SlideShare a Scribd company logo
1 of 6
Download to read offline
Ligand- and mutation-induced conformational
selection in the CCR5 chemokine G
protein-coupled receptor
Ravinder Abrola,1,2,3
, Bartosz Trzaskowskia,1,4
, William A. Goddard IIIa,1,2
, Alexandre Nesterovb
, Ivan Olaveb
,
and Christopher Ironsb
a
Materials and Process Simulation Center, California Institute of Technology, Pasadena, CA 91125; and b
PharmSelex Inc., Seattle, WA 98102
Contributed by William A. Goddard III, July 27, 2014 (sent for review January 25, 2014)
We predicted the structural basis for pleiotropic signaling of the C-C
chemokine type 5 (CCR5) G protein-coupled receptor (GPCR) by
predicting the binding of several ligands to the lower-energy
conformations of the CCR5 receptor and 11 mutants. For each case,
we predicted the ∼20 most stable conformations for the receptor
along with the binding sites for four anti-HIV ligands. We found that
none of the ligands bind to the lowest-energy apo-receptor confor-
mation. The three ligands with a similar pharmacophore (Maraviroc,
PF-232798, and Aplaviroc) bind to a specific higher-energy receptor
conformation whereas TAK-779 (with a different pharmacophore)
binds to a different high-energy conformation. This result is in
agreement with the very different binding-site profiles for these
ligands obtained by us and others. The predicted Maraviroc bind-
ing site agrees with the recent structure of CCR5 receptor cocrystal-
lized with Maraviroc. We performed 11 site-directed mutagenesis
experiments to validate the predicted binding sites. Here, we inde-
pendently predicted the lowest 10 mutant protein conformations
for each of the 11 mutants and then docked the ligands to these
lowest conformations. We found the predicted binding energies to
be in excellent agreement with our mutagenesis experiments. These
results show that, for GPCRs, each ligand can stabilize a different
protein conformation, complicating the use of cocrystallized struc-
tures for ligand screening. Moreover, these results show that a sin-
gle-point mutation in a GPCR can dramatically alter the available
low-energy conformations, which in turn alters the binding site,
potentially altering downstream signaling events. These studies val-
idate the conformational selection paradigm for the pleiotropic
function and structural plasticity of GPCRs.
protein structure prediction | ligand–protein binding prediction |
conformational ensemble | functional selectivity | GEnSeMBLE
Since the finding that individuals lacking the C-C chemokine
receptor type 5 (CCR5) gene are resistant to HIV, the CCR5
receptor has been established as a coreceptor for macrophage-
tropic viruses, including HIV, to enter host cells (1). Several
drugs aimed at disrupting the HIV–CCR5 coupling have been
developed. To understand the structural mechanism of ligand
binding to G protein-coupled receptors (GPCRs) and how these
CCR5 targeting drugs work, we predicted the low-energy struc-
tures of the apo CCR5 receptor, which in turn were used to
predict the binding sites for various ligands. The focus was on
ligands already known to inhibit CCR5: Maraviroc (MVC) (2),
PF-232798 (PF) (3), Aplaviroc (APL) (4), and TAK-779 (TAK) (5)
because there are abundant experimental data on their binding to
numerous CCR5 mutants, to test the predicted structures. The li-
gand structures are shown in Scheme S1. For three ligands (MVC,
PF, and APL), we predict a similar binding site pharmacophore
centered on a protonated nitrogen, whereas the other ligand (TAK)
containing a quaternary nitrogen leads to a very different binding
site pharmacophore. To validate our predictions, we identified 11
singly mutated CCR5 receptor forms selected to test the predicted
binding region for the ligands. The predicted structures of ligand-
bound CCR5 complexes presented here have already been used in
two recent studies to provide a structural basis for (i) biophysical
measurements (6) mapping the binding mode of MVC to CCR5
using genetically encoded photo–cross-linkers and (ii) viral entry
experiments (7) probing the basis for G protein-coupled and
-uncoupled CCR5 receptors by MVC-resistant and -sensitive HIV-
1 viruses.
While this work was being finalized for submission, the CCR5
receptor was crystallized with MVC (8), allowing a direct struc-
tural validation of the predicted structures as discussed in
Prediction of Ligand–CCR5 Structures and Comparison with the
Crystal Structure.
Results
Prediction of the CCR5 Structure. To predict the ensemble of
low-energy structures (conformations) for CCR5, we used the
GEnSeMBLE method (9), which performs a complete conforma-
tional sampling (∼11 billion) of transmembrane (TM) helix confor-
mations accessible in the membrane. This method starts with
templates for positioning the seven TM helices based on
available experimental/predicted structures and then samples
a large number of helix orientation angles as defined in Fig. 1.
During conformational sampling, we first sampled only the
helix rotation angles (η) (on a 15° grid), and the BiHelix energy
Significance
The C-C chemokine receptor type 5 (CCR5) G protein-coupled re-
ceptor (GPCR) is a prime target for preventing HIV invasion. A
major difficulty in developing effective therapeutics is that the
CCR5 exhibits an ensemble of ∼10–20 distinct low-energy con-
formations, each of which might favor binding to different
ligands and/or lead to different downstream functions. X-ray
structures generally provide only one of these conformations. We
applied the GEnSeMBLE methodology to predict this ensemble,
and we designed and carried out 11 experiments to validate the
ability of this ensemble to predict binding of an HIV therapeutic
to CCR5. We found that each of the mutations changes the
binding site. The predicted effects of mutations on binding are in
excellent agreement with experiments, providing CCR5 structures
for designing new ligands.
Author contributions: R.A. and W.A.G. designed research; R.A., B.T., W.A.G., A.N., I.O.,
and C.I. performed research; R.A., B.T., W.A.G., A.N., I.O., and C.I. analyzed data; and R.A.,
B.T., and W.A.G. wrote the paper.
The authors declare no conflict of interest.
1
R.A., B.T., and W.A.G. contributed equally to this work.
2
To whom correspondence may be addressed. Email: wag@wag.caltech.edu or abrolr@
csmc.edu.
3
Present address: Departments of Medicine and Biomedical Sciences, Cedars-Sinai Medical
Center, Los Angeles, CA 90048.
4
Present address: Centre of New Technologies, University of Warsaw, 02-089, Warsaw,
Poland.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1413216111/-/DCSupplemental.
13040–13045 | PNAS | September 9, 2014 | vol. 111 | no. 36 www.pnas.org/cgi/doi/10.1073/pnas.1413216111
(10) was used to evaluate the energy of all (24)7
∼ 4.6 billion TM
helix bundle conformations. For the human CCR5 receptor
structure prediction, this sampling was done starting with each of
the following four X-ray–derived templates: bovine rhodopsin
(PDB ID code 1u19) (11), human β2 adrenergic receptor (PDB
ID code 2rh1) (12), human adenosine A2A receptor (PDB ID
code 3eml) (13), and turkey β1 adrenergic receptor (PDB ID
code 2vt4) (14). The full seven-helix bundle was built for the top
1,000 conformations from each template, and the side chains
were optimized using the SideChain Rotamer Energy Analysis
Method (SCREAM) side-chain placement protocol (15), fol-
lowed by minimization (10 steps using Dreiding force field) (16).
The β2 receptor template-based conformational ensemble led to
the lowest energies (most stable conformations). We then se-
lected the most stable 16 CCR5 conformations, for each of which
we simultaneously sampled all three helix orientation angles (θ, ϕ,
and η) allowing −10°, 0°, and +10° for the θ tilt angle and the −15°,
0° , and +15° range for both ϕ azimuthal and η rotation angles.
This procedure led to a total of (27)7
∼11 billion TM bundle
conformations, for each of which we evaluated the energy rapidly
using the SuperBiHelix method (17). Then the lowest-energy
2,000 conformations were built into seven-helix bundles and op-
timized, from which we selected the 20 lowest-energy con-
formations (labeled WT1 to WT20) as shown in Table S1. From
these conformations, we selected eight structurally diverse seven-
helix structures (highlighted rows in Table S1) for further analysis
and ligand docking. The lowest-energy conformation WT1 cor-
responds to the predicted apo conformation of the receptor. None
of the experimentally obtained GPCR structures (with the ex-
ception of opsin) has been ligand-free so it remains to be con-
firmed whether WT1 resembles the apo conformation of the
CCR5 receptor.
These best eight diverse structures were then used to predict
the binding of CCR5 ligands. The next section shows that MVC
binds most strongly to the WT7 conformation rather than the
lowest-energy WT1 conformation.
Prediction of Ligand–CCR5 Structures and Comparison with the
Crystal Structure. The four ligands (MVC, PF, APL, and TAK)
were minimized using the B3LYP flavor of density functional
theory (DFT) (with the 6–311G** basis set) using the Jaguar
software package (Jaguar, version 7.8; Schrödinger, LLC). A
conformational search was performed over the rotatable bonds
for each ligand, and ∼20–30 conformations were selected (based
on energy and diversity) (SI Materials and Methods) for each li-
gand. Then, each of these conformations was docked to the
previously selected eight CCR5 conformations using the
DarwinDock/GenDock method (18). This method for predicting
ligand–protein structures samples a complete set (∼50,000) of
ligand poses in the potential binding regions, which are then
clustered based on root-mean-squared (rms) distances to obtain
∼2,000 family heads whose energies are then evaluated to select
∼200 families, for each of which the energies of all members are
evaluated. The top 100 poses from this process are optimized to
finally select the best protein–ligand structure. These ligand–
CCR5 structure predictions were performed independently for
each of the eight CCR5 structures using the DREIDING force
field (16).
The final predicted most stable structure for the MVC/CCR5
complex is shown in Fig. 2, overlaid with the complex structure
from the recent crystal structure (8) and showing the most im-
portant ligand–residue interactions. Binding of MVC stabilized
the WT7 structure of CCR5, which was also stabilized by PF and
APL (Fig. 3 A and B) whereas TAK selected the WT10 confor-
mation (Fig. 3C). Similar results have been obtained earlier for the
adenosine A3AR receptor, where all four selective agonists sta-
bilized WT15 whereas all four selective antagonists stabilized WT2
or WT3 (19) conformations; for none of these eight ligands was
the complex with WT1 of lowest energy. These results emphasize
the importance of having the ensemble of low-energy structures,
and they suggest that ligands can control the final conformation
and perhaps downstream function of CCR5 as well as GPCRs
in general.
The biggest changes in the CCR5 structure upon binding
MVC (WT1 to WT7) were in the η rotation angles of helices 2
and 7 (compare the binding site residues in a 2D representation
in Fig. S1 A and D). The 20-degree rotation of helix 2 completely
changes the interhelical hydrogen bond network of CCR5. In the
wild-type protein, the structure of CCR5 is stabilized by several
interactions, including a conserved N48 (TM1)–D76 (TM2) hy-
drogen bond, S63 (TM2)–R126 (TM3) hydrogen bond, and fi-
nally a 1-2-7 hydrogen bond network Y37 (TM1)–W86 (TM2)–
T284 (TM7). The 1-2-7 network and the W86–T284 hydrogen
bond, in particular, result in a compact structure for the protein
that is not suitable to develop strong interactions with large
ligands having a tertiary nitrogen atom, such as MVC. This is
because the interaction with T284 forces W86 to be close to
E283, the crucial residue for interacting with the basic part of the
ligand. To overcome this problem, the W86-T284 hydrogen bond
must break, which leads to a slight destabilization of the struc-
ture, which forces helices 2 and 7 to rotate to find favorable
interactions with other residues from other helices, such as π–π
stacking with Y108. The existence of a similar receptor activation
network was proposed recently (20), but it did not include T284;
Fig. 1. Coordinate frame for the TM helix orientation in a membrane
(reprinted with permission from ref. 28). The x-y plane defined as the middle
plane of the membrane bilayer. The helical axis, defined as the least moment
of inertia axis, intersects this plane at position x,y with residue “h.” The
helical axis is tilted by an angle θ with respect to the z axis that is perpen-
dicular to the membrane. The azimuthal angle ϕ of this tilt in the x-y plane and
the rotation angle of the helix around this axis is the angle η.
Fig. 2. Comparison of the predicted MVC binding site (MVC in yellow and
binding site residues in cyan) with the experimental X-ray MVC binding site
(MVC in purple and binding site residues in green).
Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13041
BIOCHEMISTRY
instead, the network consisting of Y37, W86, Y108, and E283
residues was proposed.
Thus, binding of MVC causes helix 2 to rotate by 20 degrees,
which opens up the binding pocket formed by helices 1, 2, and 7.
This rotation also allows the W86 residue to find an interaction
with Y108 on helix 3, while interacting with the ligand, whereas
T284 still forms a hydrogen bond with Y37 on TM1. Addition-
ally these rotations aid the hydrogen bonding between N71
(TM2)–N293 (TM7) due to a more favorable geometry of these two
residues. Moreover, the 15-degree rotation of H7 allows the E283
residue to find a better orientation, optimizing the interaction be-
tween the carboxylic group and the basic nitrogen from the tropane
ring of MVC, as well as the hydroxyl group of Y251 on helix 6.
The binding pocket of the predicted complex model is in very
good agreement with the crystal structure of the MVC:CCR5
complex published just recently (8). Fig. 2 also shows the MVC–
CCR5 interactions observed in the crystal structure of the
complex for comparison with the predicted structure. Fig. S1 A
and B shows these interactions using a 2D representation. All
important interactions between the MVC and CCR5 residues
