throbber
Lipopbilicity and drug activity
`
`By Hugo Kubinyi
`Chemical Research and Development of BASF Pharma Division,
`Knoll AG, D-6700 LudwigshafeniRhein, Germany
`
`Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`1
`The additivity concept. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`2
`Lipophilicity parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3
`3.1 Partition coefficients. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3.2 Calculation of partition coefficients. nand fvalues . . . . . . . . . . . . . . . . .
`3.3 Dissociation and partition coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3.4 The rate constants of drug partitioning. . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3.5 Chromatographic parameters ..... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3.6 Molecular connectivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`3.7 Other parameters related to lipophilicity . . . . . . . . . . . . . . . . . . . . . . . . . .
`4
`Linear dependence of biological activity on lipophilicity. . . . . . . . . . . . .
`5 Nonlinear dependence of biological activity on lipophilicity. . . . . . . . . .
`5.1 Reasons for nonlinear relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`5.2 The parabolic model and related empirical models. . . . . . . . . . . . . . . . .
`5.3 Equilibrium models of drug partitioning. . . . . . . . . . . . . . . . . . . . . . . . . .
`5.4 Kinetic models of drug transport. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`5.5 The bilinear model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`5.6 Models for transport and distribution of ionizable drugs. . . . . . . . . . . . .
`5.7 Consequences for additive de novo models. . . . . . . . . . . . . . . . . . . . . . . .
`6
`The future development of QSAR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`Acknowledgments.. . ... .. ..... . . . .. . ... . . . . . . .. . . ... . .. . . .. . . .
`References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
`
`97
`
`98
`99
`102
`102
`107
`112
`119
`122
`126
`132
`137
`148
`148
`151
`156
`161
`168
`174
`185
`187
`189
`189
`
`H. Kubinyi et al., Progress in Drug Research / Fortschritte der Arzneimittelforschung / Progrès des
` recherches pharmaceutiques © Birkhäuser Verlag Basel 1979
`
`Micro Labs Exhibit 1063
`Micro Labs v. Santen Pharm. and Asahi Glass
`IPR2017-01434
`
`

`

`98
`
`1
`
`Introduction
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`The understanding of drug potency in biological systems requires an un(cid:173)
`derstanding of chemical structures in terms of physical and chemical
`properties: transport and distribution of a drug in a biological multicom(cid:173)
`partment system, the affmity of the drug to a complementary - structur(cid:173)
`ally unknown - receptor, and the interaction of the drug with its receptor
`obviously depend on these properties.
`In drug design the fIrst step is the more or less systematic variation of a
`lead compound to derive some hypotheses of relationships between
`chemical structure and biological activity. In the next step these hypo(cid:173)
`theses are used to arrive at improved derivatives of the original lead com(cid:173)
`pound with minimal effort. The introduction of the Hansch model in
`1964 enabled medicinal chemists to formulate their hypotheses of struc(cid:173)
`ture-activity relationships in quantitative terms and to check these hypo(cid:173)
`theses by means of statistical methods. From such quantitative structure(cid:173)
`activity relationships (QSAR) it is possible to elucidate the influence of
`various physicochemical properties on drug potency and to predict activ(cid:173)
`ity values for new compounds within certain limits.
`The main purpose of this review is to sum up some developments of
`QSAR since 1971, when the review 'On the understanding of drug po(cid:173)
`tency' [1] appeared in this series. Three excellent books on QSAR have
`been published in the meantime, the introductory book by Purcell, Bass
`and Clayton [2] and two comprehensive monographs by Martin [3] and
`Seydel and Schaper [4]. Monographs on selected topics [5, 6], several
`symposia proceedings [7-10], review articles [11-28], and a rapidly in(cid:173)
`creasing number of publications reflect the growing importance of QSAR
`in medicinal chemistry.
`During the last decade QSAR started.to develop from a merely intuitive
`and empirical discipline to a more and more theoretically based science.
`Drug design will remain a sophisticated art all the time; however, from
`QSAR medicinal chemists gained new insights which allow the applica(cid:173)
`tion of more rational approaches, especially in lead structure optimiza(cid:173)
`tion. The largest progress has been made in describing the lipophilicity of
`drugs and in understanding the dependence of drug activity on lipophil(cid:173)
`icity. Therefore special emphasis is placed on lipophilicity and drug ac(cid:173)
`tivity in this review.
`
`Micro Labs Exhibit 1063-2
`
`

