throbber
3928
`
`J. Med. Chem. 1998, 41, 3928-3939
`
`Articles
`
`Investigations of Neurotrophic Inhibitors of FK506 Binding Protein via Monte
`Carlo Simulations
`
`Michelle L. Lamb and William L. Jorgensen*
`
`Department of Chemistry, Yale University, New Haven, Connecticut 06520-8107
`
`Received January 27, 1998
`
`The binding and solution-phase properties of six inhibitors of FK506 binding protein (FKBP12)
`were investigated using free energy perturbation techniques in Monte Carlo statistical
`mechanics simulations. These nonimmunosuppressive molecules are of current interest for
`their neurotrophic activity when bound to FKBP12 as well as for their potential as building
`blocks for chemical inducers of protein dimerization. Relative binding affinities were computed
`and analyzed for ligands differing by a phenyl ring, an external phenyl or pyridyl substituent,
`and a pipecolyl or prolyl ring. Such results are, in general, valuable for inhibitor optimization
`and, in the present case, bring into question some of the previously reported binding data.
`
`Introduction
`The R-ketoamide functionality of the immunosup-
`pressant natural product FK506 (Figure 1) is retained
`in many of the highest affinity ligands that have been
`developed to inhibit the rotamase (cis-trans peptidyl-
`prolyl isomerase, or PPIase) activity1 of the FK506
`binding protein (FKBP12, MW ) 12 kDa).2 Originally,
`interpretation of the crystal structure of FK506-
`FKBP12 led to the belief that the R-ketoamide mimics
`a twisted-amide transition state of peptide bond isomer-
`ization, although an endogenous substrate for FKBP12
`had not been discovered. It was thought that blockage
`of the isomerase active site prevented modification of
`downstream proteins necessary for T-cell activation, and
`this was the source of the observed immunosuppression.
`A similar mechanism had been proposed for the activity
`of the undecapeptide cyclosporin A (CsA), which inhibits
`the PPIase cyclophilin, although neither the natural
`products nor the proteins are homologous. However,
`evidence that rotamase inhibition was not sufficient for
`immunosuppression soon began to mount.3 Rapamycin
`(Figure 1), another fungal molecule structurally similar
`to FK506, inhibited FKBP12 but appeared to influence
`a later stage of the T-cell cycle. Schreiber and co-
`workers4 made a significant contribution with the syn-
`thesis of a molecule which retained the FKBP12 binding
`domain of FK506 and rapamycin (pyranose ring, R-ke-
`toamide, pipecolate ester, and cyclohexyleth(en)yl groups),
`but in which the macrocycle was contracted. This
`molecule was a rotamase inhibitor but did not prevent
`T-cell proliferation.
`It later became clear that the formation of an immu-
`nosuppressant-immunophilin complex results in a gain
`of function for the protein. The CsA-cyclophilin and
`FK506-FKBP12 pairs each present a recognition sur-
`face to the calcium-dependent, serine/threonine phos-
`phatase, calcineurin (CN).5 The FK506-FKBP12 com-
`plex binds at least 10 Å from the active site of CN and
`
`Figure 1. Structures and atom numbering for the immuno-
`suppressants FK506 and rapamycin.
`
`S0022-2623(98)00062-4 CCC: $15.00 © 1998 American Chemical Society
`Published on Web 09/19/1998
`
`

`
`Neurotrophic Inhibitors of FK506 Binding Protein
`
`Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21 3929
`
`Table 1. Experimental Activities for Selected FKBP12 Ligands
`
`a Data on neurite outgrowth from chick dorsal root ganglia reported in ref 2. b Data from Guilford Pharmaceuticals, refs 2 and 59.
`c Data from SmithKline Beecham, refs 13 and 16.
`
`must block binding of subsequent phosphorylated pro-
`teins and thus the T-cell signaling pathway.6,7 Reports
`of the association of calcium channels containing -Leu-
`Pro- sequences with both FKBP12 and CN are filling
`in another long-standing piece of the FKBP12 puzzle,
`as these may represent endogenous “ligands” for FKBP12
`mimicked by FK506.8 In contrast, rapamycin-FKBP12
`interrupts a distinct signaling cascade through its
`interaction with another protein, generally termed
`FRAP (FKBP-rapamycin-associated protein).9-11 A
`crystallographic structure of this ternary complex con-
`firms the recognition requirements for rapamycin.12 In
`both FKBP12 ligands, it is the portion of the macrocycle
`opposite the R-ketoamide-pipecolic acid moiety, the
`“effector” region, which contacts calcineurin.