were predicted, including (i) the salt bridge between the tropane
nitrogen atom and Glu283, (ii) the hydrogen bond between the
carboxyamide nitrogen and Tyr251, and (iii) the hydrophobic
interactions of the phenyl group with Tyr108, Phe109, and
Phe112. The predicted structure is not perfect and missed some
interactions: (i) the hydrogen bond between the amine moiety of
the triazole group and Tyr37 and hydrogen bonds between one
of the fluorines on the cyclohexane ring with Thr195 and Thr259.
The accuracy of the predicted CCR5 structure (obtained using
the human β2 adrenergic receptor template) shows the strength
of the BiHelix and SuperBihelix methodology. This approach
allows sampling of a complete set of TM rotations and local tilts,
going well beyond simple homology modeling. This method also
allows the residues crucial for ligand binding to be correctly
positioned in the binding pocket, leading to high-quality protein
structures starting from relatively low-homology templates (25%
sequence identity between β2AR and CCR5 in the TM region).
The MVC ligand does not select the lowest-energy receptor
conformation (WT1) for binding but a higher-energy one (WT7).
The helix rotation η angles change from [0, −120, 0, 0, 15, 0, 0]
for the WT1 conformation to [15, 0, 0, 105, 15, 0, 0] for the WT7
conformation, which is very close to the [15, 0, 10, 105, −15, 0, 0]
angles calculated for the crystal conformation of CCR5 bound to
MVC (8). The predicted TM domains are also in good agree-
ment with the crystal structure (crystal TMs/predicted TMs):
TM1, 26–57/26–57; TM2, 64–92/64–93; TM3, 97–132/98–131;
TM4, 142–166/142–166; TM5, 187–222/192–224; TM6, 229–261/
229–259; and TM7, 269–297/276–299.
The predicted PF binding mode (Fig. 3A) is similar to that of
MVC, based also on the same protein configuration WT7. This
result is reasonable because the ligand structures are rather similar.
Thus, the mechanism of PF binding to CCR5 protein is similar to
that of MVC binding, involving also the rotation of helices 2 and 7,
which opens up the 1-2-7 binding pocket to allow the ligand to bind
to E283. Fig. S1C shows the PF binding site in a 2D representation.
Although APL has a slightly different molecular scaffold,
it shares with MVC and PF a strongly basic nitrogen atom
positioned in the center of the molecule (see central N atom in
the ligand structures shown in Scheme S1). Indeed, it interacts
strongly with the E283 anchor point (Fig. 3B), similar to the
MVC and PF cases. Thus, the preferred CCR5 configuration is
WT7, just as for MVC and PF. However, the benzoic acid group
of APL provides an additional anchoring point for the ligand, far
from helix 7 and E283. It interacts with the K191 residue located
on the verge of helix 5 and extracellular loop 2, leading to a large
reduction of binding upon K191A mutation. This result is con-
sistent with the experimental >100-fold decrease in binding af-
finity of APL upon K191A mutation (21, 22).
The mechanism of binding of APL also involves breaking or
weakening the 1-2-7 hydrogen bond network (Y37-W86-T284),
with E283 providing the anchor for the middle part of the ligand.
Interestingly, the benzoic acid group also interacts with R168 on
extracellular loop 2, also shown to be important for ligand
binding in experimental assays (23).
The TAK antagonist is based on a completely different scaf-
fold containing a quaternary amine group in place of the pro-
tonated nitrogen for the other three ligands, and giving the
whole ligand a formal charge of +1. We predict that TAK prefers
to bind to the WT10 conformation (Fig. 3C). The overall
mechanism of opening the binding site via disruption of the
W86-T284 hydrogen bond is preserved. As in the previous cases,
the 1–2-7 hydrogen bond network is broken or weakened due to
the 20-degree rotation of helix 2 and formation of a π-π stacking
interaction between W86 and Y108. The T284-Y37 hydrogen
bond stabilizes the ligand-bound structure, in agreement with the
experimental data that show lower binding affinity of TAK upon
the Y37A mutation (21). Additionally, the 15-degree rotation of
helix 5 allows the T195 residue to rotate away from the toluyl
group of the ligand, making more space available in the 3-5-6
pocket. On the other hand, the relatively large rotation of weakly
interacting H4 does not seem to have any impact on the structure
of the protein or ligand binding.
A detailed analysis of the TAK binding site shows that W86 is
the strongest interacting residue with this ligand, with a −3.7 kcal/mol
contribution to the overall interaction energy, which agrees
well with the experimental data. This favorable interaction comes
from an imperfect π–π stacking between the tryptophan residue
and the benzo[7]annulene part of the ligand. We predict that
the contribution of E283 residue is the second strongest one
(−3.3 kcal/mol). On the other hand, the interaction ofthe ligand with
Y108 is very weak so that the relatively large effect of the Y108A
mutation seen experimentally results because WT22 becomes the
most stable TAK bound conformation in the Y108A receptor
mutant. Thus, the large effect of the mutation is indirect through
stabilizing a different protein conformation rather than a change
in direct interaction with the ligand. All other residues in-
vestigated experimentally show very weak interactions with TAK,
consistent with the experimental mutation data.
Mutation Experiments. Based on the binding structure for MVC
to WT CCR5, we selected 15 single mutants to validate the
predicted structure: F109Y, F112A, Q194A, Y251F, D276A,
Q277A, W86A, A90H, T105A, Y108A, F109A, I198A, Y251A,
Q280A, and E283A. For 11 of these mutants, binding studies of
Fig. 3. The predicted binding sites for (A) PF-232798 (PF), (B) Aplaviroc (APL), and (C) TAK-779 (TAK).
13042 | www.pnas.org/cgi/doi/10.1073/pnas.1413216111 Abrol et al.
MVC and PF were carried out. The predicted results were
dramatically inconsistent with experiments for the A90H and
F109A mutants. Also, the experiments suggested a bigger
change in binding energy for PF vs. MVC in the A90H mutant.
A fundamental assumption in our selection of mutations was
that the structure of the binding site would not change signifi-
cantly (the established paradigm for assessing the effect of
mutations). These experimental results suggested that this as-
sumption might be invalid for CCR5 and potentially other
GPCRs. This result is also consistent with our earlier studies on
binding of the ligand BX471 to CCR1 receptor, where we found
that a single mutation of human CCR1 to mouse and rat CCR1
dramatically changed the binding (24).
Rather than starting all over to predict the hierarchy of CCR5
structures for each of the 15 mutant receptors, we simplified the
procedure as follows. We went back to the 100 lowest-energy
structures from the conformational hierarchy predicted for WT
CCR5; we mutated the single residue in question for each of
those structures, and then we reassigned side chains, using
SCREAM (15), and minimized the structures. The net result is
often a dramatic reordering in the ranking of these 100 receptor
structures. We then selected the 20 lowest-energy structures for
docking the four ligands.
This approach leads to dramatically different hierarchies, as
shown in Table S2. For example, the W86A mutant makes the
WT30 conformation fourth best. This outcome is easy to un-
derstand from the structures. WT30 conformation has the W86
residue up against TM7 (Fig. S2), but replacing this residue by
Ala in the W86A mutant removes this steric clash, dramatically
improving the rank of the WT30 conformation to fourth.
Next, we docked the 20–30 conformations for each of the four
ligands to each of the ∼10 receptor conformations for each of the
mutated proteins using the DarwinDock/GenDock methods
(18). The results are shown in Table 1. For the top-scoring poses,
a water displacement penalty was applied if a polar residue was
mutated to a nonpolar residue and could have accommodated a
water molecule in the apo protein conformation. This penalty
affected only the Q194A mutant.
The results in Table 1 show the pronounced effect of muta-
tions in the active site. Thus, the WT7 configuration stabilized by
MVC binding to WT CCR5 is not the best binding configuration
for any of the 11 mutants, all of which were aimed at modifying the
binding! Instead, WT5 conformation appears seven times, WT2
four times, WT3 three times, and WT10 once. These mutants were
generated experimentally (as described in SI Materials and Methods,
Experimental Generation of CCR5 Mutants) and used to measure Ki
values for the binding of different ligands to CCR5 WT and mutant
receptors (SI Materials and Methods, Ligand-Binding Experiments).
Comparing relative Ki values to relative predicted binding energies
for each case led to the results in Fig. 4, which show excellent
agreement with experimentally obtained Ki values for MVC binding
to the mutants relative to the WT receptors, with the binding
constants varying over a range larger than 3 logs. The trend for the
other three ligands is also correct, but the correlation is weaker than
that for MVC, due in part to the fact that the experimental changes
in binding upon mutations span less than three orders of magnitude
(less than 3 logs) for these ligands, which is only slightly larger than
the best expected resolution from the calculations of ∼2 logs [Fig.
S3 (PF), Fig. S4 (APL), and Fig. S5 (TAK)]. These results overall
confirm the prediction that different conformations can be stabi-
lized under different conditions (mutants and/or ligands).
Discussion
Analysis of the Ligand–CCR5 Binding Site in the Context of Mutations
and Ligand Structure–Activity Relationship Series. The experimental
assays confirmed the predictions that the binding modes of MVC
and PF232798 are rather similar, but with one exception. The
W86A mutation gives a 629-fold decrease in binding of PF and
a 29.8-fold decrease in binding for MVC. [Kondru et al. reported
a 10-fold decrease in MVC binding upon W86A mutation (22) in
slightly different conditions.] Indeed, the calculations find that
for W86A, both MVC and PF bind to WT7, but with predicted
energy changes of 3.1 kcal/mol and 3.5 kcal/mol.
The difference in impact of the W86A mutation on MVC/PF
binding comes in part from structural differences between those
two ligands. Although they share a similar scaffold, they have
a different stereochemistry for the five-membered nitrogen-rich
ring that forces the boat conformation of the tropane ring for PF
whereas MVC retains the favorable chair conformation. Also
there is an additional cyclohexane ring in the terminal part of
MVC that affects the pKa value of the N3 atom, making it more
basic. The quantum mechanical calculations predict the pKa
value of the nitrogen in the five-membered ring to be 5.0–6.0 for
MVC, but 6.0–7.0 for PF, suggesting that MVC may exist in
a doubly protonated state in the experimental environment (the
tropane N atom is very basic for both MVC and PF, with a pre-
dicted pKa of ∼9). Because, for WT, the residue W86 is close to
the five-membered ring of both MVC and PF, this change in the
protonation state could have been responsible for the predicted
different interaction with tryptophan through a strong cation–π
interaction (25).
Table 1. Predicted lowest-energy protein structures for wild-
type and mutant receptors in the apo and ligand-bound forms
Mutations
Lowest-energy CCR5 conformations
Apo MVC PF APL TAK
Wild-type wt1 wt7 wt7 wt7 wt10
F109Y wt3 wt5 wt5 wt5 wt3
F112A wt1 wt5 wt5 wt5 wt3
Q194A wt1 wt5 wt5 wt5 wt4
Y251F wt1 wt2 wt22 wt4 wt3
D276A wt1 wt2 wt7 wt3 wt2
Q277A wt1 wt5 wt5 wt5 wt3
W86A wt1 wt7 wt7 wt7 wt3
A90H wt3 wt10 wt5 wt10 wt3
T105A wt1 wt2 wt6 wt6 wt3
Y108A wt3 wt2 wt22 wt5 wt22
F109A wt1 wt5 wt7 wt7 wt3
I198A wt1 wt5 wt5 wt5 wt25
Y251A wt2 wt3 wt5 wt5 wt3
Q280A wt3 wt5 wt7 wt7 wt7
E283A wt3 wt3 wt22 wt22 wt3
Ligands used are Maraviroc (MVC), PF-232798 (PF), Aplaviroc (APL), and
Tak-779 (TAK). This shows that a single mutation can dramatically affect the
binding site.
Fig. 4. Comparison of the predicted effect of Maraviroc binding to mutants
vs. experimental data for the 11 cases studied experimentally. The red line is
the fit (R2
= 0.89) whereas the green lines indicate a deviation of 1 kcal/mol.
Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13043
BIOCHEMISTRY
However, our predicted structure makes this hypothesis un-
likely because W86 is oriented in a way not allowing for the
π–cation interaction in the PF. Despite a similar orientation of
W86 in the MVC-bound and PF-bound models, we predict
a difference in the interaction energy between those two ligands
and the tryptophan residue. The binding site energy analysis
allows the ligand–protein interaction to be partitioned into
contributions from each residue. For the MVC binding site, we
found a ligand interaction energy of −2.0 kcal/mol with W86
(which is less than the E283 contribution of −3.2 kcal/mol) whereas
for PF it is −4.2 kcal/mol (almost identical to −4.3 kcal/mol
contribution of E283). This difference can be understood in
terms of the geometries of the ligands (the different stereo-
chemistry of the tropane ring-five membered ring connections)
and the different position of the phenyl ring with respect to the
tropane ring for MVC and PF, leading to different interaction
energies.
A second important difference in the experimental mutation
data for these two ligands occurs for the A90H mutation, which
changes the PF binding significantly, but has almost no effect on
MVC binding. PF prefers the WT5 conformation for the A90H
mutant whereas MVC prefers WT10. Even though both MVC
and PF are anchored by E283 and occupy a similar position in
the protein, they are structurally different in the part that
interacts with helix 2. The bulkier pyridyl group of PF positions
the ligand closer to the A90 residue, which does not affect the
binding unless this residue is mutated to a larger moiety: i.e., the
histidine residue. Indeed, MVC is relatively far from the A90
residue (which is on top of helix 2); as a result, replacing A90 by
the bulky histidine does not affect its binding. However, the li-
gand PF is positioned closer to A90 so that the mutation to
