SlideShare a Scribd company logo
1 of 6
Download to read offline
A dynamic Asp–Arg interaction is essential for catalysis
in microsomal prostaglandin E2 synthase
Joseph S. Brocka,1
, Mats Hamberga,1
, Navisraj Balagunaseelana
, Michael Goodmanb
, Ralf Morgensternc
,
Emilia Strandbacka
, Bengt Samuelssona,2
, Agnes Rinaldo-Matthisa
, and Jesper Z. Haeggströma,2
a
Department of Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden; b
Department of Chemistry, Vanderbilt University
School of Medicine, Nashville, TN 37232-6304; and c
Institute of Environmental Medicine, Karolinska Institutet, S-171 77 Stockholm, Sweden
Contributed by Bengt Samuelsson, December 18, 2015 (sent for review October 9, 2015; reviewed by Lawrence J. Marnett, Charles N. Serhan, and
Takao Shimizu)
Microsomal prostaglandin E2 synthase type 1 (mPGES-1) is responsible
for the formation of the potent lipid mediator prostaglandin E2 under
proinflammatory conditions, and this enzyme has received consider-
able attention as a drug target. Recently, a high-resolution crystal struc-
ture of human mPGES-1 was presented, with Ser-127 being proposed
as the hydrogen-bond donor stabilizing thiolate anion formation
within the cofactor, glutathione (GSH). We have combined site-directed
mutagenesis and activity assays with a structural dynamics analysis to
probe the functional roles of such putative catalytic residues. We found
that Ser-127 is not required for activity, whereas an interaction be-
tween Arg-126 and Asp-49 is essential for catalysis. We postulate that
both residues, in addition to a crystallographic water, serve critical roles
within the enzymatic mechanism. After characterizing the size or
charge conservative mutations Arg-126–Gln, Asp-49–Asn, and Arg-
126–Lys, we inferred that a crystallographic water acts as a general
base during GSH thiolate formation, stabilized by interaction with
Arg-126, which is itself modulated by its respective interaction with
Asp-49. We subsequently found hidden conformational ensembles
within the crystal structure that correlate well with our biochemical
data. The resulting contact signaling network connects Asp-49 to distal
residues involved in GSH binding and is ligand dependent. Our work
has broad implications for development of efficient mPGES-1 inhibitors,
potential anti-inflammatory and anticancer agents.
inflammation | prostaglandin | mPGES-1 | MAPEG | mechanism
Prostaglandin E2 (PGE2) is an abundant lipid mediator that
signals via four receptors (EP1–4) to induce an array of important
biological actions in physiology as well as pathophysiology (1). Under
proinflammatory conditions, biosynthesis of PGE2 proceeds
from arachidonic acid, which is converted to the unstable endoper-
oxide PGH2 by cyclooxygenase type 2 (COX-2). PGH2 is further
isomerized into PGE2 by microsomal PGE synthase type 1 (mPGES-
1) (2, 3). mPGES-1 is encoded by PTGES and is up-regulated by
mitogens and cytokines in a pathway that is functionally coupled to
COX-2 (2, 4). Because of its key role in PGE2 synthesis, mPGES-1
has attracted attention as a potential drug target in the areas of
inflammation, pain, fever, and cancer (5).
mPGES-1 is a member of the MAPEG (Membrane-Associated
Proteins in Eicosanoid and Glutathione metabolism) superfamily of
enzymes (6), which also includes two key proteins in the leukotriene
(LT) cascade, viz. 5-lipoxygenase activating protein and LT C4 syn-
thase (LTC4S). All MAPEG members are integral, homotrimeric
membrane proteins, and structural information on this family has
been scarce. However, significant progress has recently been made in
this area with several high-resolution structures being solved by X-ray
crystallography (7–9). In particular, the crystal structures of human
LTC4S provided detailed structural information, including an argi-
nine residue that was later shown to activate the glutathione (GSH)
thiolate (10, 11). We have previously proposed that this conserved
arginine residue is also essential for enzymatic activity in mPGES-1 as
Arg-126 (12). The recent structural determination of mPGES-1,
however, at an exceptionally high resolution of 1.16 Å, uncovered
several unanticipated structural features (13). The active sites, found
at the three monomeric interfaces, show that Ser-127 is positioned
near the GSH thiol group, indicating that it may act as a hydrogen-
bond donor to assist in thiolate formation during catalysis. Further-
more, Arg-126 and Asp-49 participate in a charge interaction that
could also contribute to catalysis. This structural information is
supported by a mesophase crystal structure of an engineered version
of mPGES-1 (14) and several inhibitor complexes that have recently
been published (15).