`

`Hugo Kubinyi: Lipophilicity and drug activity
`
`99
`
`2
`
`The additivity concept
`
`The fundamental basis of all quantitative structure-activity analyses is the
`concept of additivity: all substructures of a drug are assumed to contrib(cid:173)
`ute to biological activity in an additive manner, each part of the structure
`irrespective of all other variations in the molecule. There is no sharp defi(cid:173)
`nition of the term substructure, each substituent and each partial struc(cid:173)
`ture with unique chemical properties can be regarded as a substructure:
`sometimes single atoms, e.g. the halogens, sometimes larger groups, e.g. a
`sulfonamido or pyridyl group, are taken as substructures.
`The environment of a substructure has of course a significant influence
`on its chemical properties and therefore different activity contributions
`may be observed for identical groups in different positions of a molecule.
`For drugs interacting with a specific receptor additional differences result
`from the asymmetric topology of the receptor. Taking this into considera(cid:173)
`tion the different biological activities of optical enantiomers are compat(cid:173)
`ible with the additivity concept.
`Is there a rationale of the additivity concept of drug-receptor interac(cid:173)
`tions? There is one: if the aflinity of a drug to its receptor depends only
`on the physicochemical properties of the complementary binding sites
`and if these physicochemical properties are additive themselves, also the
`receptor affInity of a drug should be an additive molecular property. It
`must be emphasized that the structure and the physicochemical proper(cid:173)
`ties of the receptor binding site need not be known because this part of
`the system remains constant. Such 'simple' drug-receptor interactions can
`be studied e.g. in isolated enzyme systems or in receptor preparations.
`As far as hydrophobic interactions are concerned, the additivity concept
`can be illustrated by the driving forces of hydrophobic interactions (fig. I )
`[3, 29, 30]. A nonpolar drug and a hydrophobic region of a receptor are
`surrounded by water molecules which are more or less ordered and
`therefore in a higher state of energy than in free solution. In the drug(cid:173)
`receptor complex a smaller number of water molecules is in contact with
`hydrophobic surfaces; the resulting increase in entropy leads to a stabili(cid:173)
`zation of the drug-receptor complex. It is obvious that the gain in free
`energy should be proportional to the number of water molecules chang(cid:173)
`ing from an ordered to an unordered state, i.e. proportional to the surface
`area of the nonpolar part of the drug. Specific polar, electronic and steric
`effects may add to these unspecific interactions.
`
`Micro Labs Exhibit 1063-3
`
`

`

`100
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`R
`
`)
`
`~
`
`Figure I
`b.c O 0 0 0
`Schematic representation of
`00000 0
`0000000""
`hydrophobic interaction.
`0 0°0 0 0 00 ('0
`R= hydrophobic part of
`OO~
`receptor covered by c water
`molecules; D= approaching
`drug enveloped by a + b water
`molecules; D-R= drug-receptor
`interaction complex with a
`representing ordered water
`molecules covering D-Rand
`b + c are the displaced,
`disordered water molecules
`(reprinted from [29] with
`permission of the copyright
`owner).
`
`O- R
`
`In more complex systems like isolated cells, bacteria, isolated organs, or
`whole animals the biological activity of a drug depends not only on the
`receptor affmity but also on the absorption and distribution of the drug:
`the more complex the system is, the more important will the influence of
`absorption and distribution be.
`Lipophilicity is the main factor governing transport and distribution of
`drugs in biological systems. Although these drug characteristics are - for
`a given biological system - unequivocally a function of chemical struc(cid:173)
`ture and time, the relationships are not as simple as in the case of drug(cid:173)
`receptor interactions. Nonlinear lipophilicity-activity relationships have
`been known since long but they could not be described mathematically
`until fifteen years ago; today the understanding of the dependence of
`drug distribution in biological systems on lipophilicity is much better
`than in the early days of QSAR.
`However, there are some other effects which cause departures from the
`additivity concept. Metabolism of drugs is - because of the specifIty of
`the involved enzymes - no simple function of a defmite molecular prop(cid:173)
`erty, but depends on the presence of certain substructures. As long as
`these substructures are common to all molecules of a series, one can be
`confident that a quantitative relationship can be derived for these meta(cid:173)
`bolic conversions too. If the metabolic conversions take place at a
`position of substituent variation, the additivity concept is seriously dis(cid:173)
`turbed. Examples for such metabolic conversions are e.g. the hydroxyl(cid:173)
`ation of aromatic rings, the reduction of nitro groups to amino groups
`and the cleavage of ethers, esters, ami des and amines.
`
`Micro Labs Exhibit 1063-4
`
`

