throbber
Bioorganic & Medicinal Chemistry 7 (1999) 851–860
`
`Estimation of the Binding A(cid:129)nities of FKBP12 Inhibitors Using
`a Linear Response Method
`
`Michelle L. Lamb,y Julian Tirado-Rives and William L. Jorgensen*
`
`Department of Chemistry, Yale University, New Haven, CT 06520-8107, USA
`
`Received 15 September 1998; accepted 2 November 1998
`
`Abstract—A series of non-immunosuppressive inhibitors of FK506 binding protein (FKBP12) are investigated using Monte Carlo
`statistical mechanics simulations. These small molecules may serve as sca€olds for chemical inducers of protein dimerization, and
`have recently been found to have FKBP12-dependent neurotrophic activity. A linear response model was developed for estimation
`of absolute binding free energies based on changes in electrostatic and van der Waals energies and solvent-accessible surface areas,
`which are accumulated during simulations of bound and unbound ligands. With average errors of 0.5 kcal/mol, this method pro-
`vides a relatively rapid way to screen the binding of ligands while retaining the structural information content of more rigorous free
`energy calculations. # 1999 Elsevier Science Ltd. All rights reserved.
`
`Introduction
`
`The binding protein of the immunosuppressant natural
`product FK506 (Fig. 1) has been the target of extensive
`investigation by both biochemical and theoretical tech-
`niques during the last 10 years. Discovery of the cis-
`trans peptidyl-prolyl
`isomerase (PPIase or rotamase)
`activity of FKBP12 (MW=12 kDa) led to dissection of
`the rotamase mechanism1–4 and hopes for the rapid
`design of low molecular weight immunosuppressant
`molecules that inhibited this activity. The crystal struc-
`ture of FK506 bound to FKBP12 was crucial to this
`endeavor, as it demonstrated that the peptidomimetic a-
`ketoamide and pipecolyl portions of the ligand were
`buried but
`that much of
`the macrocycle remained
`exposed to solvent.5 However, in later studies rotamase
`inhibition was found to be insu(cid:129)cient for immunosup-
`pression; the FK506–FKBP12 complex associates with
`the surface of calcineurin (CN), a serine/threonine
`phosphatase, and hinders binding of subsequent pro-
`teins in the T-cell signaling pathway.6 The ‘‘e€ector’’
`region of FK506, which contacts CN, is opposite the a-
`ketoamide-pipecolic acid moiety.7,8 Thus, the PPIase
`inhibitors developed through structure-based design
`e€orts (e.g. compounds 1–7 in Table 1) formed a set of
`potential
`sca€olds
`for
`immunosuppressive
`e€ector
`components.9,10
`
`Key words: Monte Carlo; linear response; FKBP12; rotamase inhibi-
`tors.
`*Corresponding author. Fax: +203-432-6299; e-mail: bill@adrik.chem.
`yale.edu
`y Current address: Dept. of Pharmaceutical Chemistry, University of
`California, San Francisco, Box 0446, San Francisco, CA 94143-0446.
`
`these non-immunosuppressive FKBP12
`Many of
`ligands are also able to promote neuronal growth in
`vitro and in vivo through binding to FKBP12.11–13
`Dose-response studies of neurite outgrowth in chick
`dorsal root ganglia resulted in an ED50 of 0.058 nM for
`for example.11 In addition,
`tethered
`compound 10,
`dimers of molecules similar to 4 have been used to
`‘‘chemically-induce’’ dimerization14 of targeted cellular
`proteins that had been adapted to include FKBP12
`domains.15 Modification of targets in appropriate sig-
`naling pathways can result in apoptosis or the induction
`of transcription; the prospects of this technology for
`gene therapy are intriguing.
`
`Consequently, we have used theoretical techniques to
`probe this class of small molecules (Table 1) to gain
`insight
`into the physical basis
`for di€erences
`in
`FKBP12-binding activity and to evaluate new methods
`for the calculation of protein–ligand binding a(cid:129)nities.
`In previous work,16 relative free energies of binding
`((cid:1)(cid:1)Gb) for compounds 2, 4, and 8–10 were calculated
`by the free energy perturbation (FEP) technique using a
`Metropolis Monte Carlo (MC) algorithm for config-
`urational sampling.17 The parameters and geometry of
`one ligand were transformed into those of another with
`these simulations, both in solution and while bound to
`the protein. At each step of the transformation, the
`system was brought to equilibrium and the free energy
`di€erence relative to the previous step was computed.
`The di€erence in binding free energy for the two ligands
`was then found from the di€erence in the total free
`energy change for each (bound and unbound) transfor-
`mation. While theoretically rigorous and accurate, cal-
`culations of this type are computationally demanding
`
`0968-0896/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved.