`As part of an effort to design low molecular weight
`PPIase inhibitors as scaffolds for the immunosuppres-
`sive effector components, the crystal structure of
`1-FKBP12 (Table 1) was solved at SmithKline Beecham
`in 1993.13 Figure 2 shows the binding mode revealed
`for the R-ketoamide and pipecolyl portion of 1. The keto
`carbonyl (O4) contacts aromatic hydrogens of Tyr26,
`Phe36, and Phe99, and the pipecoline ring sits over Trp59.
`The 3-phenylpropyl moiety binds in the solvent-exposed
`FK506-cyclohexyl groove of FKBP12 between Ile56 and
`
`Figure 2. Position of compound 1 (yellow) in the aromatic
`binding pocket of FKBP12 (green).13 Molecular graphics im-
`ages were produced using the MidasPlus software system from
`the Computer Graphics Laboratory, University of California,
`San Francisco.60
`
`Tyr82, and these residues form hydrogen bonds with the
`ester (O2) and amide (O3) carbonyl oxygens of the
`ligand. The 1-phenyl substituent interacts with Phe46
`and the tertiary pentyl group of the inhibitor. A
`
`

`
`3930 Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21
`
`Lamb and Jorgensen
`
`Scheme 1
`
`Figure 3. FKBP12-bound conformation of 1 (yellow) overlaid
`with that of FK506 (red).13,15
`
`comparison of the bound conformation of 1 and FK506
`presented in Figure 3 demonstrates that this mode is
`consistent with that found in crystallographic structures
`of FK506-FKBP12 and rapamycin-FKBP12.14,15 How-
`ever, the ability to form hydrogen bonds to the Glu54
`carbonyl observed in complexes with FK506 (C24-OH)
`and rapamycin (C28-OH) is not present in this ligand.
`Binding patterns similar to those for 1 may be expected
`for compounds 2 and 3 (Table 1) as well.13,16 An ex-
`cellent analysis of FKBP12-ligand interactions, includ-
`ing discussion of previously unpublished atomic struc-
`tures, is included in a review of protein-ligand recog-
`nition motifs by Babine and Bender.17
`An additional activity for rotamase inhibitors of this
`class has expanded interest in these compounds beyond
`their potential in immunosuppressant drug design. As
`reviewed recently by Hamilton and Steiner,2 FK506 has
`been shown to induce the regeneration of damaged
`nerves in animal models of Parkinson’s and Alzheimer’s
`diseases. Furthermore, the enriched concentration of
`FKBP12 in neurons has been associated with nitric
`oxide synthesis, neurotransmitter release, and neurite
`extension. Potent, nonimmunosuppressive FKBP12
`ligands, such as V-10,36718-20 and GPI-1046 (6, Table
`
`1),21-23 are able to promote neuronal growth in vitro and
`in vivo without the addition of exogenous growth factors.
`They have a better therapeutic potential than growth
`factors in that they are orally bioavailable and able to
`cross the blood-brain barrier. The requirement of
`binding to FKBP12 for neuronal activity has also been
`demonstrated, but there is no linear relationship be-
`tween rotamase inhibition and activity in neuronal
`cells.2 FKBP12 binding is apparently necessary but not
`sufficient for stimulation of nerve growth, suggesting
`
`that, as in T-cells, the complex may modify the function
`of an additional target.
`Another use of this class of FKBP12 ligands has also
`emerged. The ability of the immunosuppressants to
`induce protein heterodimerization and the knowledge
`of ligand modifications that prevent this association has
`been exploited for control of cellular signaling pathways,
`protein translocation, and gene activation.24,25 Target
`proteins are first artificially attached to the immuno-
`philins (FKBP12 or cyclophilin), CN, or FRAP. The
`ligands themselves or synthetic homo- or heterodimers
`of FK506, CsA, or rapamycin then bring their protein
`partners together, resulting in the proximity of the tar-
`get proteins and transmission of signal.24-30 Recently,
`dimers of 7 have been used to effect cellular apoptosis
`and to induce transcription, again without the immu-
`nosuppressive effects of further binding to calcineurin.31
`This technique of “chemically induced dimerization”,
`used with small, cell-permeable molecules such as 7, is
`
`designed to have application in cellular gene therapy.