histidine causes a steric clash, decreasing binding. This difference
arises from the difference in the receptor conformation for the
A90H mutant preferred by MVC (WT10) versus PF (WT5).
On the other hand, all 11 mutations investigated force the
TAK-bound CCR5 system to switch to a different conformation.
This conformational change has, however, little impact on the
computational mutation results. For all six mutations studied
experimentally for TAK, the results from both structures were in
good agreement with experimental data.
Unfortunately, the structure–activity relationship (SAR) series
for the PF compound showed (3) little structural diversity. The
experimental data presented in ref. 3 agrees well with the
PF232798 binding pose predicted here. All 3-Substituted 1,4,6,7-
Tetrahydro-imidazo[4,5-c]pyridines that were tested experimentally
(compounds 39 and 41a–h from ref. 3) show a very high affinity, on
the level of the PF232798 system. In the predicted docking pose,
this part of the ligand interacts with the flexible N terminus of the
protein and there is enough space to incorporate any substituent
suggested in the original work. The modification of the ligand
yielding 3-substituted derivatives (compounds 20a–h and 29–32
from ref. 3) positions the terminal part of the ligand closer to
TM2, but avoiding steric clashes with any protein residues. This
short SAR series further supports the predicted PF232798
binding pose.
APL (Aplaviroc) has been developed from a series of spi-
rodiketopiperazine derivatives showing a relatively high CCR5
activity (5). In that study (5), four different spirodiketopiperazine
derivatives were evaluated for anti–HIV-1 activity, and molecule
E913 was selected as the lead compound for a newer round of
optimizations. Our predicted docking pose of APL is in agree-
ment with the experimental data for this series of compounds.
The main difference among the earlier spirodiketopiperazine
derivatives E910, E913, E916, and E917 is the length of the
molecule’s terminal part targeting the TM3-TM4-TM5 region of
CCR5. According to the predicted pose, there is enough space to
incorporate even the longest, heptylbenzene tail present in E910.
This region is also highly hydrophobic, with a number of residues
(F109, F112, F168, and I198) able to make relatively strong
interactions with the phenyl ring present at the end of the hy-
drophobic tail of all these ligands. The difference responsible for
the strong binding of APL is the addition of the benzoic acid
group to the end of this tail, which is the most important dif-
ference between APL and E913. In APL, the phenoxy linker is
able to position the new carboxylic group to promote a strong
interaction with K191, known to be important in APL binding
from the experimental mutation data.
For TAK-779, the original paper (4) does not disclose any
SAR series. Other Takeda compounds with known CCR5 ac-
tivity (like TAK-220) are completely different.
Implications for the Conformational Ensemble View of GPCRs. Fig. 5
summarizes the results for all ligands and mutants. Fig. 5A shows
that MVC, PF, and APL prefer to bind to the WT7 receptor
conformation, not to the lowest-energy WT1 conformation of
the apo-protein. TAK prefers to bind to the WT10 conforma-
tion. This difference between TAK and other ligands arises be-
cause TAK has a quaternary nitrogen group in place of the
tertiary nitrogen of the other ligands. Fig. 5B shows how differ-
ent mutants of the receptor stabilize distinct conformations and
indicates which conformation is preferred by ligands for different
mutants. These changes in conformation of the mutated apo-
protein explain the differential mutational data for W86A and
A90H mutants on binding of MVC vs. PF compounds. This re-
sult has direct implications for the functional selectivity of
GPCRs because it shows that different receptor conformations
are stabilized under different conditions (ligand binding and/or
Fig. 5. (A) Conformational shuffling upon binding of different ligands. (B) CCR5 conformations preferred by different ligands and mutants.
13044 | www.pnas.org/cgi/doi/10.1073/pnas.1413216111 Abrol et al.
mutations), which potentially can have distinct downstream
effects on signaling.
These results show that CCR5 is a relatively plastic protein
with distinctly different protein conformations stabilized by dif-
ferent ligands. A single mutation can considerably alter the en-
ergetics of the low-lying conformations to alter the preferred
binding site. This outcome reinforces the conformational selec-
tion basis of the GPCR pleiotropic function. These results sug-
gest that the paradigm of designing a ligand for a particular
protein structure will not be sufficient for developing new ther-
apeutics to CCR5 (and likely other GPCRs). One must consider
all of the low-energy protein conformations in the design and in
determining structure–activity relationships. Moreover, one must
be cautious about interpreting mutation data. Thus, we consider
that methods presented in this work for predicting the effect of
mutations on GPCR structure will be essential to developing
highly selective and active ligands targeting these receptors.
Conclusions
We developed an ensemble-based method that samples billions
of conformations to eventually select an ensemble of ∼20 low-
energy protein conformations for the CCR5 receptor and
showed how this conformational ensemble changes upon binding
of four ligands and how it changes because of mutations. To
assess the validity of the binding sites, 15 mutations were pre-
dicted in CCR5 that should significantly modify the binding.
Experimental binding studies were carried out for 11 of these
mutants, and several cases were found that disagreed signifi-
cantly from the expectations based on the predicted pose for
binding of the ligand MVC to the WT protein. This result mo-
tivated us to predict the ensemble of low-energy protein struc-
tures for each of the 15 mutants, some of which showed major
changes in the conformations available for ligand binding. We
predicted the optimum ligand pose for the ensemble of low-en-
ergy CCR5 conformations for each of 11 mutants and compared
with our experimental binding results. We found that ligand
binding to the predicting mutated receptor conformations agrees
well with the experimental mutation data for all four ligands.
This work leads to several important general conclusions. (i) Dif-
ferent ligands can stabilize dramatically different seven-helix bundle
conformations, complicating the use of experimental structures for
designing new strongly interacting ligands. (ii) For CCR5 (and
likely other GPCRs), even a single mutation can dramatically
change the binding site, complicating experimental validation of
ligand-binding structures. This result invalidates the standard par-
adigm of using directly the GPCR structure bound to the ligand to
explain mutation-binding data. (iii) The changes in protein con-
formation stabilized by different ligands and/or mutants might lead
to changes in functionality (a structural basis for GPCR pleiotropy).
Indeed functional differences between agonists and inverse
agonists (or antagonists) to GPCRs are well established. For ex-
ample, recent studies show that single, double, and triple mutant
forms of the CB1 receptor exhibit dramatic changes in function,
switching from full inactivity to high constitutive activity and back
for sequential mutations (26, 27). (iv) The GEnSeMBLE approach
for predicting the ensemble of thermally accessible protein–ligand
complexes provides a tool for accurate GPCR–ligand predictions
consistent with the GPCR pleiotropy paradigm. The validation by
both the mutation data and the subsequently published crystal
structure for the MVC:CCR5 complex indicates that this approach
may be trusted for new GPCRs for which there are little or no data.
Materials and Methods
The computational methods used in this study for the structure prediction of
CCR5 receptor in the WT and mutant forms as well as ligand docking were
described briefly in Results. Key aspects of the method for predicting re-
ceptor structure are that we do a very complete sampling of helix tilts and
rotations (∼11 billion conformations) and use energy as the sole criterion to
select a conformational ensemble for docking the ligands. Key aspects of the
method for predicting the binding site are that for each of the ligands’
multiple torsional conformations, we docked ∼50,000 poses to the receptor’s
full conformational ensemble, from which we eventually select the best pose
solely on the basis of energy. SI Materials and Methods contains detailed
descriptions of those computational methods along with the experimental
methods used in this study for the receptor mutant generation and ligand-
binding assays.
ACKNOWLEDGMENTS. We thank Soo-Kyung Kim and Andrea Kirkpatrick
for useful discussions. A portion of this research was funded by a gift
from Accelerator/PharmSelex. The balance was from gifts to the Materials
and Process Simulation Center.
1. Dragic T, et al. (1996) HIV-1 entry into CD4+ cells is mediated by the chemokine re-
ceptor CC-CKR-5. Nature 381(6584):667–673.
2. Dorr P, et al. (2005) Maraviroc (UK-427,857), a potent, orally bioavailable, and selective
small-molecule inhibitor of chemokine receptor CCR5 with broad-spectrum anti-human
immunodeficiency virus type 1 activity. Antimicrob Agents Chemother 49(11):4721–4732.
3. Stupple PA, et al. (2011) An imidazopiperidine series of CCR5 antagonists for the
treatment of HIV: The discovery of N-(1S)-1-(3-fluorophenyl)-3-[(3-endo)-3-(5-iso-
butyryl-2-methyl-4,5,6,7-tetrahydro-1H-imidazo[4,5-c]pyridin-1-yl)-8-azabicyclo[3.2.1]
oct-8-yl]propylacetamide (PF-232798). J Med Chem 54(1):67–77.
4. Baba M, et al. (1999) A small-molecule, nonpeptide CCR5 antagonist with highly
potent and selective anti-HIV-1 activity. Proc Natl Acad Sci USA 96(10):5698–5703.
5. Maeda K, et al. (2001) Novel low molecular weight spirodiketopiperazine derivatives
potently inhibit R5 HIV-1 infection through their antagonistic effects on CCR5. J Biol
Chem 276(37):35194–35200.
6. Grunbeck A, et al. (2012) Genetically encoded photo-cross-linkers map the binding site
of an allosteric drug on a G protein-coupled receptor. ACS Chem Biol 7(6):967–972.
7. Berro R, et al. (2013) Use of G-protein-coupled and -uncoupled CCR5 receptors by
CCR5 inhibitor-resistant and -sensitive human immunodeficiency virus type 1 variants.
J Virol 87(12):6569–6581.
8. Tan Q, et al. (2013) Structure of the CCR5 chemokine receptor-HIV entry inhibitor
maraviroc complex. Science 341(6152):1387–1390.
9. Abrol R, Griffith AR, Bray JK, Goddard WA, III (2012) Structure prediction of G protein-
coupled receptors and their ensemble of functionally important conformations. Mem-
brane Protein Structure and Dynamics: Methods and Protocols, Methods in Molecular
Biology, eds Vaidehi N, Klein-Seetharaman J (Humana, New York), Vol 914, pp 237–254.
10. Abrol R, Bray JK, Goddard WA, III (2011) Bihelix: Towards de novo structure prediction
of an ensemble of G-protein coupled receptor conformations. Proteins 80(2):505–518.
11. Okada T, et al. (2004) The retinal conformation and its environment in rhodopsin in
light of a new 2.2 A crystal structure. J Mol Biol 342(2):571–583.
12. Cherezov V, et al. (2007) High-resolution crystal structure of an engineered human
beta2-adrenergic G protein-coupled receptor. Science 318(5854):1258–1265.
13. Jaakola VP, et al. (2008) The 2.6 angstrom crystal structure of a human A2A adenosine
receptor bound to an antagonist. Science 322(5905):1211–1217.
14. Warne T, et al. (2008) Structure of a beta1-adrenergic G-protein-coupled receptor.
Nature 454(7203):486–491.
15. Kam VWT, Goddard WA, III (2008) Flat-bottom strategy for improved accuracy in
protein side-chain placements. J Chem Theory Comput 4(12):2160–2169.
16. Mayo SL, Olafson BD, Goddard WA, III (1990) Dreiding: A generic force-field for
molecular simulations. J Phys Chem-Us 94(26):8897–8909.
17. Bray JK, Abrol R, Goddard WA, III, Trzaskowski B, Scott CE (2014) SuperBiHelix method for
predicting the pleiotropic ensemble of G-protein–coupled receptor conformations. Proc
Natl Acad Sci USA 111(1):E72–E78.
18. Goddard WA, III, et al. (2010) Predicted 3D structures for adenosine receptors bound
to ligands: comparison to the crystal structure. J Struct Biol 170(1):10–20.
19. Kim S-K, Riley L, Abrol R, Jacobson KA, Goddard WA, III (2011) Predicted structures of agonist
and antagonist bound complexes of adenosine A3 receptor. Proteins 79(6):1878–1897.
20. Hall SE, et al. (2009) Elucidation of binding sites of dual antagonists in the human
chemokine receptors CCR2 and CCR5. Mol Pharmacol 75(6):1325–1336.
21. Maeda K, et al. (2006) Structural and molecular interactions of CCR5 inhibitors with
CCR5. J Biol Chem 281(18):12688–12698.
22. Kondru R, et al. (2008) Molecular interactions of CCR5 with major classes of small-
molecule anti-HIV CCR5 antagonists. Mol Pharmacol 73(3):789–800.
23. Berro R, et al. (2011) Multiple CCR5 conformations on the cell surface are used dif-
ferentially by human immunodeficiency viruses resistant or sensitive to CCR5 in-
hibitors. J Virol 85(16):8227–8240.
24. Vaidehi N, et al. (2006) Predictions of CCR1 chemokine receptor structure and BX 471
antagonist binding followed by experimental validation. J Biol Chem 281(37):27613–27620.
25. Dougherty DA (1996) Cation-pi interactions in chemistry and biology: A new view of
benzene, Phe, Tyr, and Trp. Science 271(5246):163–168.
26. Ahn KH, Scott CE, Abrol R, Goddard WA, III, Kendall DA (2013) Computationally-
predicted CB1 cannabinoid receptor mutants show distinct patterns of salt-bridges
that correlate with their level of constitutive activity reflected in G protein coupling
levels, thermal stability, and ligand binding. Proteins 81(8):1304–1317.
27. Scott CE, Abrol R, Ahn KH, Kendall DA, Goddard WA, III (2013) Molecular basis for
dramatic changes in cannabinoid CB1 G protein-coupled receptor activation upon
single and double point mutations. Protein Sci 22(1):101–113.
28. Abrol R, Kim S-K, Bray JK, Griffith AR, Goddard WA, III (2011) Characterizing and
predicting the functional and conformational diversity of seven-transmembrane
proteins. Methods 55(4):405–414.
Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13045
BIOCHEMISTRY