`P I I : S 0 9 6 8 - 0 8 9 6 ( 9 9 ) 0 0 0 1 5 - 2
`
`

`
`852
`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`Table 1. FKBP12 inhibitorsa
`
`Compound
`
`Ki,app, nM
`
`Figure 1. The immunosuppresent drug FK506.
`
`and impractical for a large set of structurally diverse
`ligands.
`
`An attractive alternative has emerged in linear interac-
`tion energy or linear response techniques (LR), as
`recently reviewed.18 Unlike most FEP calculations,
`absolute free energies of binding ((cid:1)Gb) for protein–
`ligand systems are estimated, and simulations of non-
`physical states (intermediate steps) are not required. In
`other respects, the molecular dynamics (MD) or MC
`simulation protocols are the same.
`
`
`
`(cid:11) (cid:135) a (cid:1)ELennard(cid:255)Jones
`
`
`
`
`As first proposed, the linear interaction energy method
`employs eq (1) for the estimation of binding free ener-
`gies.19
`(cid:1)Gb (cid:136) b (cid:1)ECoulomb
`(cid:133)1(cid:134)
`These terms represent the change in interaction between
`a ligand and its environment (solvent and/or protein)
`upon binding. The electrostatic energy di€erences are
`obtained from average Coulombic interaction energies
`between the ligand and solvent (E Coulomb
`l(cid:255)water ) and the ligand
`and protein (E Coulomb
`l(cid:255)FKBP ) accumulated during simulations
`of the bound and unbound ligands. The Lennard–Jones
`(van der Waals) energy di€erences are found in an ana-
`logous manner. The original value of b=0.5 is con-
`sistent with analyses of the response of polar solutions
`to changes in electric fields, such as the charging of an
`ion in water. Observed linear correlations between the
`molecular size (surface area, chain length) and solvation
`free energies of hydrocarbons suggested van der Waals
`interactions might respond linearly as well. A scale fac-
`tor a=0.161 was obtained empirically from fitting to
`experimental binding data for a small set of endothia-
`pepsin inhibitors.19 Recently, A˚ qvist and co-workers
`have advocated the assignment of one of four reduced
`values of b according to the charge-state or polarity of
`the ligand and have obtained an appropriate value for a
`by fitting to data for a larger set of protein–ligand
`pairs.20–22 An extended linear response equation,23,24 in
`which a cavitation term based on solvent-accessible sur-
`face areas (SASA) is added and all parameters are
`empirically determined, is given in eq (2) below. This
`model was first applied to the calculation of free energies
`of hydration for organic solutes; however, optimization of
`
`(cid:11)
`
`aRotamase inhibition data taken from refs 9 and 10 unless otherwise
`noted.
`bData from ref 11.
`cKi reported for a mixture of R and S isomers.
`
`

`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`853
`
`scale factors to b=0.131, a=0.131 and g=0.014 yielded
`acceptable estimates of
`thrombin-binding a(cid:129)nity as
`well.25 In addition, the extended linear response equation
`has potential for predictions of other properties important
`for pharmacological activity, such as the estimation of
`ligand lipophilicity.24 It has been observed that the con-
`tribution of the (cid:1)SASA term is often nearly constant, so
`
`
`the case in which the simple addition of a constant
`improves the fit has also been considered [eq (3)].20,25
`h
`(cid:1)Gb (cid:136) b (cid:1)ECoulomb
`
`
`(cid:1)Gb (cid:136) b (cid:1)ECoulomb
`
`(cid:11) (cid:135) g (cid:1)SASA
`(cid:11) (cid:135) a (cid:1)ELennard(cid:255)Jones
`
`
`(cid:11) (cid:135) a (cid:1)ELennard(cid:255)Jones
`
`
`(cid:11) (cid:135) g
`
`i
`(cid:133)2(cid:134)
`
`(cid:133)3(cid:134)
`
`In contrast to most proteins studied previously with this
`technique, FKBP12 has a distinctly hydrophobic binding
`pocket lined with aromatic residues, and only two inter-
`molecular hydrogen bonds are observed in the crystal
`structure of inhibitor 4 with the protein.9 Consequently, it
`was expected that the electrostatic behavior of these mole-
`cules might deviate from linear response and that the van
`der Waals contributions to binding might be larger than
`previously observed for other systems. To determine the
`suitability of the linear response approximation for binding
`to FKBP12, MC simulations of the bound and unbound
`states for all 11 inhibitors in Table 1 were performed.
`
`Computational Details
`
`The modeling strategy employed here was consistent
`with the FEP study described previously,16 based on the
`4-FKBP12 crystal structure (1FKG)9 and the OPLS
`force field.26–28 (A full listing of parameters for these
`molecules may be found in ref 29) The first five inhibi-
`tors were easily built from 4. Atoms of the 1-phenyl
`substituent were simply removed to obtain 2, and for
`compound 1, a united-atom cyclohexyl group30 was
`positioned in the plane of the original 3-phenyl ring.