`Given the diverse biological applications of these
`R-ketoamide ligands and that only slight differences in
`structure can have profound effects on activity, we have
`used theoretical techniques to probe the binding of
`compounds 1-6 (Table 1) at the atomic level, in both
`structural and energetic terms. Previous simulations
`of FKBP12 have addressed the rotamase mechanism
`applied to peptide substrates32,33 and the importance of
`Tyr82 in binding FK506.34 Our current approach has
`focused on free energy perturbation (FEP) calculations,
`using Monte Carlo (MC) methods rather than molecular
`dynamics (MD) for sampling. Computed relative free
`energies of binding, which are obtained from simula-
`tions of the ligands in solution and bound to the protein,
`may be compared with those obtained from experimen-
`tal binding constants (Scheme 1). Averages of the
`computed structures may then be used to analyze the
`origin of the differences in binding affinities.
`The MC method used here has been validated with a
`study of benzamidine inhibitors of trypsin35 and was
`further applied to the analysis of orthogonal CsA-
`cyclophilin pairs as components of a system for chemi-
`cally induced dimerization.36 The present study is
`aimed at understanding factors that influence the
`binding of 1 and its analogues. In particular, the effects
`of removal of the 1-phenyl group, conversion of the
`3-phenyl to 3-(3-pyridyl), and ring contraction of the
`
`

`
`Neurotrophic Inhibitors of FK506 Binding Protein
`
`Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21 3931
`
`pipecolyl ring to prolyl are examined. There are dis-
`crepancies in the binding data from the two experimen-
`tal sources, as indicated by the results for 2 and 3 in
`Table 1. From the crystal structure for 1 bound (Figure
`2), the pyridine nitrogen of 3 is anticipated to be solvent
`exposed. Thus, it would normally not be expected to
`favor the lower dielectric environment of a protein ((cid:15) (cid:25)
`4) over that of bulk water ((cid:15) (cid:25) 80),37 in contrast to the
`binding results from Guilford Pharmaceuticals. This
`was pursued through computations for the 2, 3 and 5,
`6 pairs. Hamilton and Steiner have also pointed out
`that 5 and 6 are the first examples of prolyl compounds
`that bind better than their pipecolyl analogues, but the
`high affinity is attributed only to “improved design”.2
`To investigate further, differences in free energies of
`binding were computed for two pairs of pipecolyl and
`prolyl ligands. Compounds 2 and 5 represent the
`unusual case with the prolyl ligand (5) as the better
`inhibitor. Compounds 1 and 4 represent the more
`common situation in which the presumably more hy-
`drophobic pipecolyl ligand (1) has higher affinity for
`FKBP12.
`
`Computational Details
`Parametrization and Initial MC Simulations.
`The crystal structure of 1-FKBP12 at 2.0 Å resolution13
`from the Brookhaven Protein Data Bank38 (entry 1fkg)
`was used as the starting point for the simulations. The
`computational protocol for the MC simulations was the
`same as in previous applications.35,36 The good precision
`that is obtainable for free energy changes with this
`methodology was addressed extensively in ref 35. The
`MC sampling included variation of all bond angles and
`dihedrals of the ligand and protein side chains as well
`as overall rotation and translation of the ligand and
`water molecules. The protein backbone atoms were held
`fixed in their crystallographic positions. This makes the
`MC simulations more rapid, and the approximation is
`justified for FKBP12. Restricted backbone motion on
`the picosecond time scale has been noted for native
`FKBP12,39 and ligand binding further rigidifies the
`protein structure, as demonstrated by the close resem-
`blance among atomic structures of FKBP12 in numerous
`FKBP12-ligand complexes.17 To be consistent with
`prior MD calculations on the FK506-FKBP12 system,40
`all 79 residues within 12 Å of FK506 in its cocrystal
`structure with FKBP1214 were sampled. This provided
`a greater number of moving side chains than would be
`found in a 12 Å region around 1.
`The OPLS united-atom force field41 with all-atom
`aromatic groups42 provided most parameters for the
`protein; parameters for the inhibitors also came from
`this source and from a previous MD study of FK506.43
`A listing of parameters for the inhibitors is provided in
`the Supporting Information. The torsional parameters
`for the amino acid residues were derived from fitting to
`torsional energy profiles obtained from ab initio calcula-
`tions with the 6-31G* basis set.44 Any missing param-
`eters were derived by fitting to MM245 energy profiles,
`which were generated using Macromodel.46 A scale
`factor of 1/2 was applied to all 1-4 nonbonded interac-
`tions. Histidines 25, 87, and 94 are known to be un-
`protonated,47 and they were designated as (cid:228)-tautomers
`based on visual inspection. This tautomeric state has
`
`also been chosen in MD simulations of FKBP12-ligand
`complexes in solution.32,34
`The unbound ligands and protein-ligand complexes
`were solvated with 22 Å spheres containing 1477 and
`939 TIP4P water molecules, respectively. A half-
`harmonic potential with a 1.5 kcal/mol Å2 force constant
`was employed to prevent waters from migrating away
`from the cluster. A 9 Å residue-based cutoff was used
`for all nonbonded interactions; if any pair of atoms from
`two residues was within this distance, all nonbonded
`interactions between the residues were included in the
`energy evaluation. The list of nonbonded interactions
`was updated every 2 (cid:2) 105 configurations during the
`simulations.