More Related Content

What's hot

V(D)J rearrangements and Antigen Antibody interactions
V(D)J rearrangements and Antigen Antibody interactionsV(D)J rearrangements and Antigen Antibody interactions
V(D)J rearrangements and Antigen Antibody interactionsTathagat Sah
 
Arf6 Reliability Paper - LinkedIn
Arf6 Reliability Paper - LinkedInArf6 Reliability Paper - LinkedIn
Arf6 Reliability Paper - LinkedInKenneth Hee
 
Basler modellers.210126reduced
Basler modellers.210126reducedBasler modellers.210126reduced
Basler modellers.210126reducedOlivier Bignucolo
 
Curran Lab 2015 poster
Curran Lab 2015 posterCurran Lab 2015 poster
Curran Lab 2015 posterHanna Kiani
 
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...Karthik Raman
 
Poster_2015_Calderon_final
Poster_2015_Calderon_finalPoster_2015_Calderon_final
Poster_2015_Calderon_finalAlissa Calderon
 
Targeting giants - the pharmacology of adhesions GPCRs
Targeting giants - the pharmacology of adhesions GPCRsTargeting giants - the pharmacology of adhesions GPCRs
Targeting giants - the pharmacology of adhesions GPCRsLaura Berry
 
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...Purification, Peptide Sequencing and Study of Antiproliferative activity of L...
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...Dr. Antik Bose
 
Scholarly paper in thesis format- FL final
Scholarly paper in thesis format- FL finalScholarly paper in thesis format- FL final
Scholarly paper in thesis format- FL finalFoo Li
 
Brandon_Boone_Yale Poster_Final Version
Brandon_Boone_Yale Poster_Final VersionBrandon_Boone_Yale Poster_Final Version
Brandon_Boone_Yale Poster_Final VersionBrandon Boone
 

What's hot (20)

870.full
870.full870.full
870.full
 
emm201548a
emm201548aemm201548a
emm201548a
 
V(D)J rearrangements and Antigen Antibody interactions
V(D)J rearrangements and Antigen Antibody interactionsV(D)J rearrangements and Antigen Antibody interactions
V(D)J rearrangements and Antigen Antibody interactions
 
Maxadilan poster
Maxadilan posterMaxadilan poster
Maxadilan poster
 
Arf6 Reliability Paper - LinkedIn
Arf6 Reliability Paper - LinkedInArf6 Reliability Paper - LinkedIn
Arf6 Reliability Paper - LinkedIn
 
Basler modellers.210126reduced
Basler modellers.210126reducedBasler modellers.210126reduced
Basler modellers.210126reduced
 
Curran Lab 2015 poster
Curran Lab 2015 posterCurran Lab 2015 poster
Curran Lab 2015 poster
 
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...
Fast-SL: An efficient algorithm to identify synthetic lethals in metabolic ne...
 
onc2013588a
onc2013588aonc2013588a
onc2013588a
 
BCSRCv1.3
BCSRCv1.3BCSRCv1.3
BCSRCv1.3
 
CCMB document
CCMB documentCCMB document
CCMB document
 
Poster_2015_Calderon_final
Poster_2015_Calderon_finalPoster_2015_Calderon_final
Poster_2015_Calderon_final
 
Targeting giants - the pharmacology of adhesions GPCRs
Targeting giants - the pharmacology of adhesions GPCRsTargeting giants - the pharmacology of adhesions GPCRs
Targeting giants - the pharmacology of adhesions GPCRs
 
NiH_Presentation
NiH_PresentationNiH_Presentation
NiH_Presentation
 
Complement final
Complement finalComplement final
Complement final
 
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...Purification, Peptide Sequencing and Study of Antiproliferative activity of L...
Purification, Peptide Sequencing and Study of Antiproliferative activity of L...
 
Scholarly paper in thesis format- FL final
Scholarly paper in thesis format- FL finalScholarly paper in thesis format- FL final
Scholarly paper in thesis format- FL final
 
Aacr2009 chip
Aacr2009 chipAacr2009 chip
Aacr2009 chip
 
Paper 5
Paper 5Paper 5
Paper 5
 
Brandon_Boone_Yale Poster_Final Version
Brandon_Boone_Yale Poster_Final VersionBrandon_Boone_Yale Poster_Final Version
Brandon_Boone_Yale Poster_Final Version
 

Viewers also liked

Claiming Love Consciousness
Claiming Love ConsciousnessClaiming Love Consciousness
Claiming Love ConsciousnessZoe Golightly
 
Valentine's Day Gifts and Gift Baskets
Valentine's Day Gifts and Gift BasketsValentine's Day Gifts and Gift Baskets
Valentine's Day Gifts and Gift BasketsGiftblooms Patel
 
Los talleres en educación infantil 1
Los talleres en educación infantil 1Los talleres en educación infantil 1
Los talleres en educación infantil 1Mileynaru
 
ประวัติส่วนตัว
ประวัติส่วนตัวประวัติส่วนตัว
ประวัติส่วนตัวratchanonrk
 
індивідуальний план самоосвіти вчителя історії
індивідуальний план самоосвіти вчителя історіїіндивідуальний план самоосвіти вчителя історії
індивідуальний план самоосвіти вчителя історіїНаталія Бондаренко
 
Lateralidad julieth palacio ruiz
Lateralidad julieth palacio ruizLateralidad julieth palacio ruiz
Lateralidad julieth palacio ruiz7e2e3a4b
 
Areas de la psicomotricidad
Areas de la psicomotricidadAreas de la psicomotricidad
Areas de la psicomotricidadpaovz
 
Chemistry of benzene
Chemistry of benzeneChemistry of benzene
Chemistry of benzeneKandarp Vyas
 
Organic Chemistry : Aromatic Compound (Benzene)
Organic Chemistry : Aromatic Compound (Benzene)Organic Chemistry : Aromatic Compound (Benzene)
Organic Chemistry : Aromatic Compound (Benzene)Thivyaapriya Sambamoorthy
 

Viewers also liked (20)

CEFR in EFL Classroom
CEFR in EFL Classroom CEFR in EFL Classroom
CEFR in EFL Classroom
 
Drinking the Kool Aid: How and Why for Internal Brands
Drinking the Kool Aid: How and Why for Internal BrandsDrinking the Kool Aid: How and Why for Internal Brands
Drinking the Kool Aid: How and Why for Internal Brands
 
Claiming Love Consciousness
Claiming Love ConsciousnessClaiming Love Consciousness
Claiming Love Consciousness
 
Valentine's Day Gifts and Gift Baskets
Valentine's Day Gifts and Gift BasketsValentine's Day Gifts and Gift Baskets
Valentine's Day Gifts and Gift Baskets
 
Riesgo físico
Riesgo físicoRiesgo físico
Riesgo físico
 
Los talleres en educación infantil 1
Los talleres en educación infantil 1Los talleres en educación infantil 1
Los talleres en educación infantil 1
 
ประวัติส่วนตัว
ประวัติส่วนตัวประวัติส่วนตัว
ประวัติส่วนตัว
 
Mapa conceptual 2do parcial
Mapa conceptual 2do parcialMapa conceptual 2do parcial
Mapa conceptual 2do parcial
 
Стратегічний напрямок II «Майстерність та професійність»
Стратегічний напрямок II «Майстерність та професійність»Стратегічний напрямок II «Майстерність та професійність»
Стратегічний напрямок II «Майстерність та професійність»
 
