opposing resonance and field effects can nearly
`cancel. The pharmacological differences can be at-
`tributed to the influence of the electron-withdrawing
`effect that the fluorine substitution causes on inter-
`action with either a biological receptor or enzyme,
`as well as its effect on the metabolic fate of the drug.
`The antineoplastic agent 5-fluorouracil (5-FU) rep-
`resents a classical example of how fluorine substitu-
`tion of a normal enzyme substrate can result in a
`derivative which can alter select enzymatic processes.
`In this instance, 5-FU is biochemically transformed
`in vivo into 5-fluoro-2¢
`-deoxyuridylic acid. Its close
`similarity to uracil allows this fluoro derivative to be
`a successful mimetic. This biochemically altered
`
`Chemical Reviews, 1996, Vol. 96, No. 8 3149
`form of 5-FU, 5-fluoro-2¢
`-deoxyuridylic acid, is ulti-
`mately responsible for the inhibition of thymidylate
`synthase, an enzyme involved in the conversion of
`uridylic acid to thymidylic acid and critical for DNA
`synthesis (Figure 1). The increased reactivity of
`5-fluoro-2¢
`-deoxyuridylic acid relative to 2¢
`-deoxy-
`uridylic acid is due to the inductive effect of fluorine
`which results in its covalent binding to thymidylate
`synthase.
`
`Figure 1.
`The application of the monovalent substitution of
`a fluorine atom for a hydrogen atom can also be seen
`in a more recent study with naphthyl-fused diaze-
`pines, which were employed as agonistic probes of
`the pharmacophore of benzodiazepine receptors.14
`Replacement of the hydrogen with fluorine at the
`ortho position of the pendent phenyl group of either
`naphthyl-fused diazepines, as illustrated in Figure
`2, resulted in enhanced affinity and efficacy for both
`naphthyl isomers (Table 4). This greater receptor
`binding affinity could again be attributed to the
`inductive effect of the fluorine atom facilitating a
`stronger interaction with the receptor.
`
`Figure 2.
`
`Table 4. Benzodiazepine Receptor Binding Affinity
`for Naphthyl-Fused Diazepines
`IC50 (nM)a
`compound
`X
`1a
`1000
`H
`1b
`260
`F
`2a
`1000
`H
`2b
`55
`F
`a In vitro potency of the compound to displace [3H]fluni-
`trazepam from the benzodiazepine receptor.
`
`Another good illustration of this monovalent bioi-
`sosteric replacement is observed in a recent series of
`anti-inflammatory corticosteroid analogues (3, Figure
`3).15
`In this study, the topical anti-inflammatory
`activity of two pairs of structurally similar corticos-
`teroids were compared. Their relative anti-inflam-
`matory activity was normalized to fluocinolone ace-
`tonide, which was assigned a potency of 100. Table
`
`Page 3 of 30
`
`

`
`+
`
`3150 Chemical Reviews, 1996, Vol. 96, No. 8
`
`+
`
`Patani and LaVoie
`
`Figure 3.
`
`Table 5. Biological Activities of Halomethyl
`Androstane-17(cid:226)-carbothionates
`
`topical
`anti-inflammatory
`activitya
`Z
`Y
`X
`compound
`3a
`42
`dCH2
`F
`H
`3b
`dCH2
`F
`F
`108
`3c
`(cid:226)-CH3
`H
`H
`27
`3d
`(cid:226)-CH3
`F
`H
`41
`a Topical anti-inflammatory activity was measured in mice
`by modifications of the croton oil ear assay.16 Fluocinolone
`acetonide served as a positive control and is assigned a relative
`potency index of 100.
`
`5 shows that, in the case of the pair of compounds
`possessing a 16-methylene substituent, the presence
`of an additional fluorine atom at the 6R position
`results in a derivative with greater activity than 3a
`or fluocinolone acetonide. With the pair of corticos-
`teroids with a 16-methyl substituent (Z ) CH3),
`replacement of hydrogen with fluorine at the 9R
`position, 3d, also increased anti-inflammatory activ-
`ity relative to 3c.
`Thus, the ability of fluorine to replace hydrogen is
`an effective method of exploring the affinity of an
`agent to the target site (receptor or enzyme) by virtue
`of its greater electronegativity while other param-
`eters such as steric size and lipophilicity17 are
`maintained.
`
`2. Interchange of Hydroxyl and Amino Groups
`The monovalent interchange of amino and hydroxyl
`groups is well known and has been successfully
`employed in the development of various pharmaco-
`logical agents. The similar steric size (Table 7),
`spatial arrangement, and the ability of these func-
`tional groups to act as either hydrogen bond acceptors
`or donors is likely responsible for their successful use
`as bioisosteres.