`

`Hugo Kubinyi: Lipopbilicity and drug activity
`
`101
`
`Other nonlinear effects may arise from steric crowding of substituents,
`leading to lower than predicted activity, and cooperative binding, leading
`to higher than predicted activity. While the effects from steric crowding
`are easy to understand, cooperative binding can be explained only in
`thermodynamic terms. Each interaction between a drug substructure and
`the complementary receptor site causes an enthalpy and an entropy
`change. Once a drug molecule is fIxed at its binding site by one or more
`specifIc interactions, no entropy loss will result from further interactions.
`Hence, if two or more substructures of a drug molecule fIt to a comple(cid:173)
`mentary structure, the overall binding force may be higher than the sum
`of the individual contributions. Although all single drug-receptor inter(cid:173)
`actions are weak bonds, the high affinity and specifIty of most drugs can
`be explained by this cooperative effect.
`While we are far from a general mathematical model including all fac(cid:173)
`tors responsible for the relationships between chemical structure and bio(cid:173)
`logical activity, the linear free energy related Hansch model in its linear
`[eq. (1)] and parabolic form [eq. (2)] [31-34] and the de novo model of
`Free and Wilson [eq. (3)] [35] have proven their utility for the quantita(cid:173)
`tive description of such relationships and have confIrmed the additivity
`concept of biological activity group contributions.
`
`10gl/C=alogP+ bO'+c,
`
`logl/C= a(1ogp)2+ blogP+ CO'+ d,
`
`logl/C= ~:ai+ fl..
`i
`
`(I)
`
`(2)
`
`(3)
`
`In these equations C is a molar concentration causing a standard bio(cid:173)
`logical response, e.g. an ED 50 or LD 50. P is the partition coefficient, 0' is
`the Hammett constant, and a, b, c and d are constants determined by
`linear multiple regression analysis. Other physicochemical parameters
`can be used instead of or in addition to P and 0' in equations (1) and (2).
`In equation (3) ai are the values of the substituent group contributions to
`biological activity, and fl. is regarded to be the activity contribution of the
`parent system (in Fujita-Ban analysis [36, 37] fl. is the theoretical bio(cid:173)
`logical activity value of the reference compound). Both the Hansch and
`Free-Wilson analysis have been reviewed in the literature [1-4, 11-19];
`only new developments concerning the methodology will be discussed in
`this review.
`
`Micro Labs Exhibit 1063-5
`
`

`

`102
`
`3
`3.1
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`Lipophilicity parameters
`Partition coefficients
`
`(4)
`
`The lipophilicity of a drug is dermed by the partition coefficient P, which
`is the ratio of equilibrium concentrations of a drug in an organic phase,
`corg> and an aqueous phase, caq [eq. (4)].
`P= corg •
`caq
`n-Octanol/water has proven to be the system of choice for measuring
`partition coefficients for QSAR studies [33, 38, 39] due to its similarity to
`biological systems: like n-octanol biological membranes are made up
`from hydrophobic alkyl chains and polar groups. In addition to this simi(cid:173)
`larity n-octanol has some theoretical and practical advantages as com(cid:173)
`pared with other organic solvents:
`- A much broader spectrum of compounds is soluble in n-octanol than
`in aliphatic or aromatic hydrocarbons; the hydroxyl group of n-octanol
`can act as a hydrogen bond donor as well as an acceptor.
`- While n-octanol is practically insoluble in water, it dissolves an appre(cid:173)
`ciable amount of water (2.3M, which corresponds to a molar ratio ofn-oc(cid:173)
`tanol:water ~4:1) under eqUilibrium conditions. Hydrogen bonds need
`not be broken during the transfer of a solvated drug molecule from the
`aqueous phase to the organic phase. Therefore n-octanol/water partition
`coefficients reflect only hydrophobic interactions, while hydrocarbon!
`water partition coefficients are additionally influenced by desolvation
`energies.
`-
`n-Octanol has low vapor pressure at room temperature.
`It is well suited for direct measurement of concentrations in the ultra-
`-
`violet region due to its low absorption over a wide range.
`-
`n-Octanol/water partition coefficients are available from the literature
`(for a review see [38]) and from the Hansch data bank for a large number
`of drugs. The hydrophobic substituent constant n and the hydrophobic
`fragmental constant f, which allow the calculation of partition coefficients
`(see below), refer to the n-octanol/water system.
`The experimental measurement of partition coefficients has been re(cid:173)
`viewed by Purcell et aL [2], Martin [3] and Rekker [6].
`The use of a model system for drug/biological system interactions is justi(cid:173)
`fied by the Collander equation (5) [40],
`logP2 = alogP 1 + b,
`
`(5)
`
`Micro Labs Exhibit 1063-6
`
`