Who We Are
Who We AreWho We Are
Who We Are
 
Melanoma seminario
Melanoma seminarioMelanoma seminario
Melanoma seminario
 
індивідуальний план самоосвіти вчителя історії
індивідуальний план самоосвіти вчителя історіїіндивідуальний план самоосвіти вчителя історії
індивідуальний план самоосвіти вчителя історії
 
Lateralidad julieth palacio ruiz
Lateralidad julieth palacio ruizLateralidad julieth palacio ruiz
Lateralidad julieth palacio ruiz
 
MSC certificate
MSC certificateMSC certificate
MSC certificate
 
Areas de la psicomotricidad
Areas de la psicomotricidadAreas de la psicomotricidad
Areas de la psicomotricidad
 
Chemistry of benzene
Chemistry of benzeneChemistry of benzene
Chemistry of benzene
 
Benzene
BenzeneBenzene
Benzene
 
Organic Chemistry : Aromatic Compound (Benzene)
Organic Chemistry : Aromatic Compound (Benzene)Organic Chemistry : Aromatic Compound (Benzene)
Organic Chemistry : Aromatic Compound (Benzene)
 
Quizbank quiz buzzer
Quizbank quiz buzzer Quizbank quiz buzzer
Quizbank quiz buzzer
 
Unidad 2-globalizacion
Unidad 2-globalizacionUnidad 2-globalizacion
Unidad 2-globalizacion
 

Similar to PNAS_CCR5_drug_binding

Structure Prediction of WDR13 and a study of its Interacting Partners
Structure Prediction of WDR13 and a study of its Interacting PartnersStructure Prediction of WDR13 and a study of its Interacting Partners
Structure Prediction of WDR13 and a study of its Interacting PartnersAshish Baghudana
 
Biophysical Society Poster
Biophysical Society PosterBiophysical Society Poster
Biophysical Society PosterBilly Nicholson
 
Chem100 poster-Thanh-V3qBq-SM1
Chem100 poster-Thanh-V3qBq-SM1Chem100 poster-Thanh-V3qBq-SM1
Chem100 poster-Thanh-V3qBq-SM1Phuc Tran
 
Presentation july 31_2015
Presentation july 31_2015Presentation july 31_2015
Presentation july 31_2015gkoytiger
 
Medicilon KRAS-targeted Drugs R&D Service.pdf
Medicilon KRAS-targeted Drugs R&D Service.pdfMedicilon KRAS-targeted Drugs R&D Service.pdf
Medicilon KRAS-targeted Drugs R&D Service.pdfmedicilonz
 
Synthesis of Novel Protected Na(x-Drug) Amino Acid Building
Synthesis of Novel Protected Na(x-Drug) Amino Acid BuildingSynthesis of Novel Protected Na(x-Drug) Amino Acid Building
Synthesis of Novel Protected Na(x-Drug) Amino Acid Buildingsagiv weintraub
 
Research Inventy : International Journal of Engineering and Science
Research Inventy : International Journal of Engineering and ScienceResearch Inventy : International Journal of Engineering and Science
Research Inventy : International Journal of Engineering and Scienceresearchinventy
 
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...BRNSS Publication Hub
 
CellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group ProgrammeCellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group Programmerichardmiddleton
 
Raypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxRaypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxDrRajeshDas
 
Raypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxRaypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxDrRajeshDas
 
bradykinin receptor antagonist and thrombin inhibitors
bradykinin receptor antagonist and thrombin inhibitorsbradykinin receptor antagonist and thrombin inhibitors
bradykinin receptor antagonist and thrombin inhibitorsBhanuSri28
 
Homology Modeling of CCR3 Receptor
Homology Modeling of CCR3 ReceptorHomology Modeling of CCR3 Receptor
Homology Modeling of CCR3 Receptordharmakarma
 
Gutell 090.bmc.bioinformatics.2004.5.105
Gutell 090.bmc.bioinformatics.2004.5.105Gutell 090.bmc.bioinformatics.2004.5.105
Gutell 090.bmc.bioinformatics.2004.5.105Robin Gutell
 
RAD51 Drug Discovery (ACS Denver 2011)
RAD51 Drug Discovery (ACS Denver 2011)RAD51 Drug Discovery (ACS Denver 2011)
RAD51 Drug Discovery (ACS Denver 2011)Anthony Coyne
 

Similar to PNAS_CCR5_drug_binding (20)

Structure Prediction of WDR13 and a study of its Interacting Partners
Structure Prediction of WDR13 and a study of its Interacting PartnersStructure Prediction of WDR13 and a study of its Interacting Partners
Structure Prediction of WDR13 and a study of its Interacting Partners
 
Biophysical Society Poster
Biophysical Society PosterBiophysical Society Poster
Biophysical Society Poster
 
QTL mapping
QTL mappingQTL mapping
QTL mapping
 
Chem100 poster-Thanh-V3qBq-SM1
Chem100 poster-Thanh-V3qBq-SM1Chem100 poster-Thanh-V3qBq-SM1
Chem100 poster-Thanh-V3qBq-SM1
 
IJBB-10803348
IJBB-10803348IJBB-10803348
IJBB-10803348
 
Presentation july 31_2015
Presentation july 31_2015Presentation july 31_2015
Presentation july 31_2015
 
RENIN1.ppt
RENIN1.pptRENIN1.ppt
RENIN1.ppt
 
3 D QSAR Approaches and Contour Map Analysis
3 D QSAR Approaches and Contour Map Analysis3 D QSAR Approaches and Contour Map Analysis
3 D QSAR Approaches and Contour Map Analysis
 
Medicilon KRAS-targeted Drugs R&D Service.pdf
Medicilon KRAS-targeted Drugs R&D Service.pdfMedicilon KRAS-targeted Drugs R&D Service.pdf
Medicilon KRAS-targeted Drugs R&D Service.pdf
 
Synthesis of Novel Protected Na(x-Drug) Amino Acid Building
Synthesis of Novel Protected Na(x-Drug) Amino Acid BuildingSynthesis of Novel Protected Na(x-Drug) Amino Acid Building
Synthesis of Novel Protected Na(x-Drug) Amino Acid Building
 
Research Inventy : International Journal of Engineering and Science
Research Inventy : International Journal of Engineering and ScienceResearch Inventy : International Journal of Engineering and Science
Research Inventy : International Journal of Engineering and Science
 
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...
Molecular Modeling of Metalloreductase STEAP2 Protein and Docking Interaction...
 
CellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group ProgrammeCellAura Technologies Fluorescent Ligand User Group Programme
CellAura Technologies Fluorescent Ligand User Group Programme
 
Raypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxRaypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptx
 
Raypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptxRaypur(NEW) QSAR.pptx
Raypur(NEW) QSAR.pptx
 
bradykinin receptor antagonist and thrombin inhibitors
bradykinin receptor antagonist and thrombin inhibitorsbradykinin receptor antagonist and thrombin inhibitors
bradykinin receptor antagonist and thrombin inhibitors
 
Homology Modeling of CCR3 Receptor
Homology Modeling of CCR3 ReceptorHomology Modeling of CCR3 Receptor
Homology Modeling of CCR3 Receptor
 
Gutell 090.bmc.bioinformatics.2004.5.105
Gutell 090.bmc.bioinformatics.2004.5.105Gutell 090.bmc.bioinformatics.2004.5.105
Gutell 090.bmc.bioinformatics.2004.5.105
 
Qsar
QsarQsar
Qsar
 
RAD51 Drug Discovery (ACS Denver 2011)
RAD51 Drug Discovery (ACS Denver 2011)RAD51 Drug Discovery (ACS Denver 2011)
RAD51 Drug Discovery (ACS Denver 2011)
 

More from Alexandre "Sasha" Nesterov (14)

TRPV Neuron 2015
TRPV Neuron 2015TRPV Neuron 2015
TRPV Neuron 2015
 
Archives
ArchivesArchives
Archives
 
MCB
MCBMCB
MCB
 
JBC1
JBC1JBC1
JBC1
 
PNAS
PNASPNAS
PNAS
 
EMBO
EMBOEMBO
EMBO
 
Oncogene1.PDF
Oncogene1.PDFOncogene1.PDF
Oncogene1.PDF
 
JBC2
JBC2JBC2
JBC2
 
Cancer Research 2
Cancer Research 2Cancer Research 2
Cancer Research 2
 
Skin_whiteners_patent_2015
Skin_whiteners_patent_2015Skin_whiteners_patent_2015
Skin_whiteners_patent_2015
 
Drug Fut
Drug FutDrug Fut
Drug Fut
 
CPB-UP302-Publication
CPB-UP302-PublicationCPB-UP302-Publication
CPB-UP302-Publication
 
JBS 2010
JBS 2010JBS 2010
JBS 2010
 
Nesterov et al_Science_signaling_2015
Nesterov et al_Science_signaling_2015Nesterov et al_Science_signaling_2015
Nesterov et al_Science_signaling_2015
 