`Several medicinal agents under investigation as
`potential clinical agents carry heteroaromatic moi-
`eties. Many of these heteroaromatic compounds are
`capable of tautomerization. The prototropic tautom-
`erism of heteroaromatic compounds includes all
`agents wherein a mobile proton can move from one
`site to another within the heteroaromatic molecule.
`Figure 4 illustrates one of the more common types
`of tautomerization involving the movement of a
`proton between a cyclic nitrogen atom and a sub-
`stituent on the neighboring carbon atom within the
`ring. Tautomerism in heterocyclic molecules has
`been extensively studied.18 In the presence of electron-
`donating atoms such as nitrogen in heterocyclic
`systems, it is known that there will be substantial
`tautomerization where a neighboring CsOH will
`tautomerize to CdO.19 In the case of a neighboring
`
`Figure 4.
`carbon containing CsNH2 (7, Figure 4), the preferred
`tautomer is the CsNH2 form.
`Perhaps the best known example of classical iso-
`steric substitution of an amino for a hydroxyl group
`is illustrated by aminopterin (8b) wherein the hy-
`droxyl substituent of folic acid (8a) has been substi-
`tuted by an amino group (Figure 5). As previously
`noted, this represents a monovalent bioisosteric
`substitution at a carbon atom adjacent to a hetero-
`cyclic nitrogen atom. Thus, this bioisosteric replace-
`ment has the capability of mimicking even the
`tautomeric forms of folic acid. The similarity as well
`as the capability of the amino group to hydrogen bond
`to the enzyme are two important factors that facili-
`tate the binding of aminopterin to the enzyme dihy-
`drofolate reductase.
`
`Figure 5.
`
`Interchange of an amino group with a hydroxyl
`moiety in the case of 6,9-disubstituted purines (Table
`6) has been shown to result in the development of
`agents with similar benzodiazepine receptor binding
`activity.20 This example further substantiates the
`ability of the amino group to mimic the hydroxyl
`group at the receptor site. In this study a series of
`6,9-disubstituted purines were tested for their ability
`to bind to the benzodiazepine receptor in rat brain
`tissue. The relative activity of the 9-(3-aminophenyl)-
`methyl derivative (9a) was compared to the 9-(3-
`hydroxyphenyl)methyl analogue (9b) (Figure 6). In
`contrast to aminopterin where a dramatic difference
`in binding affinity was observed relative to the
`normal substrate, these bioisosteric 6,9-disubstituted
`
`Figure 6.
`
`Table 6. Benzodiazepine Receptor Binding Activity
`of Substituted 6-(Dimethylamino)-9-benzyl-9H-purines
`compound
`R
`IC50 ((cid:237)M)a
`9a
`NH2
`0.9
`9b
`OH
`1.2
`a Concentration of compound that decreased specific binding
`of 1.5 nM [3H]diazepam to rat brain receptors by 50%.
`
`Page 4 of 30
`
`

`
`+
`
`+
`
`Bioisosterism: A Rational Approach in Drug Design
`
`Chemical Reviews, 1996, Vol. 96, No. 8 3151
`
`purines exhibited similar activity with regard to their
`affinity for the benzodiazepine receptor.
`In this
`example of bioisosteric replacement, pharmacological
`activity was retained. It is important to note that
`retention of biological activity based on in vitro data
`can be critical in those instances where differences
`between bioisosteric analogues exist with regard to
`in vivo parameters which may include absorption,
`distribution, metabolism, or elimination. While one
`may only observe retention of activity associated with
`interaction of drug with the pharmacophore, bioiso-
`steres may differ dramatically in their in vivo ef-
`ficacy. Additional examples of this bioisosteric re-
`placement will be discussed in the next section on
`monovalent replacement of hydroxyl and thiol groups.
`3. Interchange of Hydroxyl and Thiol Groups
`The interchange of thiol for hydroxyl can be con-
`sidered as an extension of the amino-hydroxyl
`replacement and has been used extensively in me-
`dicinal chemistry. This replacement is based on the
`ability of both these functional groups to be hydrogen
`bond acceptors or donors. A classical illustration of
`this replacement being guanine (10a) and 6-thiogua-
`nine (10b, Figure 7).21
`
`Figure 7.
`As discussed in the previous section, when part of
`a heteroaromatic ring, these functional groups can
`exist in different tautomeric forms. Figure 8 il-
`lustrates the most common example wherein a mobile
`proton on a nitrogen atom in the aromatic ring can
`be transferred to the heteroatom attached to the
`adjacent carbon resulting in the different tautomers.