Here, we initially confirmed the necessity for GSH thiolate during
catalysis via incubations with the analogous tripeptide γ-Glu–Ser–Gly
(GOH). We then used site-directed mutagenesis to analyze the
functional roles of active site residues. We found that Ser-127 is
nonessential for catalysis, whereas Arg-126 and Asp-49 are crucial
and mutually dependent for native isomerase activity of the enzyme.
Because the latter codependence of activity could be rationalized by
a dynamic functional role of these residues, we turned to the high-
resolution X-ray data (13, 15) deposited in the Protein Data Bank
(PDB; www.rcsb.org) (16) to provide evidence of their dynamic
motion within the crystal structure. Several recent studies have
shown that the information present in such data is often under-
estimated and that it is possible to refine multiple conformations of
residues simultaneously, each with individually refined occupancy
and B factors, without overfitting the data (17–22). By such sampling
of low-level electron density, discrete, “hidden” conformations are
revealed, facilitating a more quantitative representation of dynamic
Significance
Microsomal prostaglandin E2 synthase type 1 (mPGES-1) is an
integral membrane protein that produces prostaglandin E2 (PGE2),
a mediator of inflammation, fever, pain, and tumorigenesis. Here
we show that a serine residue implicated by the crystal structure
is not required for function, whereas an arginine and aspartate
residue in the active site, observed to be interacting within the
crystal structure, are essential and mutually dependent during
catalysis. We also demonstrate that a contact signaling network
can interrupt the arginine–asparagine interaction and facilitate
their participation in the chemical mechanism. Our work has
broad implications for development of effective mPGES-1 inhib-
itors, potential drugs with clinical application in treatment of in-
flammatory diseases and cancer.
Author contributions: M.H., B.S., A.R.-M., and J.Z.H. designed research; J.S.B., M.H., N.B.,
and E.S. performed research; M.G. and R.M. contributed new reagents/analytic tools;
J.S.B., M.H., N.B., A.R.-M., and J.Z.H. analyzed data; and J.S.B. and J.Z.H. wrote the paper.
Reviewers: L.J.M., Vanderbilt University Medical Center; C.N.S., Brigham and Women’s
Hospital/Harvard Medical School; and T.S., National Center for Global Health
and Medicine.
The authors declare no conflict of interest.
Freely available online through the PNAS open access option.
1
J.S.B. and M.H. contributed equally to this work.
2
To whom correspondence may be addressed. Email: jesper.haeggstrom@ki.se or bengt.
samuelsson@ki.se.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1522891113/-/DCSupplemental.
972–977 | PNAS | January 26, 2016 | vol. 113 | no. 4 www.pnas.org/cgi/doi/10.1073/pnas.1522891113
motion within the crystal lattice. Furthermore, it has been shown that
this information is often essential for understanding enzymatic
function (23) and mechanism (24, 25) and successfully achieving
structure-based drug design (26).
We quantified the dynamic conformations of active site residues
using the software qFit (20) and CONTACT (27). This method
generated significant improvements in the quality indicators for
PDB ID codes 4AL0 and 4AL1 [1.16 and 1.95 Å, respectively (13)]
and revealed a ligand-dependent contact network that corroborates
the mechanism suggested from biochemical data. These van der
Waals interactions within the binary complex with GSH (PDB ID
code 4AL0) reveal an extensive network of correlated side-chain
motions within the cytoplasmic “C-domain” that forms the bottom
of the active site and confirm a dynamic role of Asp-49 in catalysis.
In comparison, the much smaller networks found in PDB ID code
4AL1, with a bisphenyl–GSH analog and detergent molecule
bound in the active site, indicate that ligand binding can influence
network signaling. This finding is also supported by our analysis of
recently published inhibitor complexes (15).
Our results suggest that the positively charged Arg-126 stabilizes
transient thiolate formation and that its dynamic interaction with
Asp-49 is essential for catalysis. We also observed a crystallo-
graphic water molecule that is ideally situated to act as a proton
acceptor during this process. Furthermore, we found a striking
contact signaling network within the active site that effects the
conformation of residues in a ligand-dependent fashion.