`All Monte Carlo simulations were performed with the
`MCPRO program.48 An advantage of using internal
`coordinate MC methods is the ability to focus sampling
`on specific regions and degrees of freedom of interest.
`Consequently, bond lengths were fixed to their crystal
`structure values, and aromatic rings were treated as
`rigid units. To prevent inversion at sp3 centers such
`as R-carbons and to enforce planarity of sp2 centers for
`more efficient sampling, improper dihedral angles were
`not varied except as noted below. Otherwise, all bond
`angles and dihedrals in the moving portion of the system
`were sampled.
`The MC simulations were carried out for 25 °C on
`Silicon Graphics workstations and on a cluster of
`personal computers using Pentium processors. It may
`be noted that the experimental results come from an
`assay for rotamase inhibition.49 This widely used
`procedure for measuring FKBP12 binding affinities is
`usually performed somewhat below room temperature,
`e.g., near 10 °C.13 The solvent was first sampled for 1
`million (M) configurations to remove any highly repul-
`sive initial contacts with the solutes. Then, 8M con-
`figurations were performed to equilibrate the 1-FKBP12
`complex. The same protocol was followed for 1 in
`solution, beginning with the bound conformation taken
`from the 1-FKBP12 structure. During equilibration,
`the conformation of the bound ligand remained sim-
`ilar to the crystal conformation; however, partial in-
`version of the pipecolyl ring occurred in solution to
`switch it from a chair to a half-chair conformation
`(Figure 4). In gas-phase optimizations of ligand 1 with
`the present force field, the adopted ring conformation
`is favored by 1.3 kcal/mol. This is likely an artifact of
`using the AMBER C2-N-CH bending parameters with
`ı0 ) 118°, which was not designed for a piperidine
`ring.50 The difference is expected to have little effect
`on the computed free energy changes since the mutated
`phenyl rings are not in contact with the pipecolyl ring
`or, in the case of the ring contraction, the chair con-
`formation was enforced (vide infra).
`Free energy changes were calculated during the MC
`simulations according to standard procedures of statis-
`tical perturbation theory.51-53 The difference in free
`energy of binding (¢¢Gb) for molecule B relative to
`molecule A (Scheme 1) may be obtained from transfor-
`mations of the ligands in solution and bound to the
`protein according to eq 1:
`¢¢Gb(AfB) ) ¢GB - ¢GA ) ¢GFKBP - ¢Gaq
`FEP Simulation Protocol for 1f2. This perturba-
`
`(1)
`
`

`
`3932 Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21
`
`Lamb and Jorgensen
`
`Figure 4. Stereoviews of unbound ligand 1. The initial geometry from the 1-FKBP12 crystal structure has the pipecolic ester
`substituent in an axial conformation (top). Subsequent equilibration resulted in partial inversion of the ring (bottom).
`
`tion involved the removal of the 1-phenyl ring of 1 to
`obtain 2. The atoms of the phenyl group were converted
`to “dummy” atoms without charge or Lennard-Jones
`parameters, and the length of the bond connecting the
`substituent to the remainder of the ligand was reduced
`to 0.65 Å with all other phenyl ring bonds reduced to
`0.35 Å. The transformation of 1f2 was carried out in
`13 windows with double-wide sampling, which yield 26
`free energy increments.51 A coupling parameter, (cid:236), was
`employed such that (cid:236) ) 0 corresponds to the initial
`state, 1, and (cid:236) ) 1 corresponds to the final state, 2. The
`first six windows used ¢(cid:236) ) (0.025, while the remain-
`ing windows used ¢(cid:236) ) (0.050. All were equilibrated
`with 2-4M configurations of sampling; the last config-
`uration of the previous window was used to start the
`next one. Averaging was done in batches of 2 (cid:2) 105
`configurations, with data collected over a total of 4-7M
`configurations in each window. For subsequent analy-
`ses of hydrogen bonding, an additional 1M configura-
`tions were generated at the endpoints of the simula-
`tions.