`
`Figure 9.
`
`Table 7. Calcium Channel Blocking Activity of
`1,4-Dihydropyrimidines
`
`van der Waal’s
`IC50 (nM)a
`radius24 (Å)
`X
`compound
`15a
`140
`1.40
`dO
`15b
`160
`1.50
`dNH
`15c
`17
`1.85
`dS
`a Concentration that produced 50% inhibition and deter-
`mined for the vasorelaxant activity with potassium-depolarized
`rabbit thoracic aorta.
`
`analogues with similar potency. However, substitu-
`tion with the thiol resulted in enhanced potency
`(Table 7). This could be explained by the fact that
`the size of the substituents, described here as the van
`der Waal’s radii, and the ability to hydrogen bond
`were the important factors influencing retention of
`activity. Therefore, replacement with the amino
`group, which has a similar size, resulted in similar
`potency. However, replacement with the sterically
`optimal thiol resulted in an analogue which was an
`order of magnitude more potent.
`The use of this replacement in the design of novel
`anti-inflammatory agents substantiates its utility as
`a monovalent bioisostere. Long term use of nonste-
`roidal anti-inflammatory drugs (NSAIDs) for the
`treatment of rheumatoid arthritis and other inflam-
`matory diseases has been associated with side effects
`such as gastrointestinal ulceration, bleeding, and
`nephrotoxicity.25,26 With a view to designing new
`drugs with an improved safety profile, certain thia-
`zoles (16, Figure 10 and Table 8) that are dual
`
`Figure 8.
`In the case of 6-thioguanine, the ability of this
`bioisosteric analogue to be viewed as a substrate by
`the salvage pathway associated with purine biosyn-
`thesis, allows for its transformation into 6-thiogua-
`nylic acid by hypoxanthine-guanine phosphoribosyl-
`transferase (HGPRT). However, the significance of
`this “fraudulent” nucleic acid with respect to its
`lethality to neoplasms is uncertain.22 It is as this
`phosphoriboside that either the de novo synthesis of
`nucleic acids is inhibited or incorporation into deoxy-
`ribonucleic acid occurs.
`In an attempt to enhance the calcium channel
`blocking capacity of certain dihydropyrimidine agents,
`a number of isosteric analogues with the general
`structure 15 (Figure 9) were synthesized.23 Substitu-
`tion of the hydroxyl with an amino resulted in
`
`Figure 10.
`Table 8. Anti-inflammatory Activity of Benzylidene
`Derivatives in Intact Rat Basophilic Leukemia
`(RBL-1) Cells
`
`IC50 ((cid:237)M)a
`5-LO
`CO
`electronegativity29
`Z
`compound
`16a
`1.4
`0.35
`3.51
`OH
`16b
`NH2
`0.77
`0.39
`2.61
`16c
`0.38
`0.012
`2.32
`SH
`a Concentration of the test compound causing 50% inhibition
`of 5-LO or CO formation.
`
`inhibitors of both cyclooxygenase (CO) and 5-lipoxy-
`genase (5-LO) are being studied as potential anti-
`inflammatory agents.27 The beneficial effects of
`NSAIDs have been attributed to the inhibition of the
`enzyme cyclooxygenase, thereby preventing produc-
`tion of pro-inflammatory prostaglandins.28 Leuko-
`
`Page 5 of 30
`
`

`
`+
`
`3152 Chemical Reviews, 1996, Vol. 96, No. 8
`
`trienes produced by the 5-lipoxygenase enzyme path-
`way may also contribute to both inflammation and
`NSAID-induced effects. Table 8 summarizes the
`concentrations of test compounds required to cause
`a 50% inhibition of 5-LO and CO formation.
`Replacement of the hydroxyl with an amino group
`resulted in more potent activity toward 5-LO while
`the potency toward CO remained the same. How-
`ever, replacement with a thiol resulted in enhanced
`potency toward both 5-LO and CO. Comparison of
`the electronegativity values of oxygen, nitrogen and
`sulphur (Table 8) suggests that this could be a factor
`that modulates the degree of inhibition of 5-LO. Size,
`however, may play a significant role with regard to
`inhibition of CO (Table 7). Thus, the thiol group may
`be a suitable and informative bioisostere for the
`amino and hydroxyl groups in several different series
`of medicinal agents by the virtue of its size, lower
`electronegativity, and ability to hydrogen bond.