`

`Hugo Kubinyi: Lipophilicity and drug activity
`
`103
`
`which relates partition coefficients from similar solvent systems. Collan(cid:173)
`der equations with slopes a and correlation coefficients r close to unity
`are observed for homologous series of compounds. As an example the
`partition coefficients of 4-alkylpyridines in different solvent systems [41]
`are correlated with their n-octanol/water partition coefficients in table 1.
`
`Table 1
`Collander equations for the partition coefficients of homologous 4-alkylpyridines
`in different solvent systems (derived from the data ofYeh and Higuchi [41]).
`log P= a log P OCI+ b.
`
`Solvent
`
`a
`
`b
`
`Hexadecane
`Octane
`Butyl ether
`Carbon tetrachloride
`Chloroform
`
`1.036 (±0.04)
`1.064 (±0.04)
`1.024 (±0.04)
`1.064 (±0.04)
`1.047 (± 0.03)
`
`-1.199 (±0.15)
`-1.139(±O.l7)
`-0.781 (±0.16)
`-0.503 (±O.l6)
`0.462 (±0.1O)
`
`n
`
`7
`7
`6
`6
`7
`
`r
`
`s
`
`1.000 0.070
`0.999
`0.080
`1.000 0.072
`1.000 0.071
`1.000 0.044
`
`No such close relationships can be expected for heterogeneous groups of
`compounds because, in addition to hydrophobic interactions, solvation
`forces will influence the interrelationships between different solvent
`systems. Equation (5) has been extended to a large number of partition(cid:173)
`ing systems by Leo and Hansch [38, 42, 43]. While close correlations are
`obtained between n-octanol/water and other polar systems, like ketone/
`water, ester/water and alcohol/water systems [e.g. eq. (6)-(8)], systems
`with less polar organic solvents, like aliphatic or aromatic hydrocarbons,
`or diethyl ether gave poor correlations with the n-octanol/water system.
`
`Oleyl alcohol
`10gP= 0.999(±0.06)10gP oct- 0.575(±0.11)
`(n= 37; r= 0.985; s= 0.225),
`
`primary butanols
`10gP = 0.697 (± 0.02)logP oct + 0.381 (± 0.03)
`(n= 57; r= 0.993; s= 0.123),
`
`methyl isobutyl ketone
`10gP= 1.094(±0.07)logP oct+ 0.050(± 0.11)
`(n= 17; r=0.993; s=0.184).
`
`(6)
`
`(I)
`
`(8)
`
`Micro Labs Exhibit 1063-7
`
`