PNAS_CCR5_drug_binding

  • 1. Ligand- and mutation-induced conformational selection in the CCR5 chemokine G protein-coupled receptor Ravinder Abrola,1,2,3 , Bartosz Trzaskowskia,1,4 , William A. Goddard IIIa,1,2 , Alexandre Nesterovb , Ivan Olaveb , and Christopher Ironsb a Materials and Process Simulation Center, California Institute of Technology, Pasadena, CA 91125; and b PharmSelex Inc., Seattle, WA 98102 Contributed by William A. Goddard III, July 27, 2014 (sent for review January 25, 2014) We predicted the structural basis for pleiotropic signaling of the C-C chemokine type 5 (CCR5) G protein-coupled receptor (GPCR) by predicting the binding of several ligands to the lower-energy conformations of the CCR5 receptor and 11 mutants. For each case, we predicted the ∼20 most stable conformations for the receptor along with the binding sites for four anti-HIV ligands. We found that none of the ligands bind to the lowest-energy apo-receptor confor- mation. The three ligands with a similar pharmacophore (Maraviroc, PF-232798, and Aplaviroc) bind to a specific higher-energy receptor conformation whereas TAK-779 (with a different pharmacophore) binds to a different high-energy conformation. This result is in agreement with the very different binding-site profiles for these ligands obtained by us and others. The predicted Maraviroc bind- ing site agrees with the recent structure of CCR5 receptor cocrystal- lized with Maraviroc. We performed 11 site-directed mutagenesis experiments to validate the predicted binding sites. Here, we inde- pendently predicted the lowest 10 mutant protein conformations for each of the 11 mutants and then docked the ligands to these lowest conformations. We found the predicted binding energies to be in excellent agreement with our mutagenesis experiments. These results show that, for GPCRs, each ligand can stabilize a different protein conformation, complicating the use of cocrystallized struc- tures for ligand screening. Moreover, these results show that a sin- gle-point mutation in a GPCR can dramatically alter the available low-energy conformations, which in turn alters the binding site, potentially altering downstream signaling events. These studies val- idate the conformational selection paradigm for the pleiotropic function and structural plasticity of GPCRs. protein structure prediction | ligand–protein binding prediction | conformational ensemble | functional selectivity | GEnSeMBLE Since the finding that individuals lacking the C-C chemokine receptor type 5 (CCR5) gene are resistant to HIV, the CCR5 receptor has been established as a coreceptor for macrophage- tropic viruses, including HIV, to enter host cells (1). Several drugs aimed at disrupting the HIV–CCR5 coupling have been developed. To understand the structural mechanism of ligand binding to G protein-coupled receptors (GPCRs) and how these CCR5 targeting drugs work, we predicted the low-energy struc- tures of the apo CCR5 receptor, which in turn were used to predict the binding sites for various ligands. The focus was on ligands already known to inhibit CCR5: Maraviroc (MVC) (2), PF-232798 (PF) (3), Aplaviroc (APL) (4), and TAK-779 (TAK) (5) because there are abundant experimental data on their binding to numerous CCR5 mutants, to test the predicted structures. The li- gand structures are shown in Scheme S1. For three ligands (MVC, PF, and APL), we predict a similar binding site pharmacophore centered on a protonated nitrogen, whereas the other ligand (TAK) containing a quaternary nitrogen leads to a very different binding site pharmacophore. To validate our predictions, we identified 11 singly mutated CCR5 receptor forms selected to test the predicted binding region for the ligands. The predicted structures of ligand- bound CCR5 complexes presented here have already been used in two recent studies to provide a structural basis for (i) biophysical measurements (6) mapping the binding mode of MVC to CCR5 using genetically encoded photo–cross-linkers and (ii) viral entry experiments (7) probing the basis for G protein-coupled and -uncoupled CCR5 receptors by MVC-resistant and -sensitive HIV- 1 viruses. While this work was being finalized for submission, the CCR5 receptor was crystallized with MVC (8), allowing a direct struc- tural validation of the predicted structures as discussed in Prediction of Ligand–CCR5 Structures and Comparison with the Crystal Structure. Results Prediction of the CCR5 Structure. To predict the ensemble of low-energy structures (conformations) for CCR5, we used the GEnSeMBLE method (9), which performs a complete conforma- tional sampling (∼11 billion) of transmembrane (TM) helix confor- mations accessible in the membrane. This method starts with templates for positioning the seven TM helices based on available experimental/predicted structures and then samples a large number of helix orientation angles as defined in Fig. 1. During conformational sampling, we first sampled only the helix rotation angles (η) (on a 15° grid), and the BiHelix energy Significance The C-C chemokine receptor type 5 (CCR5) G protein-coupled re- ceptor (GPCR) is a prime target for preventing HIV invasion. A major difficulty in developing effective therapeutics is that the CCR5 exhibits an ensemble of ∼10–20 distinct low-energy con- formations, each of which might favor binding to different ligands and/or lead to different downstream functions. X-ray structures generally provide only one of these conformations. We applied the GEnSeMBLE methodology to predict this ensemble, and we designed and carried out 11 experiments to validate the ability of this ensemble to predict binding of an HIV therapeutic to CCR5. We found that each of the mutations changes the binding site. The predicted effects of mutations on binding are in excellent agreement with experiments, providing CCR5 structures for designing new ligands. Author contributions: R.A. and W.A.G. designed research; R.A., B.T., W.A.G., A.N., I.O., and C.I. performed research; R.A., B.T., W.A.G., A.N., I.O., and C.I. analyzed data; and R.A., B.T., and W.A.G. wrote the paper. The authors declare no conflict of interest. 1 R.A., B.T., and W.A.G. contributed equally to this work. 2 To whom correspondence may be addressed. Email: wag@wag.caltech.edu or abrolr@ csmc.edu. 3 Present address: Departments of Medicine and Biomedical Sciences, Cedars-Sinai Medical Center, Los Angeles, CA 90048. 4 Present address: Centre of New Technologies, University of Warsaw, 02-089, Warsaw, Poland. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1413216111/-/DCSupplemental. 13040–13045 | PNAS | September 9, 2014 | vol. 111 | no. 36 www.pnas.org/cgi/doi/10.1073/pnas.1413216111
  • 2. (10) was used to evaluate the energy of all (24)7 ∼ 4.6 billion TM helix bundle conformations. For the human CCR5 receptor structure prediction, this sampling was done starting with each of the following four X-ray–derived templates: bovine rhodopsin (PDB ID code 1u19) (11), human β2 adrenergic receptor (PDB ID code 2rh1) (12), human adenosine A2A receptor (PDB ID code 3eml) (13), and turkey β1 adrenergic receptor (PDB ID code 2vt4) (14). The full seven-helix bundle was built for the top 1,000 conformations from each template, and the side chains were optimized using the SideChain Rotamer Energy Analysis Method (SCREAM) side-chain placement protocol (15), fol- lowed by minimization (10 steps using Dreiding force field) (16). The β2 receptor template-based conformational ensemble led to the lowest energies (most stable conformations). We then se- lected the most stable 16 CCR5 conformations, for each of which we simultaneously sampled all three helix orientation angles (θ, ϕ, and η) allowing −10°, 0°, and +10° for the θ tilt angle and the −15°, 0° , and +15° range for both ϕ azimuthal and η rotation angles. This procedure led to a total of (27)7 ∼11 billion TM bundle conformations, for each of which we evaluated the energy rapidly using the SuperBiHelix method (17). Then the lowest-energy 2,000 conformations were built into seven-helix bundles and op- timized, from which we selected the 20 lowest-energy con- formations (labeled WT1 to WT20) as shown in Table S1. From these conformations, we selected eight structurally diverse seven- helix structures (highlighted rows in Table S1) for further analysis and ligand docking. The lowest-energy conformation WT1 cor- responds to the predicted apo conformation of the receptor. None of the experimentally obtained GPCR structures (with the ex- ception of opsin) has been ligand-free so it remains to be con- firmed whether WT1 resembles the apo conformation of the CCR5 receptor. These best eight diverse structures were then used to predict the binding of CCR5 ligands. The next section shows that MVC binds most strongly to the WT7 conformation rather than the lowest-energy WT1 conformation. Prediction of Ligand–CCR5 Structures and Comparison with the Crystal Structure. The four ligands (MVC, PF, APL, and TAK) were minimized using the B3LYP flavor of density functional theory (DFT) (with the 6–311G** basis set) using the Jaguar software package (Jaguar, version 7.8; Schrödinger, LLC). A conformational search was performed over the rotatable bonds for each ligand, and ∼20–30 conformations were selected (based on energy and diversity) (SI Materials and Methods) for each li- gand. Then, each of these conformations was docked to the previously selected eight CCR5 conformations using the DarwinDock/GenDock method (18). This method for predicting ligand–protein structures samples a complete set (∼50,000) of ligand poses in the potential binding regions, which are then clustered based on root-mean-squared (rms) distances to obtain ∼2,000 family heads whose energies are then evaluated to select ∼200 families, for each of which the energies of all members are evaluated. The top 100 poses from this process are optimized to finally select the best protein–ligand structure. These ligand– CCR5 structure predictions were performed independently for each of the eight CCR5 structures using the DREIDING force field (16). The final predicted most stable structure for the MVC/CCR5 complex is shown in Fig. 2, overlaid with the complex structure from the recent crystal structure (8) and showing the most im- portant ligand–residue interactions. Binding of MVC stabilized the WT7 structure of CCR5, which was also stabilized by PF and APL (Fig. 3 A and B) whereas TAK selected the WT10 confor- mation (Fig. 3C). Similar results have been obtained earlier for the adenosine A3AR receptor, where all four selective agonists sta- bilized WT15 whereas all four selective antagonists stabilized WT2 or WT3 (19) conformations; for none of these eight ligands was the complex with WT1 of lowest energy. These results emphasize the importance of having the ensemble of low-energy structures, and they suggest that ligands can control the final conformation and perhaps downstream function of CCR5 as well as GPCRs in general. The biggest changes in the CCR5 structure upon binding MVC (WT1 to WT7) were in the η rotation angles of helices 2 and 7 (compare the binding site residues in a 2D representation in Fig. S1 A and D). The 20-degree rotation of helix 2 completely changes the interhelical hydrogen bond network of CCR5. In the wild-type protein, the structure of CCR5 is stabilized by several interactions, including a conserved N48 (TM1)–D76 (TM2) hy- drogen bond, S63 (TM2)–R126 (TM3) hydrogen bond, and fi- nally a 1-2-7 hydrogen bond network Y37 (TM1)–W86 (TM2)– T284 (TM7). The 1-2-7 network and the W86–T284 hydrogen bond, in particular, result in a compact structure for the protein that is not suitable to develop strong interactions with large ligands having a tertiary nitrogen atom, such as MVC. This is because the interaction with T284 forces W86 to be close to E283, the crucial residue for interacting with the basic part of the ligand. To overcome this problem, the W86-T284 hydrogen bond must break, which leads to a slight destabilization of the struc- ture, which forces helices 2 and 7 to rotate to find favorable interactions with other residues from other helices, such as π–π stacking with Y108. The existence of a similar receptor activation network was proposed recently (20), but it did not include T284; Fig. 1. Coordinate frame for the TM helix orientation in a membrane (reprinted with permission from ref. 28). The x-y plane defined as the middle plane of the membrane bilayer. The helical axis, defined as the least moment of inertia axis, intersects this plane at position x,y with residue “h.” The helical axis is tilted by an angle θ with respect to the z axis that is perpen- dicular to the membrane. The azimuthal angle ϕ of this tilt in the x-y plane and the rotation angle of the helix around this axis is the angle η. Fig. 2. Comparison of the predicted MVC binding site (MVC in yellow and binding site residues in cyan) with the experimental X-ray MVC binding site (MVC in purple and binding site residues in green). Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13041 BIOCHEMISTRY