Results and Discussion
GSH Thiolate Is Essential for Catalysis. Fig. 1 depicts a substrate-
limited assay that measured absolute product formation (nano-
grams) as described in Materials and Methods. Incubations with
native enzyme in the presence of GSH resulted in near total con-
version of PGH2 to PGE2 and are concurrent with the specific
activity previously reported in the literature (∼4 μmol·min−1
·mg−1
at
0 °C) (28). A negative control involved microsomal preparations
of WT mPGES-1 resuspended in buffer containing the GSH
analog, GOH, that differs from the native cofactor only by the
replacement of the thiol moiety by a hydroxyl group and did not
produce product above background levels. This finding provides
strong evidence that thiolate anion is the chemical species of the
cofactor essential for catalysis.
Ser-127 Is Nonessential for Catalysis. Judging from the orientation
of Ser-127 in the recently published crystal structure (13), the
authors’ hypothesis that its hydroxyl group may act as a hydrogen-
bond donor to stabilize a GSH thiolate is apt, because this mech-
anism of thiol activation is a common theme within the evolution of
soluble GSH transferases (29). However, because we had pre-
viously proposed Arg-126 as a strong candidate for this role (12),
we investigated the function of Ser-127 in the conversion of PGH2
into PGE2.
To detail the role of Ser-127 in mPGES-1, we exchanged this
residue for an alanine by site-directed mutagenesis. After expres-
sion in Pichia pastoris and purification, aliquots of recombinant
protein were incubated with PGH2. Formation of PGE2 was ana-
lyzed by GC-MS. The combined measurements obtained from at
least three different preparations of enzyme are depicted in Fig. 1
and show that Ser-127–Ala exhibits the same level of PGE2 syn-
thase activity as WT mPGES-1. This finding was true for both
purified and microsomal preparations of the enzyme (Fig. 1). In
addition, the dual conformations observed for this residue in the
Fig. 1. Mutagenic analysis of mPGES-1 active site residues. Aliquots of WT,
S127A, D49N, R126Q, and R126K mPGES-1 were incubated with 12 μM PGH2 and
analyzed for PGE2 formation by GC-MS, as described in Materials and Methods.
The total amount (nanograms) of PGE2 formed is shown from both purified and
microsomal preparations of enzyme. PGE2 formation was also monitored for
microsomal preparations of WT enzyme incubated with GOH. Values represent
the combined measurements from at least three different preparations of en-
zyme (n = 3), with error bars representing their SD. The levels of PGF2α formed
were also measured by this method as shown in Fig. S1.
Fig. 2. The active site architecture of mPGES-1. The active sites of PDB ID codes 4AL0 (A) and 4AL1 (B) are compared with post-qFit conformational fitting and re-
finement. The coordinates of PDB depositions are overlaid in translucent over the refined qFit ensembles shown as opaque conformers in stick representation. The 2mFo-
DFc electron density corresponding to a mechanistically relevant solvent molecule is shown as blue mesh contoured at 1 rmsd, and the rotation plane of the R126
guanidinium is shown relative to the carboxylate of D49 (44.7°) (A). This water is absent within the qFit-refined bis-phenyl complex (B), potentially due to a reduced
capacity for GSH thiol interaction. A molecule of octyl glucoside bound at the C-domain and low occupancy GSH (∼13%) within the active site of B have been omitted for
clarity (cf. Fig. S3). Polar interactions are shown as dashed lines with distances given in Å.
Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 973
BIOCHEMISTRY
crystal structure indicate the absence of a strong hydrogen-bonding
interaction. Conversely, Arg-126 is observed in a single confor-
mation with an Nη-GSH thiol distance of 3.4 Å. We believe that
this active site geometry also substantiates strong evidence for a
mechanism of GSH thiol activation by an Arg-126 guanidinium
interaction (30). Hence, despite compelling structural evidence,
Ser-127 does not play a critical role in mPGES-1 catalysis.
Mutation of Arg-126 and Asp-49 Compromises PGE2 Synthase Activity,
but Allows PGH2 Reduction to PGF2α. In light of the new structural
data (13), we wanted to reexamine the functional role of Arg-126
and mutated this residue into both a glutamine and a lysine residue
using site-directed mutagenesis. According to the crystal structure
of mPGES-1 (13), Arg-126 and Asp-49 participate in an inter-
monomeric charge interaction. Therefore, we also mutated the
negatively charged counterpart, Asp-49, into an asparagine residue.