`FEP Simulation Protocol for 2f3, 5f6. The next
`transformation addressed was the conversion of a phen-
`yl moiety to a 3-pyridyl ring. This perturbation is
`straightforward; the analogous perturbation of benzene
`to pyridine had been performed in the development of
`OPLS all-atom (OPLS-AA) parameters for pyridine.54
`As before, the standard phenyl ring structure was
`transformed to a pyridine geometry determined from
`microwave experiments.54 A model of 5 was required
`prior to the conversion of 5f6 and was obtained by
`mapping a prolyl ring onto the final structure of 2 from
`the 1f2 FEP calculation. In each simulation, the prolyl
`or pipecolyl ring was flexible. The perturbation protocol
`for these calculations was slightly modified from that
`
`used for 1f2 to take advantage of the acquisition of a
`new parallel computing system within our laboratory.
`Seven double-wide windows were run in parallel, with
`4-8M configurations sampled during the equilibration
`phase and with data collected over 4-12M configura-
`tions. A gas-phase FEP calculation was also performed
`for 5f6 to allow estimation of the relative free energies
`of hydration of the two ligands.
`FEP Simulation Protocol for 2f5, 1f4. In our
`experience, perturbations between different cyclic sys-
`tems require much care to implement and can be
`particularly slow to converge. The necessity of account-
`ing for both changes in bonded and nonbonded interac-
`tions within the ring as one atom disappears makes this
`a technically difficult perturbation. One way to simplify
`the present calculations is to drive the ring from one
`fixed six-membered ring conformation to a fixed five-
`membered ring conformation, “disappearing” the re-
`maining atom and simultaneously reeling it in toward
`the others. For this rigid perturbation, changes in
`energy within the ring need not be monitored, as these
`intraligand differences should be very similar in each
`environment (bound and unbound). However, other
`possible conformations for the rings would not be taken
`into account, and the results could be sensitive to the
`path chosen.
`The simulations for the unbound and bound trans-
`formations were started from the final bound conforma-
`tions of 2-FKBP12 or 1-FKBP12 above with the chair
`conformation for the pipecolyl ring. The final prolyl ring
`geometry was obtained from a gas-phase optimization
`of the bound conformation of 2 with one ring atom
`converted to a dummy atom, as illustrated in Figure 5.
`Other than for the internal structures of the pipecolyl
`and prolyl rings, the sampling for the ligands included
`
`

`
`Neurotrophic Inhibitors of FK506 Binding Protein
`
`Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21 3933
`
`Figure 5. The initial and final ring conformations of the 2f5 perturbation. The dummy atom in 5 is noted Du.
`
`all bond angles and dihedral angles, as before. Prior to
`the FEP calculation, the unbound ligands, 2 and 1, were
`each resolvated and relaxed with 4.4M configurations
`of solvent-only sampling, followed by 8M configurations
`of full equilibration. Twenty-one windows were used
`to perform the ring contraction in small steps. Fortu-
`nately, convergence was rapidly achieved within the 4M
`configurations performed for both equilibration and
`averaging.
`
`Results and Discussion
`Effect of the 1-Phenyl Group, 1f2. Removal of
`the 1-phenyl moiety from 1 is a large perturbation but
`a computationally attractive choice given the available
`structural information and the sizable difference in
`binding affinity (Table 1). The free energy change as a
`function of (cid:236) from the FEP calculations for the unbound
`and bound ligands proceeded smoothly (Figure 6a), with
`¢Gaq ) 10.25 ( 0.31 kcal/mol and ¢GFKBP ) 11.69 (
`0.31 kcal/mol (Table 2). The resultant relative free
`energy of binding (¢¢Gb) of 1.4 kcal/mol obtained
`according to Scheme 1 is then in excellent agreement
`with the experimental observations of 1.4-1.6 kcal/mol
`(Table 2).
`A comparison of the averaged structures from the
`simulations with the crystal structure yields several
`interesting observations. First, the average root-mean-
`squared (rms) deviations for non-hydrogen atoms be-
`tween the structures sampled for the complex of 1 and
`the crystal structure13 were computed. The average rms
`for the atoms in the side chains that were varied is 0.7
`Å, and the average rms for the ligand 1 is 1.1 Å. The
`corresponding maximum rms values for individual
`structures did not exceed 0.8 and 1.4 Å, and the values
`at the end of the MC run were 0.7 and 0.9 Å. Thus, the
`MC sampled structures did not drift far from the crystal
`structure, though some differences emerged.