`

`104
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`Only after categorization of the solutes into 'hydrogen bond donors' and
`'hydrogen bond acceptors' satisfactory correlations could be obtained
`[e.g. eq. (9)-(12)].
`
`Benzene, H-donor solutes
`
`logP= 1.0l5(±0.11)logPoct-1.402(±0.14)
`(n=33; r=0.962; s=0.234),
`
`benzene, H -acceptor solutes
`
`logP= 1.223(±0.19)logP oct- 0.573(±0.20)
`(n= 19; r= 0.958; s= 0.291),
`
`diethyl ether, H-donor solutes
`
`logP= 1. 130(± 0.04)logP oct-O.l70(±0.05)
`(n=71; r=0.988; s=0.186),
`
`diethyl ether, H-acceptor solutes
`
`logP= 1.142(±0.13)logPoct -1.070(±0.12)
`(n= 32; r= 0.957; s= 0.326).
`
`(9)
`
`(10)
`
`(11)
`
`(12)
`
`In the case of chloroform and carbon tetrachloride a third equation was
`needed for 'neutral' solutes, having both donor and acceptor ability [eq.
`(13)-(15)] [42].
`
`Chloroform, H-donor solutes
`
`logP= 1.126(±0.12)logPoct -1.343(±0.21)
`(n= 28; r= 0.967; s= 0.308),
`
`chloroform, neutral solutes
`
`logP= 1. 1O(±0.12)logP oct- 0.649(±0.18)
`(n= 23; r= 0.971; s= 0.292),
`
`chloroform, H-acceptor solutes
`logP= 1.276(±0.14)logPoct+ 0.171(±0.17)
`(n=21; r=0.976; s=0.25l).
`
`(13)
`
`(14)
`
`(15)
`
`Micro Labs Exhibit 1063-8
`
`

`

`Hugo Kubinyi: Lipophilicity and drug activity
`
`105
`
`While the slopes of equations (13)-(15) are parallel, the intercepts show
`significant differences; the intercept of equation (14) is in the middle be(cid:173)
`tween the intercepts of equations (13) and (15), which is to be expected.
`However, the categorization of solutes into donors and acceptors may be
`a too rough differentiation for correlating partition coefficients from dif(cid:173)
`ferent solvent systems. For a group of substituted phenols Leo et aL [38]
`derived equation (16):
`.
`10gP oct = 0.5010gP cyc10hexane + 2.43
`(n=9; r=0.79l; s=0.39l).
`
`(16)
`
`If they accounted for the hydrogen binding ability of the phenols by add(cid:173)
`ing a term 10gKHB [44], they obtained equation (17) with a much better
`correlation coefficient and a significantly lower standard deviation.
`10gP oct = l.00logP cyc10hexane + 1.20logKHB+ 2.35
`(n= 9; r= 0.979; s= 0.140).
`
`(17)
`
`From a much larger set of partition coefficients in cyc10hexane/water and
`aliphatic hydrocarbon/water systems Seiler derived an additive constant
`IH as a measure of the hydrogen binding ability of different substituents
`and substructures [eq. (18)] [45]. Some representative IH values are given
`in table 2.
`
`(18)
`
`Table 2
`IHvalues ofsubstituents and substructures [45].
`
`SubstituenVsubstructure
`
`Aromatic-COOH
`Aromatic-OH
`-CONH-
`-S02NH-
`Aliphatic-OH
`AUphlltic-NH2
`Aromatic-NH2
`-NR IR2 (R I, R2+ H)
`-N0 2
`~ C=O
`-C=N
`-0-
`artha-Substitution to -OH, -COOH, -NRIR2
`
`2.87
`2.60
`2.56
`l.93
`l.82
`1.33
`1.18
`0.55
`0.45
`0.31
`0.23
`0.11
`-0.62
`
`Micro Labs Exhibit 1063-9
`
`

`

`106
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`The continous range of IH values suggests that the boundaries between
`H-donors and H-acceptors are fluid. The regression coefficients a= 1.00
`in equations (17) and (18) give some evidence that all possible Collander
`equations are strictly parallel if the effect of solvation is accounted for by
`introducing IH or a similar parameter into the correlation equation. The
`only exceptions are systems with very polar organic phases, like butanols,
`pentanols, cyclohexanol or butanone, which lead to regression coeffi(cid:173)
`cients a< 1.0 [compare eq. (7)]; these solvents contain so much water
`(approx. 9-17%) under equilibrium conditions, that the organic phase is
`significantly 'diluted'. On the other hand, dilution of the aqueous phase
`with an organic solvent which is only slightly soluble in the organic phase
`leads to a similar decrease of the regression coefficient a [eq. (19)] [38].
`
`Ether/50"110 aqueous dimethylformamide
`
`10gP = O.400logP oct - 0.321
`(n= 6; r= 0.988; s= 0.058).
`
`(19)
`
`Although equations (6)-(15) and similar equations may be used to con(cid:173)
`vert partition coefficients from one solvent system to another system, the
`results should be considered with suspicion. The standard deviations of
`most Collander equations are higher than those of the equations pre(cid:173)
`sented in table 1 and much higher than the standard deviations of experi(cid:173)
`mental measurements. While partition coefficients predicted within a
`closely related series of compounds will be reliable, only approximate
`values may result for structurally diverse compounds.
`Equation (5) also applies to the binding of drugs to organic and bio(cid:173)
`logical macromolecules and to the distribution of organic compounds be(cid:173)
`tween aqueous phases and biological membranes [e.g. eq. (20)-(22)] [11,
`33,38].
`
`Binding of acetanilides to nylon [46]
`
`logK=0.69(±0.23)logP-7.I6(±0.37)
`(n=7; r=0.96I; s=0.203),
`
`(20)
`
`binding of misc. neutral compounds to bovine serum albumin (1: 1) [47]
`
`10gI/C= 0.75I(±0.07)logP+ 2.300(±0.I5)
`(n=42; r=0.960; s=0.I59)
`
`(21)
`
`Micro Labs Exhibit 1063-10
`
`