`The vinyl group of 3 was also represented with united-
`atom parameters and was positioned at the minimum of
`the CH3(cid:255)C(cid:255)CH(cid:136)CH2 torsional energy profile (180.0(cid:14))
`and aligned with an edge of the 1-phenyl ring of 4.29
`One side of this ring was within 3.2 A˚ of His87 and
`Tyr82, while the other was positioned more than 3.5 A˚
`from Phe46 and Glu54. Accordingly, the orientation
`which maximized hydrophobic contact between the
`protein and 3 was chosen. Inhibitor 5 incorporated the
`crystal structure orientations of the cyclohexyl and tert-
`pentyl groups from 5-FKBP12 (1FKH).9 In 1 and 5, the
`cyclohexyl groups were treated as rigid units, as were all
`ligand and protein aromatic groups. It was thought that
`the conformational flexibility within the rings of these
`substituents would be less important to binding than the
`overall flexibility of the ligand, and thus sampling was
`focused accordingly. Starting geometries
`for
`the
`unbound ligand simulations were taken from these
`initial bound conformations.
`
`All simulations were performed using the MCPRO
`program and Monte Carlo configurational sampling.31
`
`The ligands and protein–ligand complexes were solvated
`with 22 A˚
`spheres of TIP4P water molecules. First,
`1(cid:2)106 (1 M) configurations of water-only equilibration
`with preferential sampling of molecules close to the inhi-
`bitor was performed, followed by 16 M configurations of
`sampling of the entire system with all solvent molecules
`sampled uniformly. Next, data was collected for 8 M
`configurations averaged in blocks of 2(cid:2)105 configura-
`tions. To ensure convergence, averaging for all unbound
`ligands was extended to 16 M configurations.
`
`To obtain protein-bound structures of 6 and 7, the final
`conformations of 4 and 5 above were epimerized within
`the FKBP12 binding pocket via a slow perturbation
`protocol. Eight sequential double-wide windows with
`(cid:1)l=10.0625 were used. In each window, 4 M config-
`urations were sampled to slowly transform between the
`two ligands in an energetically reasonable way. This
`procedure was repeated with the free ligands in solution.
`The simulations of ligands 8–11 were started from the
`final
`structures of FEP simulations
`reported pre-
`viously.16,29 In each case, energy components were
`averaged over 8 M (bound) and 16 M (unbound)
`configurations of the system.
`
`As was mentioned above, the initial structure for 2 had
`been generated from the bound conformation of 4 with the
`1-phenyl atoms removed. Following an FEP calculation
`from 4 to 2, further sampling was carried out within the
`first windows of both a phenyl!pyridyl FEP16 and a car-
`bonyl!hydroxyl FEP.29 The final conformations of these
`windows were then used to start two additional linear
`response MC simulations, 2a and 2b, respectively. These
`additional data points provide one gauge of precision.
`
`Solvent-accessible surface areas for the ligands in both
`aqueous and protein environments were calculated
`using the SAVOL2 program.32 This algorithm has been
`incorporated into MCPRO, and the necessary atomic
`radii are calculated from the corresponding OPLS s
`parameters via 1/2 (21/6s). Using the standard solvent
`radius for water, 1.4 A˚ , the SASAs of the ligands were
`calculated for the structure at the end of each block of
`MC configurations and were averaged.
`
`Average energy and solvent-accessible surface area dif-
`ferences were fit
`to the experimental binding data
`(Table 1) to obtain linear response parameters a, b, and
`g according to eqs (1)–(3). This procedure was per-
`formed with a Simplex-based algorithm. As inhibition
`data for the majority of the ligands in this set has been
`reported by Holt and co-workers,9,10 in cases where two
`values have been measured the values from these
`authors were used for fitting.
`
`Results and Discussion
`
`Average intermolecular energy components and SASAs
`
`Each ligand and protein–ligand complex was solvated
`with a sphere of explicit water molecules and sampled as
`described above. Average energy components and
`
`

`
`854
`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`ligand SASAs accumulated during the simulations are
`reported in Table 2. As has been observed with throm-
`bin inhibitors, the ligand–solvent Coulombic energy
`components, E Coulomb
`l(cid:255)water , fluctuate the most during the
`simulations.25 However, both the magnitude of the
`energy contribution and the magnitude of the fluctua-
`tion are reduced compared to the positively charged or
`zwitterionic protease inhibitors, as expected for the
`neutral ligands of this set. All of the ligands in solution
`have undergone hydrophobic collapse relative to their
`bound conformations but to di€ering extents, as sig-
`nificant flexibility is observed in the propyl side chain.