`In the
`R-ketoamide region of the ligand, the crystallographic
`interaction of O4 with an (cid:15)-hydrogen of Phe99 is not
`maintained, but there is a frequent interaction of the
`œ-hydrogen with the amide carbonyl oxygen (O3). The
`hydrogen bonds to Ile56 and Tyr82 are unaffected by the
`perturbation (Figure 2). Within compound 1, the 1-phen-
`yl and isopentyl groups remain in contact. However,
`the 1-phenyl group moves away some from Phe46; the
`shortest contact between aromatic carbons increases
`from 4.5 Å in the crystal structure to 5.6 Å in the
`simulation. In both the crystal structure and from the
`
`Figure 6.
`(a) Free energy profile for the transformation of
`1f2, with error bars for each window shown. (b) Profiles for
`2f3 and 5f6 (squares).
`(c) Profiles for 2f5 and 1f4
`(squares). Solid lines represent the unbound simulations, and
`the dashed lines result from simulations of FKBP12-ligand
`complexes.
`
`MC simulations, the 1-phenyl and isopentyl groups pack
`well into the hydrophobic pocket that is outlined by
`Phe46, Phe36, Ile90, Ile91, Tyr82, and His87 (Figure 2). Loss
`of hyrdrophobic contacts upon removal of the 1-phenyl
`group is unfavorable for binding.
`The simulations also suggest that some specific
`contacts with the 1-phenyl group may be relevant. As
`highlighted in Figure 7, the 1-phenyl substituent makes
`aryl CH(cid:226)(cid:226)(cid:226)O contacts with the backbone oxygen of Glu54
`and the Tyr82 hydroxyl oxygen, and it has an amino-
`aromatic interaction with the (cid:15)-nitrogen of His87. While
`the interactions of the 1-phenyl group in 1 with Tyr82
`and Glu54 are found in most of the structures analyzed,
`the interaction with His87 occurs in only 29% of the
`analyzed structures. Average distances and frequencies
`
`

`
`3934 Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21
`
`Lamb and Jorgensen
`
`Table 2. Experimental Binding Free Energies, Calculated Free Energy Changes, and a Comparison of Experimental and Calculated
`Relative Binding Free Energiesa
`
`exptlb
`
`calcd
`exptlc
`AfB
`¢GA
`¢¢Gb
`¢¢Gb
`¢GFKBP
`¢Gaq
`¢GB
`11.69 ( 0.31
`10.25 ( 0.31
`-9.0, -9.5
`-10.6, -10.9
`1f2
`1.4
`1.6, 1.4
`-0.4, 0.3
`2.32 ( 0.08
`1.52 ( 0.09
`-9.4, -9.2
`-9.0, -9.5
`2f3
`0.8
`-1.0
`1.48 ( 0.07
`0.59 ( 0.08
`-11.1
`-10.1
`5f6
`0.9
`-1.7
`-1.1
`-10.07 ( 0.10
`-8.36 ( 0.09
`-10.1
`-9.0, -9.5
`2f5
`-1.9
`-8.76 ( 0.11
`-6.86 ( 0.11
`-9.9
`-10.6, -10.9
`1f4
`0.7
`a All free energies in kcal/mol. b Absolute binding free energies are derived from rotamase inhibition data given in Table 1, using ¢G
`) RT ln Ki and T ) 25 °C (298 K). c Relative binding free energies are only listed when experimental data from the same source may be
`compared.
`
`calcd
`
`tion, 2f3 and 5f6. Experimental data from Guilford
`Pharmaceuticals suggests that modification to a pyridyl
`substituent improves binding of 3-phenylpropyl com-
`pounds, although in one case this is contradicted by data
`obtained at SmithKline Beecham (Table 2). Considering
`the ca. 4 kcal/mol more favorable free energy of hydra-
`tion of pyridine than of benzene54 and assuming that
`the pyridyl compounds bind in a manner similar to 1
`in the solvent-exposed region of the binding pocket, the
`phenyl to pyridyl conversion would be expected to
`decrease binding affinity. Simply put, a dipole is better
`solvated in a medium with a higher average dielectric
`constant.
`First, the solvation of the ligands was addressed. The
`aqueous transformation of 2f3 resulted in a free energy
`difference of 1.52 ( 0.09 kcal/mol, somewhat larger than
`that for 5f6 (0.59 ( 0.08 kcal/mol). The gas-phase
`transformation of 5 to 6 yielded a free energy change of
`3.15 ( 0.04 kcal/mol, which combines for a net relative
`free energy of hydration of -2.6 kcal/mol favoring 6.