We anticipated that the size and charge conservative mutations of
these residues could serve in probing their role in the enzymatic
mechanism, while minimizing steric and electrostatic repulsion
effects, such as disruption of the monomer interface. Although we
also attempted to create the charge conservative mutant Asp-49–
Glu, the resulting transformed construct failed to express, pre-
sumably because it resulted in an unstable quaternary structure.
After solubilization with detergent and purification via Ni-affinity
chromatography, these mutants were assayed for PGE2 synthase
activity as described above. For three different purifications of each
isoform, we found that the mutated enzymes did not convert PGH2
into PGE2 above background levels. After preparations of micro-
somal fractions, however, we found that the charge conservative
mutation Arg-126–Lys still retained a low level of isomerase activity,
indicating that a native membrane environment and a formal posi-
tive charge at position 126 are important factors for catalysis (Fig. 1).
From these results, we conclude that both Arg-126 and Asp-49 are
key to the PGE2 synthase activity of mPGES-1.
That both of these residues are essential for catalysis is intriguing,
because one could expect Arg-126 to be precluded from participating
Fig. 3. Contact signaling within mPGES-1. (A) The van der Waals contact network identified by qFit conformational fitting and subsequent CONTACT analysis
of mPGES-1 PDB entries are shown with translucent molecular surface representations over alternate conformers and correspondingly colored node diagrams.
The nodes are connected by edges whose width is weighted according to the number of networks involving the pair they connect. The network observed
within PDB ID code 4AL0, which includes the active site residue D49, is shown in red, both from the perspective of the membrane plane (Left) and the cy-
toplasm (Right). (B) The corresponding contact networks identified from the qFit ensemble complex within the bis-phenyl GSH complex (PDB ID code 4AL1)
are much smaller and are shown in cyan and red. The latter contains the active site residue R126, now observed in dual conformations, possibly due to
negation of GSH thiol interaction. One possible pathway is show in more detail within Movie S1.
974 | www.pnas.org/cgi/doi/10.1073/pnas.1522891113 Brock et al.
in thiolate stabilization if it was already engaged in a stable salt bridge
interaction with Asp-49. Analysis of the relative torsional angles,
however, shows the out-of-plane angle of the Asp-49 carboxylate
relative to the Arg-126 guanidinium to be 44.7° (Fig. 2A). This value
is far in excess of the ∼8–10° found to be typical of bidentate inter-
actions for structures of a similar resolution as reported in a recent
comprehensive review (31). Therefore, the Asp-49–Arg-126 in-
teraction cannot be classified as the energetically stable, bidentate
interaction of a formal salt bridge and implies that the energetic
barrier for its disruption would be low. This finding provides evidence
for the capacity of these residues to dynamically participate in active
site chemistry and is corroborated by conformational fitting with
qFit (27), which reveals hidden conformations of both residues
depending on the identity of the adjacent ligand (Fig. 2).
As we had previously observed for Arg-126 mutants (12), we
found that other catalytically inactive mutants assayed in this
study displayed a promiscuous reductase activity, converting
PGH2 into PGF2α. Notably, the most pronounced activity in this
respect was again observed for microsomal preparations of the
Arg-126–Lys mutant (Fig. S1).
We confirmed that all mutants possessed the same tertiary fold
as native enzyme via comparison of circular-dichroism spectra
(Fig. S2).