`The Lennard–Jones interactions for the a-ketoamide
`molecules in solution scale generally with molecular
`size; 1 and 2 (and 8–10) have the least favorable average
`energies, followed by compound 3, and finally there is a
`small range of energies for 4–7. The solvent accessible
`surface area of amide carbonyl O3 is generally ca. 10 A˚ 2
`less than that of the keto (O4) or ester (O2) atoms, and
`usually only one water molecule is found to interact
`with this atom. This is consistent with molecular
`dynamics results for trans-FK506.29 Often, O4 interacts
`strongly with one water molecule (H-O4 distance, ca.
`1.8 A˚ ), while a second molecule hovers 0.5 A˚
`further
`away. Also, water molecules are found to interact with
`the faces of aromatic rings. In the case of 2, an unusual
`long-lived solvation pattern is noted. The water mole-
`cule that hydrogen bonds to O3 further associates with
`one directed into the center of the less-accessible face of
`the phenyl ring (Fig. 2).
`
`Typical of many FKBP12-ligand complexes,33 the crys-
`tal structure of 4-FKBP12 contains contacts between
`
`Figure 2. The hydrogen bonding network in 2, with waters that bridge
`between the amide O3 and phenyl ring.
`
`O4 and aromatic hydrogens of Tyr26, Phe36, and Phe99,
`with the pipecolyl ring sitting over Trp59.9 The 3-phe-
`nylpropyl moiety binds in the solvent-exposed FK506-
`cyclohexyl groove of FKBP12 between Ile56 and Tyr82,
`and these residues form hydrogen bonds with the ester
`and amide carbonyl oxygens of the ligand. The 1-phenyl
`substituent interacts with Phe46 and the tertiary pentyl
`group of the inhibitor. The two crystallographic inter-
`molecular hydrogen bonds are maintained throughout all
`of the FKBP12-ligand simulations, and none of the
`ligands deviates significantly from the original orientation
`
`Table 2. Average interaction energies (kcal/mol) and solvent-accessible surface areas of the inhibitors (A˚ 2) from the aqueous and FKBP12 MC
`simulationsa
`
`Compound
`
`1 aq
`1FKBP
`2 aq
`2a aq
`2b aq
`2FKBP
`2aFKBP
`2bFKBP
`3 aq
`3FBKP
`4 aq
`4FKBP
`5 aq
`5FKBP
`6 aq
`6FBKP
`7 aq
`7FKBP
`8 aq
`8FBKP
`9 aq
`9FBKP
`10 aq
`10FKBP
`11 aq
`11FKBP
`
`ECoulomb
`l(cid:255)water
`(cid:255)23.16(0.32)
`(cid:255)0.24(0.15)
`(cid:255)31.28(0.46)
`(cid:255)29.91(0.35)
`(cid:255)27.11(0.38)
`(cid:255)5.15(0.31)
`(cid:255)2.14(0.19)
`(cid:255)1.62(0.19)
`(cid:255)27.68(0.36)
`(cid:255)1.41(0.20)
`(cid:255)31.89(0.42)
`(cid:255)5.30(0.38)
`(cid:255)28.43(0.50)
`(cid:255)6.49(0.19)
`(cid:255)32.84(0.55)
`(cid:255)8.00(0.19)
`(cid:255)29.64(0.34)
`(cid:255)3.92(0.23)
`(cid:255)30.53(0.43)
`(cid:255)13.51(0.48)
`(cid:255)29.88(0.50)
`(cid:255)1.43(0.19)
`(cid:255)33.45(0.40)
`(cid:255)11.95(0.31)
`(cid:255)42.57(0.43)
`(cid:255)2.72(0.34)
`
`EL(cid:255)J
`l(cid:255)water
`(cid:255)35.96(0.15)
`(cid:255)15.09(0.12)
`(cid:255)34.59(0.21)
`(cid:255)33.90(0.16)
`(cid:255)33.26(0.16)
`(cid:255)13.74(0.09)
`(cid:255)13.99(0.12)
`(cid:255)13.95(0.14)
`(cid:255)36.52(0.16)
`(cid:255)15.31(0.13)
`(cid:255)39.63(0.22)
`(cid:255)17.84(0.11)
`(cid:255)40.12(0.19)
`(cid:255)16.54(0.10)
`(cid:255)38.35(0.12)
`(cid:255)18.10(0.13)
`(cid:255)38.48(0.15)
`(cid:255)16.10(0.15)
`(cid:255)32.75(0.18)
`(cid:255)12.03(0.15)
`(cid:255)31.67(0.17)
`(cid:255)14.04(0.09)
`(cid:255)32.89(0.22)
`(cid:255)12.82(0.10)
`(cid:255)31.24(0.21)
`(cid:255)15.19(0.10)
`
`ECoulomb
`l(cid:255)FKBP
`
`(cid:255)19.14(0.16)
`
`(cid:255)19.73(0.21)
`(cid:255)21.56(0.16)
`(cid:255)20.97(0.20)
`(cid:255)21.15(0.20)
`(cid:255)21.25(0.15)
`(cid:255)20.58(0.14)
`(cid:255)17.26(0.15)
`(cid:255)17.09(0.18)
`(cid:255)18.03(0.16)
`(cid:255)23.59(0.16)
`(cid:255)20.02(0.18)
`(cid:255)18.23(0.15)
`
`EL(cid:255)J
`l(cid:255)FKBP
`
`(cid:255)39.53(0.14)
`
`(cid:255)40.29(0.13)