`

`Hugo Kubinyi: Lipophilicity and drug activity
`
`107
`
`partitioning of alcohols between human erythrocyte membranes and
`buffer [48]
`
`10gP= 1.003(±O.l3)logP oct- 0.883(±0.39)
`(n= 5; r=0.998; s=0.082).
`
`(22)
`
`The slope a= 1.00 in equation (22) is a further justification for the use of
`n-octanoVwater as a model system for the interaction of drugs with bio(cid:173)
`logical systems; slopes between 0.5 and 0.8 are observed for equations
`correlating the binding of organic compounds to biological macromole(cid:173)
`cules with n-octanoVwater partition coefficients, indicating the more or
`less polar nature of these binding sites.
`
`3.2
`
`Calculation of partition coefficients.7l and fvalues
`
`If a linear free energy related model like the Hansch model is used to
`describe and predict biological activities in terms of physicochemical
`parameters, these parameters must be known for all compounds included
`in the analysis. In order to simplify QSAR analyses it is desirable to
`predict physicochemical properties from chemical structures; such pre(cid:173)
`dictions can then be used to estimate the biological activity of new,
`hitherto unknown compounds.
`One of the pioneering achievements of Hansch was the demonstration
`that partition coefficients are - in the logarithmic scale - an additive
`constitutive molecular property, like molar volume, molar refractivity,
`parachor and Hammett (J values. In this context additive means that
`partition coefficients can be calculated by simply adding increments of
`partial structures and constitutive means that the increment values de(cid:173)
`pend on the relative position and environment of the partial structures.
`Hansch et al. [49, 50] defmed the hydrophobic substituent constant 1l
`[eq. (23)] of a substituent X as the difference of 10gP values of the substi(cid:173)
`tuted compound R - X and the unsubstituted compound R - H.
`
`(23)
`
`From eight different series of aromatic compounds 1l values for the most
`important aromatic substituents could be derived [49]. Slightly different
`values were obtained for meta and para substituents indicating some elec(cid:173)
`tronic interactions. Today the most often used values are the 1l values
`
`Micro Labs Exhibit 1063-11
`
`