A Crystallographic Water Molecule Is Ideally Situated to Participate in
the Mechanism. Analysis of the active site architecture also sug-
gests that the α-carboxylate of GSH is involved in thiolate for-
mation, via a tightly bound crystallographic water (2Fo − Fc peak
of ∼5 rmsd, ADP = 21.9 Å2
) within the active site (Fig. 2A). By
forming a hydrogen-bonding network from the α-carboxylate
moiety of GSH to its thiol group, it is ideally placed to assist in
deprotonation of the latter during catalysis. The pKa of the
α-carboxylate, in turn, is undoubtedly lowered by the side-on,
out-of-plane interaction with the guanidinium of Arg-38 (torsion
angle 45.9°), which is itself engaged in solvent-mediated inter-
actions with the main-chain carboxyl groups of Ala-43 and Arg-
60. This architecture is highly reminiscent of the “electron-sharing
network” that is functionally conserved in all classes of soluble
GSTs for the same purpose (32), and the use of a bridging water
molecule to transfer the thiol proton to the α-carboxylate of
GSH has been shown to be energetically favorable within an
alpha class soluble GST (33). Crucially, density for this water is
absent for the Phenix (34) refined bis-phenyl GSH complex
(PDB ID code 4AL1), in which the relative occupancies to GSH
were refined as 0.87:0.13, respectively. After conformational change
of Asp-49, we hypothesize that Arg-126 can further decrease the
GSH thiol pKa via charge stabilization. The crystallographic
water molecule could then function as the yet-unidentified base
that accepts a proton from GSH during thiolate formation,
concurrently forming a transient hydronium ion or shuttling the
proton to the α-carboxylate. Once formed and stabilized by in-
teraction with Arg-126, we expect attack of GSH thiolate upon
the endoperoxide ring at the C-9 position, resulting in O–O bond
cleavage and proton donation via the hydronium ion. Asp-49,
liberated from its interaction with Arg-126, would now be free to
function as a base within the resulting transition state, facilitating
a decrease of the C-9 proton pKa, and spontaneous decomposition
to yield the product PGE2 and regenerated GSH (Fig. 4). Although
an alternative mechanism in which thiolate would act as a general
base abstracting the C-9 proton has been suggested to be more
energetically favorable in model systems (35, 36), the probability
of either pathway would ultimately be determined by the precise
orientation of substrate relative to cofactor within the enzymatic
active site. Although the apparent ability of the active site mu-
tants characterized here to produce PGF2α via reduction of a
putative sulphenic acid ester intermediate speaks in favor of the
former (Fig. S1), this alternative mechanism is shown in Fig. S4.
A Contact Signaling Network Modulates Active Site Residues in mPGES-1.
We submitted the PDB entries associated with the recently published
crystal structure of mPGES-1 (PDB ID codes 4AL0 and 4AL1)
(13) to the qFit server (smb.slac.stanford.edu/qFitServer/) (20). This
software automatically samples conformational heterogeneity that is
interpretable by fitting partial occupancy conformational ensembles
into low-level electron density. The CONTACT algorithm was then
used to calculate resulting van der Waals contact networks that in-
dicate a probable correlation of conformations at each site (27).
Post-qFit conformational fitting and subsequent refinement by
Phenix (34) of the 1.16-Å binary complex with GSH (PDB ID code
4AL0) resulted in a small, but significant, improvement of structure
quality indicators, including the decrease of R/Rfree values from
12.2/13.0% to 11.6/12.8%, respectively. Subsequent analysis of the
structure with CONTACT revealed an extensive network of cor-
related side chain interactions centered upon the short, cytoplasmic
helix separating transmembrane helices I and II that was referred to
by Sjögren et al. (13), and will be hereafter, as the C-domain. Of
most interest is that the network facilitates signal transduction from
residues involved in the recognition of GSH to the active site res-
idue Asp-49. This process could facilitate the disruption of its
interaction with Arg-126, facilitating the latter’s role in thiolate
stabilization on a time scale specific to catalysis. Thr-34 and Leu-69,
located on helices I and II, respectively, make hydrophobic contacts
with the γ-glutamyl moiety of GSH and initiate series of correlated
van der Waals overlaps that ultimately affect Asp-49, e.g., Leu-69 →
Thr-34 → Cys-68 → Asp-64 → Lys-41 → Arg-40 → Leu-39 → His-
53 → K42 → H53 → Asp-49 (Fig. 3A and Movie S1).
Conversely, Arg-126, with which it forms an intermonomeric
interaction, is fitted as the single conformation observed within the
crystal structure (13) (Fig. 2A). As discussed above, we believe this
active site geometry is strong evidence of a GSH thiol–Arg-126
interaction. Although the pKa of GSH thiol has been measured as
9.42 in solution (37), the dynamic interaction of GSH thiol with
Arg-126, combined with the solvent restricted electrostatics of the
active site, may allow GSH to transiently form thiolate during ca-
talysis via a mechanism of charge redistribution. Specifically, the
hydrogen-bonding network formed by a crystallographic water
molecule between the α-carboxylate and thiol moieties of GSH
may be crucial in this respect (Figs. 2A and 4).
This finding is corroborated by comparison with the qFit-
generated structural ensembles of the bis-phenyl complex (PDB
ID code 4AL1), in which Arg-126, now with a reduced potential
for interaction with thiol, is observed to be in dynamic motion
(Figs. 2B and 3B) (see below).
Fig. 4. Proposed mechanism of mPGES-1. For details, please see Results and
Discussion.
Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 975
BIOCHEMISTRY
Contact Signaling Is Ligand-Dependent. After an iterative fitting of
alternate conformations with qFit (20), building of N-terminal
residues into density, and subsequent refinement with Phenix
(34) (described in Materials and Methods), a significant im-
provement of quality indicators was achieved for the 1.95-Å
resolution mPGES-1 complex with bisphenyl–GSH (PDB ID
code 4AL1), with a reduction of R/Rfree values from 16.3/17.2%
to 13.7/16.6%, respectively. Subsequent CONTACT analysis
lacked the extensive signaling network found within the GSH
complex, however, which were instead focused on opposing sides
of the active site. The ensemble structure was found to contain
two networks of four and five residues, respectively, the latter of
which occurs in the cytoplasmic loop between helices III and IV
and contains alternate conformations of Arg-126 (Fig. 3B). This
finding suggests that the activation of dynamic contact networks
in mPGES-1 may be dependent upon the identity of the ligand
bound at the active site. Although the difference in structural
information inherent in the two datasets (1.16 Å cf. 1.95 Å)
should be considered when drawing comparisons between the
qFit-generated ensemble structures, the resolution of the 4AL1
dataset is still significantly higher than the upper limit of 2.1 Å
suggested by the software developers (smb.slac.stanford.edu/
qFitServer/) (20). In addition, we performed a qFit/CONTACT
analysis of high-resolution (1.41–1.52 Å) mPGES-1 inhibitor
complexes recently published (15) (PDB ID codes 4YK5, 4YL0,
4YL1, and 4YL3). These four compounds are also observed to
bind in the intermonomeric active site, making extensive contact
with the C-domain. Although the four inhibitors are varied in
structure and binding modes, they all share a common interaction
with the C-domain and lack the extensive networks found in the
holoenzyme complex with GSH (PDB ID code 4AL0). Intriguingly,
the same interaction is also fulfilled by an octyl glucoside (n-octyl-β-D-
glucoside) detergent molecule (not shown in Figs. 2 and 3 for clarity;
cf. Fig. S3) within the bis-phenyl complex (PDB ID code 4AL1),
whose polar head group also makes contact with the turn/helix
C-domain motif and whose hydrophobic tail stacks against the bis-
phenyl moiety of the GSH analog (13) (Fig. S3). This finding indi-
cates that stabilizing contacts within this region may disrupt potential
for signal transduction (Fig. S3). As shown in Fig. 3A and Movie S1,
the dynamic conformations of Lys-41, Arg-40, Leu-39, and His-53
are essential to the transmission of the contact network within the
C-domain, and ultimately to the active site residue, Asp-49. There-
fore, it is possible that their mode of inhibition is mediated by fa-
voring certain conformations of these residues from the structural
ensemble and subsequent interruption of signaling (26).
This mechanism could be a common theme of potent mPGES-1
inhibitors. In a recent analysis of binding sites via mass spectrometry
hydrogen/deuterium exchange experiments (38), the authors found
that the greatest differences common to the two most potent inhibitors
were observed in residues 37–54, corresponding to the C-domain.
Conclusions
The combined results of site-directed mutagenesis, functional
assays, structural ensemble, and contact network analysis pre-
sented herein provide strong evidence for a mechanism of PGE2
synthesis by mPGES-1 that features an activation of GSH thio-
late by Arg-126, modulated via its respective interaction with
Asp-49. Furthermore, we show that conformations of the latter
can be affected by a ligand-dependent contact signaling, con-
necting it to distal residues involved in GSH recognition, with the
potential to dynamically alter the Asp-49–Arg-126 interaction
during catalysis (Fig. 3).
We propose a previously unidentified mechanism of PGH2 isom-
erization by mPGES-1 that features a prominent role of a water-
mediated interaction with the α-carboxylate of GSH and an Asp-
49–mediated thiolate stabilization by Arg-126 (Fig. 4). We hypothesize
that the active site of mPGES-1 lowers the pKa of GSH thiol
and the C-9 proton of PGH2 concurrently via respective interac-
tions with Arg-126 and Asp-49, facilitated by their dynamic con-
formational change in response to contact network signaling.
Charge conservation in this solvent-restricted environment could
thus be achieved via proton shuffling by the crystallographic
water/α-carboxylate hydrogen-bonding network (Fig. 4).
This work has broad implications for the pharmacological efforts
to inhibit this enzyme, which are a current topic of discussion
within the literature (39).