`4. Fluorine and Hydroxyl, Amino, or Methyl Groups as
`Replacements for Hydrogen (Grimm’s Hydride
`Displacement Law)
`This monovalent group of isosteres is a result of
`the direct adaptation of Grimm’s Hydride Displace-
`ment Law. The basis for the fluorine-hydrogen
`interchange and the hydroxyl-amino interchange
`was discussed previously. The existence of this
`larger group of isosteres might be attributable to a
`greater tolerance of the different physicochemical
`parameters of these functionalities within a particu-
`lar series of agents. However, in the studies outlined
`in this section, an attempt was made to correlate a
`physicochemical parameter of this group of bioisos-
`teres with the observed effect on biological activity.
`In designing agents for the treatment of cardio-
`vascular diseases, it may be beneficial to associate
`the hypotensive effects resulting from the inhibition
`of angiotensin II formation with the diuretic and
`natriuretic responses. Diuretic and natriuretic ef-
`fects can be mediated by protection of the endogenous
`atrial natriuretic peptide (ANP) from inactivation by
`inhibition of epithelial neutral endopeptidase (NEP).
`Inhibition of angiotensin II formation may be brought
`about by inhibition of endothelial angiotensin-
`converting enzyme (ACE). A series of dual metal-
`lopeptidase inhibitors have been designed on the
`basis of the characteristics of the active sites of both
`enzymes. Monovalent substitution by fluorine, hy-
`droxyl, and amino in place of hydrogen has recently
`been used in the design of these metallopeptidase
`inhibitors (Figure 11, Table 9).30
`In this study optically pure N-[2-(mercaptomethyl)-
`3-phenylbutanoyl] amino acids (17) were evaluated
`as dual inhibitors of NEP and ACE. Substitution
`with isosteres (-F, -OH, -NH2) as described by
`Grimm’s Hydride Displacement Law (Table 2) re-
`sulted in retention of activity. It was observed within
`this series, however, that the increase in the effective
`van der Waal’s radii of the isosteric substituents
`resulted in a decrease in activity (Table 9). In this
`instance, no significant alteration in preferential
`activity with either of the peptidases, ACE or NEP,
`was observed for these bioisosteres.
`The empirical approach used to advance the struc-
`ture-activity relationships with these peptidase
`
`+
`
`Patani and LaVoie
`
`Figure 11.
`Table 9. In Vitro Inhibition of NEP and ACE by
`N-[2-(Mercaptomethyl)-3-phenylbutanoyl] Amino
`Acids
`
`IC50 (nM)
`effective
`angiotensin
`van der Waal’s
`neutral
`converting
`radii (Å)31
`compound R
`endopeptidase
`enzyme
`17a
`2.5
`4.3
`1.20
`H
`17b
`5.9
`6.9
`1.47
`F
`17c
`9.0
`7.9
`1.53
`OH
`17d
`NH2
`1.79
`12.0
`16.0
`inhibitors is useful despite the fact that a more
`selective ACE inhibitor was not developed. Retention
`of activity within this series of bioisosteres permits
`an assessment of the validity of a possible correlation
`with one or more specific physicochemical param-
`eters. This study, for example, did provide insight
`into structural features which were critical to their
`activity as inhibitors of these peptidases.
`Recently, several 8-substituted O6-benzyl guanines
`(18, Figure 12) were evaluated for their ability to
`inactivate the human DNA repair protein, O6-alkyl-
`guanine-DNA alkyltransferase (AGT) (Table 10).32
`Inactivation of the human DNA repair protein O6-
`alkylguanine-DNA alkyltransferase by exposure to
`compounds such as O6-benzylguanine leads to a
`dramatic enhancement in the cytotoxic response of
`human tumor cells and tumor xenografts to chemo-
`therapeutic drugs. This effect is principally observed
`for chemotherapeutic agents whose mechanism of
`action involves modification of DNA guanine residues
`at the O6-position. In this study the effect of the
`interchange of NH2, OH, as well as CF3 (a bioisostere
`for a methyl group based on the replacement of
`hydrogen with fluorine) on activity was assessed.
`Analogues possessing electronegative groups at the
`8-position were more effective as inactivators of AGT
`in human HT29 colon tumor cell extracts. The
`relative activities of these bioisosteres based on the
`dose required for 50% inhibition (ED50) along with
`their electronegativities are outlined in Table 10.
`
`Figure 12.
`Table 10. Alkyl Guanine Transferase Inactivating
`Activity of 6-(Benzyloxy)purine Derivatives
`compound
`R
`electronegativity29
`ED50 ((cid:237)M)a
`18a
`NH2
`2.61
`2.0
`18b
`CF3
`3.46
`0.25
`18c
`OH
`3.51
`0.15
`a Effective dose required to produce 50% inactivation in HT-
`29 cells upon incubation for 4 h.