`(cid:255)38.01(0.17)
`(cid:255)37.92(0.21)
`(cid:255)44.06(0.15)
`(cid:255)42.73(0.19)
`(cid:255)44.00(0.16)
`(cid:255)38.46(0.20)
`(cid:255)43.65(0.21)
`(cid:255)38.80(0.19)
`(cid:255)39.40(0.17)
`(cid:255)38.77(0.15)
`(cid:255)38.61(0.13)
`
`SASAb
`
`665.4(1.2)
`202.1(1.6)
`653.9(1.9)
`629.3(2.3)
`620.6(2.3)
`202.9(5.1)
`222.4(6.8)
`188.6(7.0)
`680.9(1.2)
`221.2(3.0)
`715.1(1.4)
`248.8(2.5)
`728.3(1.8)
`244.9(2.1)
`712.2(1.6)
`232.4(3.0)
`688.3(1.9)
`282.3(7.1)
`618.4(1.8)
`176.0(4.7)
`624.9(2.2)
`179.3(6.4)
`630.8(2.2)
`183.2(6.5)
`626.7(1.4)
`185.5(4.4)
`
`aThe standard error of the means is given in parentheses.
`bCalculated from structures saved every 2(cid:2)105 configurations, N=40 (8 M) or 80 (16 M). (Standard deviations range from 10–50 A˚ 2.)
`
`

`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`855
`
`of 4. In general, the aromatic rings of the inhibitors are
`oriented perpendicularly to the ring of Tyr82, although a
`clearly T-shaped interaction is not dominant. Consistent
`with the hydrophobic binding pocket, the largest con-
`tribution to the protein–ligand interaction energy in all
`cases comes from the Lennard–Jones terms. In addition,
`the most favorable van der Waals interactions with
`FKBP12 are observed for the more hydrophobic 3, 5,
`and 7 compared to their more aromatic or smaller
`counterparts.
`
`An estimate of the precision of the simulations was
`determined from the three simulations of 2 in solution.
`One model was derived directly from the 4-FKBP12
`crystal structure, while the others (2a and 2b) were
`obtained through earlier FEP simulations from 4. In
`solution, there is a 4 kcal/mol range in the E Coulomb
`1(cid:255)water
`components and a 2 kcal/mol range in E L(cid:255)J
`l(cid:255)water. Varia-
`tions among the average energy components reported
`for all simulations of 2-FKBP12 are on the order of 2–
`3 kcal/mol. It must be noted that the terms from simu-
`lations 2a and 2b are more similar to each other than to
`those from the original 2 simulation.
`
`As energy and SASA di€erences between the protein
`and aqueous environments are the quantities that are
`scaled to estimate binding a(cid:129)nity, these values are
`recorded in Table 3. Two of the lowest a(cid:129)nity inhibi-
`tors, 3 and 11 have the largest van der Waals energy
`di€erences between bound and unbound ligands.
`Ligands 6, 7, and 11 have the most unfavorable
`(cid:1)ECoulomb, while only 8 has a net favorable electrostatic
`interaction upon binding. Approximately 450 A˚ 2 of each
`ligand is buried upon binding to FKBP12. This repre-
`sents 65% of the surface of inhibitor 4, for example. The
`largest change is noted for inhibitors 5 and 6 (ca. 480 A˚ 2
`buried), while the smallest di€erence is found for 7 (ca.
`410 A˚ 2).
`
`Optimization of the LR equations
`
`The energies and surface areas of Table 2 were used first
`with previously determined scaling parameters to compute
`
`FKBP12-binding a(cid:129)nities. The original parameters of
`A˚ qvist19 do not describe binding to this protein well at
`all; the free energies of binding are underestimated with
`an average unsigned error (<jerrorj>) of 9.3 kcal/mol.
`Results with the thrombin-derived parameters are of
`more appropriate magnitude ((cid:255)9.3 (5) to (cid:255)6.3 (11) kcal/
`mol, <jerrorj>=1.4 kcal/mol), although both the set
`of atomic radii for SASAs and the all-atom representa-
`tion of the thrombin inhibitors di€ered from the study
`presented here. However, to improve the correspon-
`dence with experiment for FKBP12, new values for a, b,
`and g were found by fitting the average energy and sur-
`face area di€erences to the experimental binding free
`energies. The results for the various models investigated
`are summarized in Table 4.