Materials and Methods
Protein Expression and Purification. Recombinant wild-type (WT) and active-site
mutants of human mPGES-1 were overexpressed in P. pastoris and purified by
Ni-affinity chromatography before exchanging buffer to 0.1 M phosphate
buffer, 0.03% dodecyl maltoside, and 2.5 mM GSH, pH 7.4. Microsomal prep-
arations were prepared via ultracentrifugation of lysed cell supernatant and
homogenization of the microsomal pellets in assay buffer (20 mM Tris·HCl, pH
7.8, 2.5 mM GSH). For further details, please refer to SI Materials and Methods.
Synthesis of GOH. The oxygen analog of GSH, GOH, was synthesized in a three-
step procedure based on a published method (40). For further details, please
see SI Materials and Methods.
Enzyme Activity Assay. Conversion of PGH2 to PGE2 by WT or mutated
mPGES-1 were quantified by using GC-MS as described (12). For further
details, please refer to SI Materials and Methods.
Analysis of Dynamic Contact Networks. The qFit Web server (smb.slac.
stanford.edu/qFitServer/) and the CONTACT algorithm (27) were used for
the quantification of conformational ensembles and functional contact
networks, respectively. Before analysis, the physiological trimer was gener-
ated from the asymmetric unit via crystallographic symmetry using the
program COOT (41). For PDB ID codes 4AL0, 4YL0, 4YL1, 4YL3, and 4YK5, the
coordinates were submitted to the qFit server and refined and prepared for
CONTACT as described (27), by using Phenix-1.9-1692 (34) without manual
intervention. For PDB ID code 4AL1, significant density improvement at the
amino terminus allowed residues 4–9 to be built into density after qFit
conformer fitting and refinement. After a second round of refinement, the
resulting improvement in quality indicators such as the Rfree value were
significant, such that the improved phase estimates were anticipated to
affect the conformational ensemble fitting. Hence, the improved coordi-
nates were resubmitted to the qFit server before being refined and pre-
pared for analysis with CONTACT as above. Settings for all CONTACT
analyses were as follows: Tstress (percentile) = 0.4, max_path_length = 100,
sc_only_flag = f (all atom), relief_threshold = 0.90.
ACKNOWLEDGMENTS. We thank Gunvor Hamberg for technical assistance
and gratefully acknowledge the late Richard Armstrong, who provided the
GOH GSH analogue. Part of this work was performed at the Karolinska
Institutet Protein Science Facility. Some computations were performed on
resources provided by the Swedish National Infrastructure for Computing
at Linköping University. This work was supported by Swedish Research Coun-
cil Grant 10350 and CERIC Linnaeus Grant; the Stockholm County Council
(Cardiovascular Program, Thematic Center Inflammation); and NovoNordisk
Foundation Grant NNF15CC0018346. J.Z.H. is the recipient of a Distinguished
Professor Award from Karolinska Institutet.
1. Samuelsson B, et al. (1978) Prostaglandins and thromboxanes. Annu Rev Biochem
47(1):997–1029.
2. Jakobsson PJ, Thorén S, Morgenstern R, Samuelsson B (1999) Identification of human
prostaglandin E synthase: A microsomal, glutathione-dependent, inducible enzyme,
constituting a potential novel drug target. Proc Natl Acad Sci USA 96(13):7220–7225.
3. Watanabe K, Kurihara K, Suzuki T (1999) Purification and characterization of mem-
brane-bound prostaglandin E synthase from bovine heart. Biochim Biophys Acta
1439(3):406–414.
4. Murakami M, et al. (2000) Regulation of prostaglandin E2 biosynthesis by inducible
membrane-associated prostaglandin E2 synthase that acts in concert with cyclo-
oxygenase-2. J Biol Chem 275(42):32783–32792.
5. Samuelsson B, Morgenstern R, Jakobsson P-J (2007) Membrane prostaglandin E syn-
thase-1: A novel therapeutic target. Pharmacol Rev 59(3):207–224.
6. Jakobsson PJ, Morgenstern R, Mancini J, Ford-Hutchinson A, Persson B (2000) Mem-
brane-associated proteins in eicosanoid and glutathione metabolism (MAPEG). A
widespread protein superfamily. Am J Respir Crit Care Med 161(2 Pt 2):S20–S24.
976 | www.pnas.org/cgi/doi/10.1073/pnas.1522891113 Brock et al.
7. Ferguson AD, et al. (2007) Crystal structure of inhibitor-bound human 5-lipoxygenase-
activating protein. Science 317(5837):510–512.