  • 3. instead, the network consisting of Y37, W86, Y108, and E283 residues was proposed. Thus, binding of MVC causes helix 2 to rotate by 20 degrees, which opens up the binding pocket formed by helices 1, 2, and 7. This rotation also allows the W86 residue to find an interaction with Y108 on helix 3, while interacting with the ligand, whereas T284 still forms a hydrogen bond with Y37 on TM1. Addition- ally these rotations aid the hydrogen bonding between N71 (TM2)–N293 (TM7) due to a more favorable geometry of these two residues. Moreover, the 15-degree rotation of H7 allows the E283 residue to find a better orientation, optimizing the interaction be- tween the carboxylic group and the basic nitrogen from the tropane ring of MVC, as well as the hydroxyl group of Y251 on helix 6. The binding pocket of the predicted complex model is in very good agreement with the crystal structure of the MVC:CCR5 complex published just recently (8). Fig. 2 also shows the MVC– CCR5 interactions observed in the crystal structure of the complex for comparison with the predicted structure. Fig. S1 A and B shows these interactions using a 2D representation. All important interactions between the MVC and CCR5 residues were predicted, including (i) the salt bridge between the tropane nitrogen atom and Glu283, (ii) the hydrogen bond between the carboxyamide nitrogen and Tyr251, and (iii) the hydrophobic interactions of the phenyl group with Tyr108, Phe109, and Phe112. The predicted structure is not perfect and missed some interactions: (i) the hydrogen bond between the amine moiety of the triazole group and Tyr37 and hydrogen bonds between one of the fluorines on the cyclohexane ring with Thr195 and Thr259. The accuracy of the predicted CCR5 structure (obtained using the human β2 adrenergic receptor template) shows the strength of the BiHelix and SuperBihelix methodology. This approach allows sampling of a complete set of TM rotations and local tilts, going well beyond simple homology modeling. This method also allows the residues crucial for ligand binding to be correctly positioned in the binding pocket, leading to high-quality protein structures starting from relatively low-homology templates (25% sequence identity between β2AR and CCR5 in the TM region). The MVC ligand does not select the lowest-energy receptor conformation (WT1) for binding but a higher-energy one (WT7). The helix rotation η angles change from [0, −120, 0, 0, 15, 0, 0] for the WT1 conformation to [15, 0, 0, 105, 15, 0, 0] for the WT7 conformation, which is very close to the [15, 0, 10, 105, −15, 0, 0] angles calculated for the crystal conformation of CCR5 bound to MVC (8). The predicted TM domains are also in good agree- ment with the crystal structure (crystal TMs/predicted TMs): TM1, 26–57/26–57; TM2, 64–92/64–93; TM3, 97–132/98–131; TM4, 142–166/142–166; TM5, 187–222/192–224; TM6, 229–261/ 229–259; and TM7, 269–297/276–299. The predicted PF binding mode (Fig. 3A) is similar to that of MVC, based also on the same protein configuration WT7. This result is reasonable because the ligand structures are rather similar. Thus, the mechanism of PF binding to CCR5 protein is similar to that of MVC binding, involving also the rotation of helices 2 and 7, which opens up the 1-2-7 binding pocket to allow the ligand to bind to E283. Fig. S1C shows the PF binding site in a 2D representation. Although APL has a slightly different molecular scaffold, it shares with MVC and PF a strongly basic nitrogen atom positioned in the center of the molecule (see central N atom in the ligand structures shown in Scheme S1). Indeed, it interacts strongly with the E283 anchor point (Fig. 3B), similar to the MVC and PF cases. Thus, the preferred CCR5 configuration is WT7, just as for MVC and PF. However, the benzoic acid group of APL provides an additional anchoring point for the ligand, far from helix 7 and E283. It interacts with the K191 residue located on the verge of helix 5 and extracellular loop 2, leading to a large reduction of binding upon K191A mutation. This result is con- sistent with the experimental >100-fold decrease in binding af- finity of APL upon K191A mutation (21, 22). The mechanism of binding of APL also involves breaking or weakening the 1-2-7 hydrogen bond network (Y37-W86-T284), with E283 providing the anchor for the middle part of the ligand. Interestingly, the benzoic acid group also interacts with R168 on extracellular loop 2, also shown to be important for ligand binding in experimental assays (23). The TAK antagonist is based on a completely different scaf- fold containing a quaternary amine group in place of the pro- tonated nitrogen for the other three ligands, and giving the whole ligand a formal charge of +1. We predict that TAK prefers to bind to the WT10 conformation (Fig. 3C). The overall mechanism of opening the binding site via disruption of the W86-T284 hydrogen bond is preserved. As in the previous cases, the 1–2-7 hydrogen bond network is broken or weakened due to the 20-degree rotation of helix 2 and formation of a π-π stacking interaction between W86 and Y108. The T284-Y37 hydrogen bond stabilizes the ligand-bound structure, in agreement with the experimental data that show lower binding affinity of TAK upon the Y37A mutation (21). Additionally, the 15-degree rotation of helix 5 allows the T195 residue to rotate away from the toluyl group of the ligand, making more space available in the 3-5-6 pocket. On the other hand, the relatively large rotation of weakly interacting H4 does not seem to have any impact on the structure of the protein or ligand binding. A detailed analysis of the TAK binding site shows that W86 is the strongest interacting residue with this ligand, with a −3.7 kcal/mol contribution to the overall interaction energy, which agrees well with the experimental data. This favorable interaction comes from an imperfect π–π stacking between the tryptophan residue and the benzo[7]annulene part of the ligand. We predict that the contribution of E283 residue is the second strongest one (−3.3 kcal/mol). On the other hand, the interaction ofthe ligand with Y108 is very weak so that the relatively large effect of the Y108A mutation seen experimentally results because WT22 becomes the most stable TAK bound conformation in the Y108A receptor mutant. Thus, the large effect of the mutation is indirect through stabilizing a different protein conformation rather than a change in direct interaction with the ligand. All other residues in- vestigated experimentally show very weak interactions with TAK, consistent with the experimental mutation data. Mutation Experiments. Based on the binding structure for MVC to WT CCR5, we selected 15 single mutants to validate the predicted structure: F109Y, F112A, Q194A, Y251F, D276A, Q277A, W86A, A90H, T105A, Y108A, F109A, I198A, Y251A, Q280A, and E283A. For 11 of these mutants, binding studies of Fig. 3. The predicted binding sites for (A) PF-232798 (PF), (B) Aplaviroc (APL), and (C) TAK-779 (TAK). 13042 | www.pnas.org/cgi/doi/10.1073/pnas.1413216111 Abrol et al.
  • 4. MVC and PF were carried out. The predicted results were dramatically inconsistent with experiments for the A90H and F109A mutants. Also, the experiments suggested a bigger change in binding energy for PF vs. MVC in the A90H mutant. A fundamental assumption in our selection of mutations was that the structure of the binding site would not change signifi- cantly (the established paradigm for assessing the effect of mutations). These experimental results suggested that this as- sumption might be invalid for CCR5 and potentially other GPCRs. This result is also consistent with our earlier studies on binding of the ligand BX471 to CCR1 receptor, where we found that a single mutation of human CCR1 to mouse and rat CCR1 dramatically changed the binding (24). Rather than starting all over to predict the hierarchy of CCR5 structures for each of the 15 mutant receptors, we simplified the procedure as follows. We went back to the 100 lowest-energy structures from the conformational hierarchy predicted for WT CCR5; we mutated the single residue in question for each of those structures, and then we reassigned side chains, using SCREAM (15), and minimized the structures. The net result is often a dramatic reordering in the ranking of these 100 receptor structures. We then selected the 20 lowest-energy structures for docking the four ligands. This approach leads to dramatically different hierarchies, as shown in Table S2. For example, the W86A mutant makes the WT30 conformation fourth best. This outcome is easy to un- derstand from the structures. WT30 conformation has the W86 residue up against TM7 (Fig. S2), but replacing this residue by Ala in the W86A mutant removes this steric clash, dramatically improving the rank of the WT30 conformation to fourth. Next, we docked the 20–30 conformations for each of the four ligands to each of the ∼10 receptor conformations for each of the mutated proteins using the DarwinDock/GenDock methods (18). The results are shown in Table 1. For the top-scoring poses, a water displacement penalty was applied if a polar residue was mutated to a nonpolar residue and could have accommodated a water molecule in the apo protein conformation. This penalty affected only the Q194A mutant. The results in Table 1 show the pronounced effect of muta- tions in the active site. Thus, the WT7 configuration stabilized by MVC binding to WT CCR5 is not the best binding configuration for any of the 11 mutants, all of which were aimed at modifying the binding! Instead, WT5 conformation appears seven times, WT2 four times, WT3 three times, and WT10 once. These mutants were generated experimentally (as described in SI Materials and Methods, Experimental Generation of CCR5 Mutants) and used to measure Ki values for the binding of different ligands to CCR5 WT and mutant receptors (SI Materials and Methods, Ligand-Binding Experiments). Comparing relative Ki values to relative predicted binding energies for each case led to the results in Fig. 4, which show excellent agreement with experimentally obtained Ki values for MVC binding to the mutants relative to the WT receptors, with the binding constants varying over a range larger than 3 logs. The trend for the other three ligands is also correct, but the correlation is weaker than that for MVC, due in part to the fact that the experimental changes in binding upon mutations span less than three orders of magnitude (less than 3 logs) for these ligands, which is only slightly larger than the best expected resolution from the calculations of ∼2 logs [Fig. S3 (PF), Fig. S4 (APL), and Fig. S5 (TAK)]. These results overall confirm the prediction that different conformations can be stabi- lized under different conditions (mutants and/or ligands). Discussion Analysis of the Ligand–CCR5 Binding Site in the Context of Mutations and Ligand Structure–Activity Relationship Series. The experimental assays confirmed the predictions that the binding modes of MVC and PF232798 are rather similar, but with one exception. The W86A mutation gives a 629-fold decrease in binding of PF and a 29.8-fold decrease in binding for MVC. [Kondru et al. reported a 10-fold decrease in MVC binding upon W86A mutation (22) in slightly different conditions.] Indeed, the calculations find that for W86A, both MVC and PF bind to WT7, but with predicted energy changes of 3.1 kcal/mol and 3.5 kcal/mol. The difference in impact of the W86A mutation on MVC/PF binding comes in part from structural differences between those two ligands. Although they share a similar scaffold, they have a different stereochemistry for the five-membered nitrogen-rich ring that forces the boat conformation of the tropane ring for PF whereas MVC retains the favorable chair conformation. Also there is an additional cyclohexane ring in the terminal part of MVC that affects the pKa value of the N3 atom, making it more basic. The quantum mechanical calculations predict the pKa value of the nitrogen in the five-membered ring to be 5.0–6.0 for MVC, but 6.0–7.0 for PF, suggesting that MVC may exist in a doubly protonated state in the experimental environment (the tropane N atom is very basic for both MVC and PF, with a pre- dicted pKa of ∼9). Because, for WT, the residue W86 is close to the five-membered ring of both MVC and PF, this change in the protonation state could have been responsible for the predicted different interaction with tryptophan through a strong cation–π interaction (25). Table 1. Predicted lowest-energy protein structures for wild- type and mutant receptors in the apo and ligand-bound forms Mutations Lowest-energy CCR5 conformations Apo MVC PF APL TAK Wild-type wt1 wt7 wt7 wt7 wt10 F109Y wt3 wt5 wt5 wt5 wt3 F112A wt1 wt5 wt5 wt5 wt3 Q194A wt1 wt5 wt5 wt5 wt4 Y251F wt1 wt2 wt22 wt4 wt3 D276A wt1 wt2 wt7 wt3 wt2 Q277A wt1 wt5 wt5 wt5 wt3 W86A wt1 wt7 wt7 wt7 wt3 A90H wt3 wt10 wt5 wt10 wt3 T105A wt1 wt2 wt6 wt6 wt3 Y108A wt3 wt2 wt22 wt5 wt22 F109A wt1 wt5 wt7 wt7 wt3 I198A wt1 wt5 wt5 wt5 wt25 Y251A wt2 wt3 wt5 wt5 wt3 Q280A wt3 wt5 wt7 wt7 wt7 E283A wt3 wt3 wt22 wt22 wt3 Ligands used are Maraviroc (MVC), PF-232798 (PF), Aplaviroc (APL), and Tak-779 (TAK). This shows that a single mutation can dramatically affect the binding site. Fig. 4. Comparison of the predicted effect of Maraviroc binding to mutants vs. experimental data for the 11 cases studied experimentally. The red line is the fit (R2 = 0.89) whereas the green lines indicate a deviation of 1 kcal/mol. Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13043 BIOCHEMISTRY