`This is reasonable given the previous benzene to pyri-
`dine results,54 and the larger, flexible ligand. Hydrogen
`bonds to water were similar for both sets of ligands; as
`expected, the pyridyl nitrogen provides an additional
`acceptor site in 3 and 6 (Table 4).
`The free energy profiles for both pairs of unbound and
`bound perturbations are displayed in Figure 6b. They
`are again notably smooth. As seen from Table 2, the
`transformations in the protein are ca. 1 kcal/mol less
`favorable than those in solution, and the net result is a
`consistent preference for the phenylpropyl ligands. No
`direct protein contacts are made by the pyridyl nitrogen
`atom; its interaction with water is maintained (Table
`4). For this hydrogen bond, the optimal interaction
`energy is -6.2 kcal/mol,54 much stronger than an aryl
`CH(cid:226)(cid:226)(cid:226)O- or N- interaction. However, the pyridyl
`nitrogen does participate in a hydrogen-bonding net-
`work with water molecules and carbonyl oxygens in the
`protein backbone. As reflected in Figure 9, a two-water
`bridge between the oxygen of Val55 and the nitrogen is
`found consistently in 6-FKBP12. In 3-FKBP12 the
`interaction is often mediated by three water molecules,
`with a few structures in which one molecule also bridges
`to the Gln53 carbonyl oxygen (Figure 9). The Val55
`oxygen has been noted previously as a consensus
`hydration site in FKBP12-ligand complexes.57 Crystal-
`lographic waters from the 1-FKBP12 structure13 were
`not explicitly included in the calculations, so it is
`gratifying that this interaction is established during the
`MC simulations. Water surrounding the ligand in the
`binding pocket can also bridge longer distances; for
`example, in 6-FKBP12, four molecules link the hy-
`
`Figure 7. Intermolecular aryl CH(cid:226)(cid:226)(cid:226)N, O contacts with His87,
`Tyr82, and Glu54 made in 1-FKBP12 that are lost on trans-
`formation to 2-FKBP12. One representative configuration
`from the Monte Carlo simulation is illustrated.
`
`of occurrence of these and other key interactions be-
`tween the ligands and FKBP12 are summarized in
`Table 3. Such aryl interactions are commonly observed
`in protein crystal structures, and their orientational
`distributions have been analyzed.55,56 To provide a
`sense of the strength of the aryl CH(cid:226)(cid:226)(cid:226)X interactions
`with the OPLS-AA force field, gas-phase optimizations
`for complexes of benzene with phenol, imidazole, N-
`methylacetamide, and water yielded the interaction
`energies shown in Figure 8. With the hydrogen bonds
`constrained to be linear, the intrinsic interaction ener-
`gies are in the 1-2 kcal/mol range.
`In both complexes, the 3-phenylpropyl moiety remains
`in the FK506-cyclohexyl region of FKBP12, and there
`is usually a water molecule well-positioned on the face
`of one or both phenyl rings in 1 (Figure 7). For
`reference, in the gas phase, the optimal (cid:240) hydrogen bond
`between benzene and a water molecule has an interac-
`tion energy of ca. -3.5 kcal/mol.42 An aromatic hydro-
`gen of the 3-phenylpropyl group often interacts with the
`carbonyl oxygen of Val55 in 1-FKBP12; this contact is
`shorter and more frequent in 2-FKBP12 (Table 3). In
`addition, while the 1-phenyl ring no longer interacts
`with the Glu54 backbone carbonyl oxygen when the
`transformation to 2 is complete, the 3-phenyl ring shifts
`to pick up this contact. Considering these observations,
`it is likely that the major contribution to the weaker
`binding for 2 than 1 is the overall reduction in hydro-
`phobic interactions and, possibly, the specific loss of the
`aryl CH(cid:226)(cid:226)(cid:226)O, N contacts between the 1-phenyl group and
`Tyr82 and His87.