`
`The first step was to employ eq (1) with b=0.5 and
`derive an appropriate value for a, as was done originally
`for endothiapepsin.19 This model resulted in a=0.626
`and yielded free energies which deviated significantly
`from experiment. In particular, the maximum unsigned
`error for model 1 was 4.4 kcal/mol, which diminishes its
`predictive value given that the range of experimental
`binding a(cid:129)nities is 3.4 kcal/mol. Furthermore, com-
`pound 3, a poor inhibitor, was predicted to bind as well
`as the highest a(cid:129)nity ligand, 8.
`
`As expected, the linear response assumption for elec-
`trostatic energies, b=0.5, does not appear to hold well
`for binding to FKBP12. When the value of b is set
`based on ligand composition20 and a derived empiri-
`cally with eq (1) (model 2) or both b and a treated as
`free parameters (model 3), average unsigned error and
`RMS to experiment were improved but the maximum
`errors are still greater than 2kcal/mol. With model 2, the
`a-ketoamide ligands required b=0.43; 11 called for
`b=0.37 due to its single hydroxyl substituent.20
`
`Fitting the data with three parameters yielded an RMS
`deviation of 0.7 kcal/mol, whether or not the SASA dif-
`ference was included explicitly (eq (2) versus eq (3)).
`Values of b=0.139, a=0.194, and g=0.0145 (model 4)
`ranked the ligands in a qualitatively reasonable way
`
`Table 3. Calculated energy and surface-area di€erencesa with representative binding a(cid:129)nities
`(cid:1)EL(cid:255)J
`
`(cid:1)SASA
`
`Compound
`
`(cid:1)ECoulomb
`
`1
`2
`2a
`2b
`3
`4
`5
`6
`7
`8
`9
`10
`11
`
`3.8
`6.4
`6.2
`4.5
`5.1
`5.3
`1.4
`7.6
`8.6
`(cid:255)1.0
`4.8
`1.5
`21.6
`
`(cid:255)18.7
`(cid:255)19.4
`(cid:255)18.1
`(cid:255)18.6
`(cid:255)22.7
`(cid:255)20.9
`(cid:255)20.4
`(cid:255)18.2
`(cid:255)21.3
`(cid:255)18.1
`(cid:255)21.8
`(cid:255)18.7
`(cid:255)22.6
`
`(cid:255)463.3
`(cid:255)451.0
`(cid:255)406.9
`(cid:255)432.0
`(cid:255)459.7
`(cid:255)466.3
`(cid:255)483.4
`(cid:255)479.8
`(cid:255)406.0
`(cid:255)442.4
`(cid:255)445.6
`(cid:255)447.6
`(cid:255)441.2
`
`model 2
`(cid:255)9.6
`(cid:255)8.9
`(cid:255)8.2
`(cid:255)9.2
`(cid:255)11.5
`(cid:255)10.2
`(cid:255)11.6
`(cid:255)7.6
`(cid:255)9.0
`(cid:255)11.3
`(cid:255)11.0
`(cid:255)10.6
`(cid:255)5.5
`
`model 4
`(cid:255)9.8
`(cid:255)9.4
`(cid:255)8.5
`(cid:255)9.2
`(cid:255)10.4
`(cid:255)10.1
`(cid:255)10.8
`(cid:255)9.4
`(cid:255)8.8
`(cid:255)10.1
`(cid:255)10.0
`(cid:255)9.9
`(cid:255)7.8
`
`aUnits: kcal/mol and A˚ 2, respectively.
`bReferences given in notes to Table 1, (cid:1)G=RTlnKi.
`cCompound 3 not included in the derivation of this model.
`dFree energy estimated using Ki=450 nM.