`

`108
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`derived from substituted phenoxyacetic acids [49] and from substituted
`benzenes [49,51,52]. n- Values [49,53], derived from substituted phenols
`and anilines, should be used for aromatic compounds bearing additional
`electron donor substituents. The differences between nand n- values
`arise from electronic interactions as can be seen from equation (24),
`which correlates n values from substituted phenols and benzenes with
`Hammett a values of the corresponding substituents [49].
`LIn = nphenoI- nbenzene= 0.8230' + 0.061
`(n=24; r=0.954; s=0.097).
`
`(24)
`
`A compilation of nand n- values for selected ortho, meta and para sub(cid:173)
`stituents has been published by Norrington et al. [53], while two parame(cid:173)
`ter collections of Hansch et al. [51, 52] contain some 170 n values derived
`from substituted benzenes.
`The decision which series of n values should be used for the calculation
`of 10gP values depends on the chemical structures of the parent com(cid:173)
`pounds: the closer the relationship between both parent structures, the
`more accurate the calculated partition coefficients. Partition coefficients
`calculated for molecules differing in their parent structure too much from
`the parent structure of the used n system wi11lead to erroneous results, a
`fact which is sometimes neglected in QSAR studies.
`Deviations from n additivity have been observed for polysubstituted aro(cid:173)
`matic compounds, e.g. 1,2,3-trimethoxybenzene [38], for o,o'-disubsti(cid:173)
`tuted phenols [38, 54, 55] and for compounds like salicylic acid, where
`hydrogen bonds between two polar ortho substituents reduce the hydro(cid:173)
`philic nature of both groups [38].
`The extension of the n concept to aliphatic compounds led to some
`serious problems. First, the set of aliphatic n values derived by Hansch et
`al. [50] is based on a too small number of data points. Secondly, the use
`of equation (25) [13] is an incorrect application of the n concept [eq.
`(23)] [6].
`
`(25)
`
`Thirdly, no differentiation was made between nCH3 and nCH2. As a con(cid:173)
`sequence, the n concept failed in the calculation of some 10gP values,
`e.g. the calculation of the 10gP value of 1,2-diphenylethane from
`logPc6H6 and nCH3 [eq. (26)] [56].
`
`Micro Labs Exhibit 1063-12
`
`

`

`Hugo Kubinyi: Lipophilicity and drug activity
`
`10gP C6H5CH2CH2C6H5= 2logP C6H6 + 27rCH3= 4.26+ 1.00= 5.26
`exptl.1ogP=4.79 [38].
`
`109
`
`(26)
`
`Folding of co-substituted phenylpropanes, C6H5CH2CH2CH2X, with in(cid:173)
`tramolecular hydrophobic bonding was assumed to explain deviations
`between observed and calculated partition coefficients [38, 57].
`From the deftnition of 7r values [eq. (23)] obviously 7rH must be zero;
`however, this does not imply that the lipophilicity contribution of a hy(cid:173)
`drogen atom per se has to be zero [58]. Rekker [59] reinvestigated the
`problem of folding and criticized the inappropriate application of the 7r
`concept. He formulated a new system of hydrophobic fragmental con(cid:173)
`stants f, defmed by equation (27) [6, 59-61].
`10gP= L ali.
`
`(27)
`
`i
`While 7r values are lipophilicity contributions relative to hydrogen sub(cid:173)
`stitution, the fi values of equation (27) are absolute contributions of sub(cid:173)
`stituents and substructures to the totallipophilicity; ai indicates how often
`a given fragment occurs in the structure. Statistical evaluation of some
`300 partition coefficients led to a ftrst set of aromatic and aliphatic f val(cid:173)
`ues which was further refmed upon by consideration of proximity effects
`and other corrective terms discussed below.
`The main difference between the 7r and f system results from the fact that
`there is a hydrophobic fragmental constant of the hydrogen atom, fH
`=0.175. No branching corrections are needed for aliphatic compounds
`because different f values for -CH3, -CH2-, -CH- and quaternary C
`resulted from the regression analyses. However, the differences between
`these values are not regular (different sets of similar f values - depending
`on the compounds and correction factors used in the calculation - have
`been published by Rekker et al. [6, 59, 60]; the values presented here and
`below are taken from the appendix of [6]):
`
`fcH3 =0.702} A=0.172
`fCH2 = 0.530} A = 0.295
`fCH =0.235 }
`fc =0.15
`(fH=0.175).
`The correct application of Rekker f values requires the consideration of
`certain corrections:
`
`A=0.085
`
`Micro Labs Exhibit 1063-13
`
`