`Effect of the 3-Phenyl to 3-(3-Pyridyl) Substitu-
`
`

`
`Neurotrophic Inhibitors of FK506 Binding Protein
`
`Journal of Medicinal Chemistry, 1998, Vol. 41, No. 21 3935
`
`Table 3. Key Intermolecular Hydrogen Bond Distances (Frequencies %) Less than 3.2 Åa
`
`1f2b
`
`2f3c
`
`ligand
`FKBP12
`O4
`Tyr26H(cid:15)
`Tyr26HŁ
`O4
`Phe36H(cid:15)
`O3
`Phe36H(cid:15)
`O4
`Gln53O
`H22
`Glu54O
`H23
`Glu54O
`H25
`Val55O
`H21
`Val55O
`H22
`Val55O
`H23
`Ile56H
`O2
`H19
`Tyr82C(cid:15)
`Tyr82H(cid:15)
`C19
`Tyr82H(cid:15)
`O3
`Tyr82HŁ
`O3
`Tyr82HŁ
`O1
`Tyr82OŁ
`H29
`His87N(cid:15)
`H28
`Phe99Hœ
`2.9 (100)
`O3
`Phe99Hœ
`2.9 (22)
`O4
`Phe99H(cid:15)
`3.0 (46)
`2.9 (72)
`2.9 (22)
`O4
`a Only interactions found in >10% of saved structures are reported. The frequency in parentheses records the percentage of the analyzed
`structures that had the feature. b 1f2 for 45 structures saved every 2 (cid:2) 105 configurations. c 2f3 and 5f6 for 50 structures saved every
`2 (cid:2) 105 configurations.
`
`5f6c
`
`2.8 (92)
`3.1 (16)
`
`2.6 (100)
`2.8 (96)
`2.8 (100)
`
`3.1 (26)
`3.0 (40)
`
`2.0 (100)
`3.0 (20)
`
`3.0 (60)
`1.8 (100)
`
`2.8 (96)
`3.0 (42)
`
`2.6 (98)
`
`2.9 (68)
`
`2.9 (40)
`2.0 (100)
`3.0 (76)
`3.0 (28)
`3.1 (52)
`1.8 (100)
`3.0 (74)
`
`2.7 (100)
`
`3.0 (82)
`2.8 (96)
`3.0 (53)
`2.8 (93)
`
`2.8 (91)
`
`2.9 (60)
`2.0 (100)
`3.0 (69)
`3.1 (42)
`2.9 (93)
`1.8 (100)
`3.1 (31)
`2.8 (93)
`2.9 (29)
`2.9 (78)
`
`2.8 (96)
`2.9 (62)
`3.0 (16)
`2.7 (96)
`
`2.8 (84)
`
`2.7 (96)
`
`2.0 (100)
`3.0 (49)
`3.0 (69)
`3.0 (69)
`1.8 (100)
`3.1 (44)
`
`2.7 (98)
`3.0 (62)
`3.0 (56)
`2.7 (96)
`
`2.8 (88)
`
`2.8 (84)
`3.0 (34)
`2.0 (100)
`3.1 (16)
`3.1 (38)
`3.0 (70)
`1.8 (100)
`3.0 (60)
`
`2.9 (94)
`2.9 (44)
`3.0 (22)
`2.6 (98)
`
`2.8 (86)
`
`3.1 (10)
`2.1 (100)
`3.0 (52)
`3.1 (40)
`3.0 (92)
`1.8 (100)
`3.0 (98)
`
`2.7 (96)
`
`2.6 (100)
`
`2.8 (92)
`
`droxyl oxygen of Tyr26 to the backbone carbonyl of Glu54
`(not shown).
`Finally, to ensure that the phenylpropyl 3-position
`selected for transformation did not influence possible
`protein-ligand contacts or the computed free energy
`difference, the final FEP window of 5f6 bound to
`FKBP12 was also started with the 3-pyridyl ring flipped
`by 180°. The free energy change calculated in this
`window was unaffected, and no new contacts with the
`protein were observed. In all, the simulations support
`the intuitive idea that binding for 3-(3-pyridyl)propyl
`compounds should be less favorable than that for
`3-phenylpropyl ligands. This finding agrees qualita-
`tively with the SmithKline observations for 3 and 2, but
`not with either pair of binding affinities (3 vs 2 or 6 vs
`5) reported by Guilford Pharmaceuticals, as sum-
`marized in Table 2.
`Effect of Ring Contraction on Binding, 2f5 and
`1f4. With the relatively hydrophobic binding pocket
`of FKBP12, one might expect the larger pipecolyl
`compounds to bind with higher affinity than prolyl
`homologues. Though this is true for most published
`FKBP12 inhibitors,2 the opposite pattern is reported for
`the 3-phenylpropyl and 3-(3-pyridyl)propyl compounds
`in Table 1.
`With this in mind, the 2f5 transforma

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