`
`(cid:1)Gb, kcal/mol
`
`model 6
`
`9.7
`(cid:255)9.4
`(cid:255)8.6
`(cid:255)9.3
`((cid:255)10.9)c
`(cid:255)10.3
`(cid:255)10.9
`(cid:255)9.0
`(cid:255)9.3
`(cid:255)10.2
`(cid:255)10.5
`(cid:255)10.0
`(cid:255)7.8
`
`exptb
`(cid:255)9.2((cid:255)9.2)
`(cid:255)9.5((cid:255)9.0)
`(cid:255)9.5((cid:255)9.0)
`(cid:255)9.5((cid:255)9.0)
`(cid:255)9.0
`(cid:255)10.9((cid:255)10.6)
`(cid:255)11.1
`(cid:255)8.7d
`(cid:255)8.9
`(cid:255)9.2((cid:255)9.4)
`(cid:255)10.1
`(cid:255)11.1
`(cid:255)7.7
`
`

`
`856
`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`Table 4. Summary of the a, b, and g parameters determined by fitting to the experimental (cid:1)Gb dataa
`max. jerrorjc
`
`gb
`
`<jerrorj>c
`
`RMS to exptc
`
`Model
`
`eq
`
`b
`
`a
`
`1
`2
`3
`4
`
`5
`6
`
`7
`
`1
`1
`1
`2
`
`3
`2
`
`3
`
`0.50d
`0.43d,e
`0.201
`0.139
`0.138(cid:139)0.016
`0.145
`0.176
`0.174(cid:139)0.013
`0.180
`
`0.626
`0.599
`0.536
`0.194
`0.191(cid:139)0.006
`0.134
`0.348
`0.348(cid:139)0.041
`0.328
`
`0.0145
`0.0146(cid:139)0.0003
`(cid:255)7.75
`0.0084
`0.0085(cid:139)0.0002
`(cid:255)4.21
`
`4.39
`2.49
`2.22
`1.39
`
`1.12
`1.09
`
`1.12
`
`1.24
`1.01
`0.71
`0.57
`
`0.55
`0.47
`
`0.47
`
`1.75
`1.25
`0.90
`0.72
`
`0.69
`0.58
`
`0.59
`
`aEnergy components and SASAs from MC simulations of 1, 2, 2a, 2b, and 3–11 were fit unless otherwise noted in the text. Cross-validation values
`with uncertainties given in italics.
`beq (2), kcal/mol.A˚ 2, eq (3), kcal/mol.
`cMaximum unsigned error, average unsigned error, and RMS deviation to experiment in kcal/mol.
`dFixed value.
`eb=0.37 for hydroxyl-containing ligand 11 (see text).
`
`(Table 3); again 3 was predicted to bind too well and
`was responsible for the largest deviation from experi-
`ment (1.4 kcal/mol). In performing the ‘‘leave-one-out’’
`cross-validation of these scaling parameters, it became
`clear that compound 3 was the notable outlier. Thus,
`the fitting was repeated without including data for this
`compound, resulting in optimal parameters b=0.176,
`a=0.348, and g=0.0084 (model 6). The predicted order
`of a(cid:129)nities for the remaining ligands improved slightly,
`and there was a corresponding decrease in RMS devia-
`tion to 0.6 kcal/mol. The average unsigned error in this
`case improves by 0.1 kcal/mol to 0.5 kcal/mol. A plot of
`calculated versus experimental binding free energies based
`on model 6 is shown in Figure 3. Additional justification
`
`for setting aside compound 3 may be found when g is a
`constant term in the linear response model. When 3 is
`included in the data set (model 5), the predicted binding
`energy results largely from the constant rather than the
`scaled energy terms (Table 4). However, when the fitting
`the value of g
`is performed without 3 (model 7),
`((cid:255)4.21 kcal/mol) accurately reflects the nearly constant
`value of g*(cid:1)SASA found with model 6.
`
`Overall, with or without 3, the LR calculations are
`ordering the binding a(cid:129)nities well. In view of the sta-
`tistical noise in the simulations and the experimental
`uncertainties, average errors of 0.5 kcal/mol are as good
`as can be hoped for.
`
`Figure 3. Calculated versus experimental (cid:1)Gb plotted for all ligands. The binding free energies were estimated using optimal parameters b=0.176,
`a=0.348, and g=0.0084 and eq (2). Compound 3, not included in this model, is not shown.
`
`

`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`857
`
`Structural observations
`
`With reasonable reproduction of the observed binding
`free energies, we now turn to structures obtained during
`the simulations for possible insights into the origins of
`the predicted a(cid:129)nities. The benchmark example in our
`earlier work was the di€erence between the 3-phenyl-
`propyl compound, 2, and the (R)-1,3-diphenylpropyl, 4.
`The reduced binding for 2 was attributed to a loss of
`hydrophobic contacts and specific aryl C(cid:255)H. . .N,O
`interactions with the protein upon removal of the 1-
`phenyl substituent.16 The relative free energy of binding
`computed with the LR method for any simulation of 2
`and 4 ((cid:1)(cid:1)Gb=0.9–1.8 kcal/mol) is in excellent agree-
`ment with the experimental and FEP-calculated values
`of 1.4 kcal/mol. Appropriately, the conformations of 2
`and 4 bound to FKBP12 resemble those reported pre-
`viously, although at the end of the simulations, the
`phenyl rings in these complexes are found nearer to
`Tyr82 than Glu54 (not shown). The mobility of the
`ligands within that region of
`the protein is seen
`throughout the MC simulations.