  • 5. However, our predicted structure makes this hypothesis un- likely because W86 is oriented in a way not allowing for the π–cation interaction in the PF. Despite a similar orientation of W86 in the MVC-bound and PF-bound models, we predict a difference in the interaction energy between those two ligands and the tryptophan residue. The binding site energy analysis allows the ligand–protein interaction to be partitioned into contributions from each residue. For the MVC binding site, we found a ligand interaction energy of −2.0 kcal/mol with W86 (which is less than the E283 contribution of −3.2 kcal/mol) whereas for PF it is −4.2 kcal/mol (almost identical to −4.3 kcal/mol contribution of E283). This difference can be understood in terms of the geometries of the ligands (the different stereo- chemistry of the tropane ring-five membered ring connections) and the different position of the phenyl ring with respect to the tropane ring for MVC and PF, leading to different interaction energies. A second important difference in the experimental mutation data for these two ligands occurs for the A90H mutation, which changes the PF binding significantly, but has almost no effect on MVC binding. PF prefers the WT5 conformation for the A90H mutant whereas MVC prefers WT10. Even though both MVC and PF are anchored by E283 and occupy a similar position in the protein, they are structurally different in the part that interacts with helix 2. The bulkier pyridyl group of PF positions the ligand closer to the A90 residue, which does not affect the binding unless this residue is mutated to a larger moiety: i.e., the histidine residue. Indeed, MVC is relatively far from the A90 residue (which is on top of helix 2); as a result, replacing A90 by the bulky histidine does not affect its binding. However, the li- gand PF is positioned closer to A90 so that the mutation to histidine causes a steric clash, decreasing binding. This difference arises from the difference in the receptor conformation for the A90H mutant preferred by MVC (WT10) versus PF (WT5). On the other hand, all 11 mutations investigated force the TAK-bound CCR5 system to switch to a different conformation. This conformational change has, however, little impact on the computational mutation results. For all six mutations studied experimentally for TAK, the results from both structures were in good agreement with experimental data. Unfortunately, the structure–activity relationship (SAR) series for the PF compound showed (3) little structural diversity. The experimental data presented in ref. 3 agrees well with the PF232798 binding pose predicted here. All 3-Substituted 1,4,6,7- Tetrahydro-imidazo[4,5-c]pyridines that were tested experimentally (compounds 39 and 41a–h from ref. 3) show a very high affinity, on the level of the PF232798 system. In the predicted docking pose, this part of the ligand interacts with the flexible N terminus of the protein and there is enough space to incorporate any substituent suggested in the original work. The modification of the ligand yielding 3-substituted derivatives (compounds 20a–h and 29–32 from ref. 3) positions the terminal part of the ligand closer to TM2, but avoiding steric clashes with any protein residues. This short SAR series further supports the predicted PF232798 binding pose. APL (Aplaviroc) has been developed from a series of spi- rodiketopiperazine derivatives showing a relatively high CCR5 activity (5). In that study (5), four different spirodiketopiperazine derivatives were evaluated for anti–HIV-1 activity, and molecule E913 was selected as the lead compound for a newer round of optimizations. Our predicted docking pose of APL is in agree- ment with the experimental data for this series of compounds. The main difference among the earlier spirodiketopiperazine derivatives E910, E913, E916, and E917 is the length of the molecule’s terminal part targeting the TM3-TM4-TM5 region of CCR5. According to the predicted pose, there is enough space to incorporate even the longest, heptylbenzene tail present in E910. This region is also highly hydrophobic, with a number of residues (F109, F112, F168, and I198) able to make relatively strong interactions with the phenyl ring present at the end of the hy- drophobic tail of all these ligands. The difference responsible for the strong binding of APL is the addition of the benzoic acid group to the end of this tail, which is the most important dif- ference between APL and E913. In APL, the phenoxy linker is able to position the new carboxylic group to promote a strong interaction with K191, known to be important in APL binding from the experimental mutation data. For TAK-779, the original paper (4) does not disclose any SAR series. Other Takeda compounds with known CCR5 ac- tivity (like TAK-220) are completely different. Implications for the Conformational Ensemble View of GPCRs. Fig. 5 summarizes the results for all ligands and mutants. Fig. 5A shows that MVC, PF, and APL prefer to bind to the WT7 receptor conformation, not to the lowest-energy WT1 conformation of the apo-protein. TAK prefers to bind to the WT10 conforma- tion. This difference between TAK and other ligands arises be- cause TAK has a quaternary nitrogen group in place of the tertiary nitrogen of the other ligands. Fig. 5B shows how differ- ent mutants of the receptor stabilize distinct conformations and indicates which conformation is preferred by ligands for different mutants. These changes in conformation of the mutated apo- protein explain the differential mutational data for W86A and A90H mutants on binding of MVC vs. PF compounds. This re- sult has direct implications for the functional selectivity of GPCRs because it shows that different receptor conformations are stabilized under different conditions (ligand binding and/or Fig. 5. (A) Conformational shuffling upon binding of different ligands. (B) CCR5 conformations preferred by different ligands and mutants. 13044 | www.pnas.org/cgi/doi/10.1073/pnas.1413216111 Abrol et al.
  • 6. mutations), which potentially can have distinct downstream effects on signaling. These results show that CCR5 is a relatively plastic protein with distinctly different protein conformations stabilized by dif- ferent ligands. A single mutation can considerably alter the en- ergetics of the low-lying conformations to alter the preferred binding site. This outcome reinforces the conformational selec- tion basis of the GPCR pleiotropic function. These results sug- gest that the paradigm of designing a ligand for a particular protein structure will not be sufficient for developing new ther- apeutics to CCR5 (and likely other GPCRs). One must consider all of the low-energy protein conformations in the design and in determining structure–activity relationships. Moreover, one must be cautious about interpreting mutation data. Thus, we consider that methods presented in this work for predicting the effect of mutations on GPCR structure will be essential to developing highly selective and active ligands targeting these receptors. Conclusions We developed an ensemble-based method that samples billions of conformations to eventually select an ensemble of ∼20 low- energy protein conformations for the CCR5 receptor and showed how this conformational ensemble changes upon binding of four ligands and how it changes because of mutations. To assess the validity of the binding sites, 15 mutations were pre- dicted in CCR5 that should significantly modify the binding. Experimental binding studies were carried out for 11 of these mutants, and several cases were found that disagreed signifi- cantly from the expectations based on the predicted pose for binding of the ligand MVC to the WT protein. This result mo- tivated us to predict the ensemble of low-energy protein struc- tures for each of the 15 mutants, some of which showed major changes in the conformations available for ligand binding. We predicted the optimum ligand pose for the ensemble of low-en- ergy CCR5 conformations for each of 11 mutants and compared with our experimental binding results. We found that ligand binding to the predicting mutated receptor conformations agrees well with the experimental mutation data for all four ligands. This work leads to several important general conclusions. (i) Dif- ferent ligands can stabilize dramatically different seven-helix bundle conformations, complicating the use of experimental structures for designing new strongly interacting ligands. (ii) For CCR5 (and likely other GPCRs), even a single mutation can dramatically change the binding site, complicating experimental validation of ligand-binding structures. This result invalidates the standard par- adigm of using directly the GPCR structure bound to the ligand to explain mutation-binding data. (iii) The changes in protein con- formation stabilized by different ligands and/or mutants might lead to changes in functionality (a structural basis for GPCR pleiotropy). Indeed functional differences between agonists and inverse agonists (or antagonists) to GPCRs are well established. For ex- ample, recent studies show that single, double, and triple mutant forms of the CB1 receptor exhibit dramatic changes in function, switching from full inactivity to high constitutive activity and back for sequential mutations (26, 27). (iv) The GEnSeMBLE approach for predicting the ensemble of thermally accessible protein–ligand complexes provides a tool for accurate GPCR–ligand predictions consistent with the GPCR pleiotropy paradigm. The validation by both the mutation data and the subsequently published crystal structure for the MVC:CCR5 complex indicates that this approach may be trusted for new GPCRs for which there are little or no data. Materials and Methods The computational methods used in this study for the structure prediction of CCR5 receptor in the WT and mutant forms as well as ligand docking were described briefly in Results. Key aspects of the method for predicting re- ceptor structure are that we do a very complete sampling of helix tilts and rotations (∼11 billion conformations) and use energy as the sole criterion to select a conformational ensemble for docking the ligands. Key aspects of the method for predicting the binding site are that for each of the ligands’ multiple torsional conformations, we docked ∼50,000 poses to the receptor’s full conformational ensemble, from which we eventually select the best pose solely on the basis of energy. SI Materials and Methods contains detailed descriptions of those computational methods along with the experimental methods used in this study for the receptor mutant generation and ligand- binding assays. ACKNOWLEDGMENTS. We thank Soo-Kyung Kim and Andrea Kirkpatrick for useful discussions. A portion of this research was funded by a gift from Accelerator/PharmSelex. The balance was from gifts to the Materials and Process Simulation Center. 1. Dragic T, et al. (1996) HIV-1 entry into CD4+ cells is mediated by the chemokine re- ceptor CC-CKR-5. Nature 381(6584):667–673. 2. Dorr P, et al. (2005) Maraviroc (UK-427,857), a potent, orally bioavailable, and selective small-molecule inhibitor of chemokine receptor CCR5 with broad-spectrum anti-human immunodeficiency virus type 1 activity. Antimicrob Agents Chemother 49(11):4721–4732. 3. Stupple PA, et al. (2011) An imidazopiperidine series of CCR5 antagonists for the treatment of HIV: The discovery of N-(1S)-1-(3-fluorophenyl)-3-[(3-endo)-3-(5-iso- butyryl-2-methyl-4,5,6,7-tetrahydro-1H-imidazo[4,5-c]pyridin-1-yl)-8-azabicyclo[3.2.1] oct-8-yl]propylacetamide (PF-232798). J Med Chem 54(1):67–77. 4. Baba M, et al. (1999) A small-molecule, nonpeptide CCR5 antagonist with highly potent and selective anti-HIV-1 activity. Proc Natl Acad Sci USA 96(10):5698–5703. 5. Maeda K, et al. (2001) Novel low molecular weight spirodiketopiperazine derivatives potently inhibit R5 HIV-1 infection through their antagonistic effects on CCR5. J Biol Chem 276(37):35194–35200. 6. Grunbeck A, et al. (2012) Genetically encoded photo-cross-linkers map the binding site of an allosteric drug on a G protein-coupled receptor. ACS Chem Biol 7(6):967–972. 7. Berro R, et al. (2013) Use of G-protein-coupled and -uncoupled CCR5 receptors by CCR5 inhibitor-resistant and -sensitive human immunodeficiency virus type 1 variants. J Virol 87(12):6569–6581. 8. Tan Q, et al. (2013) Structure of the CCR5 chemokine receptor-HIV entry inhibitor maraviroc complex. Science 341(6152):1387–1390. 9. Abrol R, Griffith AR, Bray JK, Goddard WA, III (2012) Structure prediction of G protein- coupled receptors and their ensemble of functionally important conformations. Mem- brane Protein Structure and Dynamics: Methods and Protocols, Methods in Molecular Biology, eds Vaidehi N, Klein-Seetharaman J (Humana, New York), Vol 914, pp 237–254. 10. Abrol R, Bray JK, Goddard WA, III (2011) Bihelix: Towards de novo structure prediction of an ensemble of G-protein coupled receptor conformations. Proteins 80(2):505–518. 11. Okada T, et al. (2004) The retinal conformation and its environment in rhodopsin in light of a new 2.2 A crystal structure. J Mol Biol 342(2):571–583. 12. Cherezov V, et al. (2007) High-resolution crystal structure of an engineered human beta2-adrenergic G protein-coupled receptor. Science 318(5854):1258–1265. 13. Jaakola VP, et al. (2008) The 2.6 angstrom crystal structure of a human A2A adenosine receptor bound to an antagonist. Science 322(5905):1211–1217. 14. Warne T, et al. (2008) Structure of a beta1-adrenergic G-protein-coupled receptor. Nature 454(7203):486–491. 15. Kam VWT, Goddard WA, III (2008) Flat-bottom strategy for improved accuracy in protein side-chain placements. J Chem Theory Comput 4(12):2160–2169. 16. Mayo SL, Olafson BD, Goddard WA, III (1990) Dreiding: A generic force-field for molecular simulations. J Phys Chem-Us 94(26):8897–8909. 17. Bray JK, Abrol R, Goddard WA, III, Trzaskowski B, Scott CE (2014) SuperBiHelix method for predicting the pleiotropic ensemble of G-protein–coupled receptor conformations. Proc Natl Acad Sci USA 111(1):E72–E78. 18. Goddard WA, III, et al. (2010) Predicted 3D structures for adenosine receptors bound to ligands: comparison to the crystal structure. J Struct Biol 170(1):10–20. 19. Kim S-K, Riley L, Abrol R, Jacobson KA, Goddard WA, III (2011) Predicted structures of agonist and antagonist bound complexes of adenosine A3 receptor. Proteins 79(6):1878–1897. 20. Hall SE, et al. (2009) Elucidation of binding sites of dual antagonists in the human chemokine receptors CCR2 and CCR5. Mol Pharmacol 75(6):1325–1336. 21. Maeda K, et al. (2006) Structural and molecular interactions of CCR5 inhibitors with CCR5. J Biol Chem 281(18):12688–12698. 22. Kondru R, et al. (2008) Molecular interactions of CCR5 with major classes of small- molecule anti-HIV CCR5 antagonists. Mol Pharmacol 73(3):789–800. 23. Berro R, et al. (2011) Multiple CCR5 conformations on the cell surface are used dif- ferentially by human immunodeficiency viruses resistant or sensitive to CCR5 in- hibitors. J Virol 85(16):8227–8240. 24. Vaidehi N, et al. (2006) Predictions of CCR1 chemokine receptor structure and BX 471 antagonist binding followed by experimental validation. J Biol Chem 281(37):27613–27620. 25. Dougherty DA (1996) Cation-pi interactions in chemistry and biology: A new view of benzene, Phe, Tyr, and Trp. Science 271(5246):163–168. 26. Ahn KH, Scott CE, Abrol R, Goddard WA, III, Kendall DA (2013) Computationally- predicted CB1 cannabinoid receptor mutants show distinct patterns of salt-bridges that correlate with their level of constitutive activity reflected in G protein coupling levels, thermal stability, and ligand binding. Proteins 81(8):1304–1317. 27. Scott CE, Abrol R, Ahn KH, Kendall DA, Goddard WA, III (2013) Molecular basis for dramatic changes in cannabinoid CB1 G protein-coupled receptor activation upon single and double point mutations. Protein Sci 22(1):101–113. 28. Abrol R, Kim S-K, Bray JK, Griffith AR, Goddard WA, III (2011) Characterizing and predicting the functional and conformational diversity of seven-transmembrane proteins. Methods 55(4):405–414. Abrol et al. PNAS | September 9, 2014 | vol. 111 | no. 36 | 13045 BIOCHEMISTRY