`

`110
`
`Hugo Kubinyi: Lipophilicity and drug activity
`
`(I) Proximity effects: if two electronegative groups like -OH, -0-,
`-COOH or -NH2 are separated by only one or two saturated carbon
`atoms, the overalllipophilicity is higher than predicted; correction factors
`0.861 and 0.574, respectively, must be added.
`(2) Hydrogen attached to an electronegative group: a hydrogen atom
`bound to -COOH, -COOR, -COR, -CONH2' etc. has an increased fH
`value of 0.462 instead of the normal value fH=0.175.
`(3) Conjugated and cross-conjugated systems: if two aromatic rings are
`conjugated (e.g. biphenyl) or cross-conjugated (e.g. benzophenone), a
`correction factor of 0.28 must be added.
`(4) Condensed aromatic systems like naphthalene require a correction
`factor of 0.31 for each pair of carbon atoms common to two aromatic
`rings.
`Whether the correction factors and the differences between aliphatic and
`aromatic f values can be explained by a 'magic' constant cM = 0.28 or
`multiples of this value [6], is still subject to discussion. However, the cor(cid:173)
`rect application of the f system does not depend on the right answer to
`this more or less philosophical problem.
`The most impressive result from the use of equation (27) is that 10gP val(cid:173)
`ues of aliphatic and araliphatic compounds can be calculated with high
`accuracy; the problems of folding and intramolecular hydrophobic bonds
`do not exist any longer.
`
`Calculation oflogP C6HS(CH2)3Cl from 7l values:
`
`10gP= 10gPbenzene+ 371CH3 + 7laliph.Cl= 2.13+ 1.50+0.39=4.02. (28)
`
`Calculation oflogP C6Hs(CH2hCl from fvalues:
`
`10gP= fC6Hs + 3fcH2 + faliph.Cl= ·1.886+ 1.590+ 0.061 = 3.537
`exptl.1ogP=3.55 [50].
`
`(29)
`
`Starting from equation (27), but using a different approach, Leo et al.
`[62] derived another set of f values. From the 10gP values of hydrogen
`and some lower hydrocarbons they calculated strictly additive fcH3,
`fcH2, fCH' fc and fH values (see table 3) and used these values for the cal(cid:173)
`culation of all other f values. The disadvantage of this approach is that
`Leo et al. had to consider negative correction factors fb for 'each single
`carbon-carbon bond after the ftrst one' when calculating 10gP values of
`
`Micro Labs Exhibit 1063-14
`
`

`

`Hugo Kubinyi: Lipopbilicity and drug activity
`
`III
`
`larger molecules. A comparison of some representative f values and cor(cid:173)
`rection factors is given in table 3 together with aromatic 7t values. As ex(cid:173)
`pected from the defInition of 7t and f values, the aromatic f values of both
`
`Table 3
`Hydrophobic fragmental constants f and 7t values of selected substructures and
`substituents.
`
`Fragment
`
`Fragmental constants f
`Reller [6]
`aliph.
`
`arom.
`
`Leo et al. [62]
`aliph.
`arom.
`
`0.175
`0.702
`0.530
`0.235
`0.15
`1.886
`1.688
`1.431
`
`-0.462
`0.061
`0.270
`0.587
`-1.491
`-1.581
`-0.954
`-1.428
`-1.825
`-0.939
`-1.970
`-1.703
`0.757
`-1.066
`0.00
`-0.51
`
`0.399
`0.922
`1.131
`1.448
`-0.343
`-0.433
`-0.093
`-0.854
`-0.964
`-0.078
`-1.109
`-0.842
`1.331
`-0.205
`0.62
`0.11
`
`H
`CH3
`CH2
`CH
`C
`CJIs
`CJI4
`CJI3
`
`F
`CI
`Dr
`I
`OH
`-0-
`COOH
`NH2
`NH
`N02
`CONH2
`>C=O
`CF3
`C=N
`SH
`-s-
`
`Correction factors
`Reller
`
`0.23
`0.89
`0.66
`0.43
`0.20
`1.90
`
`-0.38
`0.06
`0.20
`0.60
`-1.64
`-1.81
`-1.09
`-1.54
`-2.11
`-1.26
`-2.18
`-1.90
`
`0.37
`0.94
`1.09
`1.35
`-0.40
`-0.57
`-0.03
`-1.00
`-1.03
`-0.02
`-1.26
`-0.32
`
`-1.28
`
`-0.34
`
`-0.79
`
`0.03
`
`Leo et al.
`
`7tbenzene
`[51]
`
`0.00
`0.56
`
`1.96
`
`0.14
`0.71
`0.86
`1.12
`-0.67
`
`-0.32
`-1.23
`
`-0.28
`-1.49
`
`0.88
`-0.57
`0.39
`
`Proximity effect 1
`Proximity effect 2
`H on electronegative group
`Aryl conjugation
`Condensed aromatic system
`
`0.861
`0.574
`0.287
`0.28
`0.31
`
`fb(chain)
`fb (cyclic)
`Chain branching
`Group branching
`
`-0.12
`-0

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