`
`Among the final structures of the bound simulations of
`compound 2, the two from the FEP-generated struc-
`tures (2a and 2b) are more similar to one another than
`to that of 2, especially with regard to the dicarbonyl
`O(cid:136)C(cid:255)C(cid:136)O dihedral angle, which is (cid:255)108(cid:14) in 2a and 2b
`and (cid:255)125(cid:14) in 2. However, the electrostatic energy dif-
`ferences for 2 and 2a are very similar (6.4 and 6.2 kcal/
`mol, respectively), while 2b yields 4.5 kcal/mol. There is
`also variation in the (cid:1)SASA terms for each model of 2
`that parallels the trend in contribution to binding from
`(cid:1)EL(cid:255)J for these ligands, in that 2 is most favorable, 2b
`is intermediate, and 2a is the least favorable. As a result,
`the predicted binding free energies for 2 and 2b agree
`well with the experimental values (Table 3), while 2a is
`estimated to bind 1 kcal/mol less well.
`
`As one would expect, compound 1, which contains no
`aromatic ring, has the least favorable average electro-
`static interaction of all ligands in solution ((cid:255)23.16 kcal/
`mol); however, its van der Waals interactions with sol-
`vent are comparable to those of 2. The flexible propyl
`side chain allows the cyclohexyl ring of 1 to pack
`against the side chain of Ile56 (Fig. 4). It is also observed
`that much less water is found within 3 A˚ of the cyclo-
`hexyl group than the phenyl group,
`in both bound
`(Fig. 4) and unbound structures.
`
`One of the best inhibitors, 5, is notable for its packing
`among aromatic side chains of the protein. This is
`found both in its crystal structure with FKBP129 and in
`the structures modeled here. One He of Tyr82 is 2.8 A˚
`from the ligand phenyl ring center and this protein side
`chain fills the space between the tert-pentyl and phenyl
`groups (Fig. 5). Furthermore, the pipecolyl and cyclo-
`hexyl rings are parallel to one another and separated by
`Phe46. The cyclohexyl moiety has more contact with
`Phe46 than does the analogous aromatic group of 4.
`Without these contacts in solution, the cyclohexyl ring
`is free to rotate ca. 30(cid:14) into a position still parallel to
`but displaced from the pipecolyl ring.
`
`Figure 4. Final simulated structures of 1-FKBP12 and 2-FKBP12.
`Protein residues and solvent molecules within 3 A˚ of the ligands are
`displayed.
`
`Figure 5. Snapshots from bound and unbound simulations of 5.
`
`The experimental and calculated binding a(cid:129)nities of 6
`and 7 are much reduced from their (R)- counterparts, 4
`and 5; the scaled electrostatic and van der Waals energy
`di€erences from model 6 each contribute ca. 1 kcal/mol
`to the less favorable (cid:1)Gb for these molecules. Thus, it
`appears that the bound conformations found for these
`ligands through epimerization of
`their counterparts
`(Fig. 6) are reasonable models. When bound to the
`protein, the two phenyl rings of 6 form a ‘‘parallel-
`stacked and displaced’’ structure with one another; and
`the distance between ring centers is 4.7 A˚ . This align-
`ment of aromatic rings has also been noted previously
`for benzene dimer in solution27 but is not seen in the
`simulations of 4-FKBP12. In 7-FKBP12, the Tyr82HZ-
`O3 hydrogen bond is the shortest of all complexes,
`1.7 A˚ , and the interaction between Ile56H and O2 is the
`longest at 2.6 A˚ . Unlike the 1-phenyl group of 6, the
`cyclohexyl ring in 7 interacts with Phe46 more closely
`than with its own tert-pentyl moiety.
`
`

`
`858
`
`M. L. Lamb et al. / Bioorg. Med. Chem. 7 (1999) 851–860
`
`the protein. This ‘‘bump-hole’’ design could be used to
`avoid unproductive binding of
`ligand by wild-type
`FKBP12 in the cytoplasm in a chemically-induced
`dimerization scheme.34,35 The electrostatic solute–sol-
`vent energy is 12–15 kcal/mol more favorable for the
`hydroxyl-containing 11 than its a-ketoamide parent,
`which is consistent with the preferential solvation of
`isopropanol compared with acetone36 and the ability of
`alcohols to be incorporated into the hydrogen bonding
`networks of water.37 The last structure in the unbound
`simulation has the hydroxyl hydrogen pointing towards
`ring (H(cid:255)O4(cid:255)C9(cid:255)C8 dihedral angle,
`the pipecolyl
`(cid:255)22(cid:14); H(cid:255)O3 distance, 2.4 A˚ ). It does make the expected
`two hydrogen bonds to water molecules, one as a
`donor, one as an acceptor. When bound to the protein,
`however there are no solvent molecules directly inter-
`acting with the ligand in the O4 binding pocket. The
`hydroxyl hydrogen of the ligand is positioned directly
`between and below the Hd and He of Phe36; these aro-
`matic hydrogens
`interact with O4 (Fig

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