`
`Page 6 of 30
`
`

`
`+
`
`+
`
`Bioisosterism: A Rational Approach in Drug Design
`
`Chemical Reviews, 1996, Vol. 96, No. 8 3153
`
`Again, the retention of activity within a series of
`bioisosteres provides the basis for the discovery of a
`possible correlation between pharmacological activity
`and the physicochemical properties of specific agents.
`As an extension to the above group defined by
`Grimm’s Hydride Displacement Law, the widespread
`use of the chlorine atom as a bioisostere has been
`observed in several different series of biologically
`active compounds. This could be attributed to the
`similarity in size between these atoms, a comparison
`of which is made in Table 11. Further, there exists
`similarity in the lipophilicity of the methyl group
`with that of chlorine which may be responsible for
`its suitability as a monovalent bioisosteric replace-
`ment.
`N-(Substituted-3-pyridyl)-N ¢
`-alkylthioureas (19, Fig-
`ure 13), which have been evaluated as novel potas-
`sium channel openers,33 are among the more recent
`illustrations of the replacement of chlorine with
`isosteres from Grimm’s Hydride Displacement Law
`(Table 11). Potassium channel openers cause va-
`sorelaxation in vascular smooth muscle through
`hyperpolarization of the cell membrane. There is an
`increased interest in these compounds based on their
`therapeutic potential in the treatment of cardiovas-
`cular diseases. Substitution at the 6-position with
`monovalent isosteres (-NH2, -CH3, -Cl) results in
`analogues with similar biological activity.
`It was
`observed that substituents with similar biological
`activity had comparable effective van der Waal’s radii
`(Table 11). The methyl group, which has a lower
`electronegativity, elicited a weaker pharmacological
`response, suggesting an additional correlation be-
`tween activity and a physicochemical property.
`
`Figure 13.
`
`Table 11. Inhibition of Spontaneous Mechanical
`Activity in Rat Portal Vein (in vitro)
`maximum
`effective
`electro-
`fall in SBPa
`van der Waal’s
`radii31 (Å)
`negativity29
`(%)
`X
`compound
`19a
`1.79
`2.61
`29
`NH2
`19b
`CH3
`1.80
`2.27
`18
`19c
`1.73
`3.0
`27
`Cl
`a Antihypertensive activity measured as maximum % fall
`in systolic blood pressure in anesthetized normotensive rat by
`iv injection.
`
`Table 12 lists the relative potency of a group of
`bioisosteres which act as inhibitors of thymidylate
`synthase. Each of these benzo[f]quinazolin-1(2H)-
`ones (20, Figure 14), inhibit thymidylate synthase by
`virtue of their structural relation to its cofactor, 5,10-
`methylenetetrahydrofolic acid.34 They are, therefore,
`referred to as folate-based thymidylate synthase
`inhibitors. These analogues differ from other folate-
`based thymidylate synthase inhibitors as the absence
`of a glutamate residue suggests that they are not
`dependent upon active folate transport and poly-
`glutamylation for activity, the two mechanisms of
`resistance that have been observed with agents such
`
`Figure 14.
`
`Table 12. Thymidylate Synthase Enzyme Inhibition
`Data for Benzo[f]quinazolin-1(2H)-ones
`thymidylate synthase
`inhibitory activity
`IC50 ((cid:237)M)a
`X
`compound
`20a
`0.025
`Cl
`20b
`CH3
`0.178
`20c
`0.48
`OH
`20d
`NH2
`0.63
`20e
`1.08
`H
`a Inhibitor concentration producing 50% inhibition and
`determined by the tritium release assay of Roberts35 as
`modified by Dev et al.36
`
`as methotrexate. Within this series, it was observed
`that hydrogen bond donors were more potent than
`the unsubstituted parent compound. These ana-
`logues, however, were less active than compact
`lipophilic groups in elevating thymidylate synthase
`inhibition. Thus, at the 9-position, optimal size and
`lipophilicity appear to be critical factors associated
`with their ability to inhibit thymidylate synthase.
`In another study aimed at designing cholinergic
`agents which would be capable of penetrating the
`central nervous system and displaying high efficacy
`at the cortical muscarinic receptors, a series of
`oxadiazole-based tertiary amines 21 (Figure 15, Table
`13) were tested. The assay used was designed to
`measure affinity and predict cortical efficacy from the
`antagonist-agonist (i.e. NMS/OXO-M) binding ratio
`in rat cort

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