8. Ago H, et al. (2007) Crystal structure of a human membrane protein involved in
cysteinyl leukotriene biosynthesis. Nature 448(7153):609–612.
9. Martinez Molina D, et al. (2007) Structural basis for synthesis of inflammatory me-
diators by human leukotriene C4 synthase. Nature 448(7153):613–616.
10. Rinaldo-Matthis A, et al. (2010) Arginine 104 is a key catalytic residue in leukotriene
C4 synthase. J Biol Chem 285(52):40771–40776.
11. Saino H, et al. (2011) The catalytic architecture of leukotriene C4 synthase with two
arginine residues. J Biol Chem 286(18):16392–16401.
12. Hammarberg T, et al. (2009) Mutation of a critical arginine in microsomal prosta-
glandin E synthase-1 shifts the isomerase activity to a reductase activity that converts
prostaglandin H2 into prostaglandin F2alpha. J Biol Chem 284(1):301–305.
13. Sjögren T, et al. (2013) Crystal structure of microsomal prostaglandin E2 synthase
provides insight into diversity in the MAPEG superfamily. Proc Natl Acad Sci USA
110(10):3806–3811.
14. Li D, et al. (2014) Crystallizing membrane proteins in the lipidic mesophase. experi-
ence with human prostaglandin E2 synthase 1 and an evolving strategy. Cryst Growth
Des 14(4):2034–2047.
15. Luz JG, et al. (2015) Crystal structures of mPGES-1 inhibitor complexes form a basis for
the rational design of potent analgesic and anti-inflammatory therapeutics. J Med
Chem 58(11):4727–4737.
16. Berman HM, et al. (2000) The Protein Data Bank. Nucleic Acids Res 28(1):235–242.
17. Kuzmanic A, Pannu NS, Zagrovic B (2014) X-ray refinement significantly underesti-
mates the level of microscopic heterogeneity in biomolecular crystals. Nat Commun
5:3220.
18. Lang PT, Holton JM, Fraser JS, Alber T (2014) Protein structural ensembles are re-
vealed by redefining X-ray electron density noise. Proc Natl Acad Sci USA 111(1):
237–242.
19. Burnley BT, et al. (2012) Modelling dynamics in protein crystal structures by ensemble
refinement. eLife 1:e00311.
20. Fraser JS, et al. (2011) Accessing protein conformational ensembles using room-
temperature X-ray crystallography. Proc Natl Acad Sci USA 108(39):16247–16252.
21. Lang PT, et al. (2010) Automated electron-density sampling reveals widespread
conformational polymorphism in proteins. Protein Sci 19(7):1420–1431.
22. Henzler-Wildman KA, et al. (2007) Intrinsic motions along an enzymatic reaction
trajectory. Nature 450(7171):838–844.
23. Yang L-Q, et al. (2014) Protein dynamics and motions in relation to their functions:
Several case studies and the underlying mechanisms. J Biomol Struct Dyn 32(3):
372–393.
24. Fraser JS, et al. (2009) Hidden alternative structures of proline isomerase essential for
catalysis. Nature 462(7273):669–673.
25. Klinman JP (2015) Dynamically achieved active site precision in enzyme catalysis. Acc
Chem Res 48(2):449–456.
26. Fischer M, Coleman RG, Fraser JS, Shoichet BK (2014) Incorporation of protein flexi-
bility and conformational energy penalties in docking screens to improve ligand
discovery. Nat Chem 6(7):575–583.
27. van den Bedem H, Bhabha G, Yang K, Wright PE, Fraser JS (2013) Automated iden-
tification of functional dynamic contact networks from X-ray crystallography. Nat
Methods 10(9):896–902.
28. Ouellet M, et al. (2002) Purification and characterization of recombinant microsomal
prostaglandin E synthase-1. Protein Expr Purif 26(3):489–495.
29. Oakley A (2011) Glutathione transferases: A structural perspective. Drug Metab Rev
43(2):138–151.
30. Zhou P, Tian F, Lv F, Shang Z (2009) Geometric characteristics of hydrogen bonds
involving sulfur atoms in proteins. Proteins 76(1):151–163.
31. Donald JE, Kulp DW, DeGrado WF (2011) Salt bridges: Geometrically specific, de-
signable interactions. Proteins 79(3):898–915.
32. Winayanuwattikun P, Ketterman AJ (2005) An electron-sharing network involved in
the catalytic mechanism is functionally conserved in different glutathione transferase
classes. J Biol Chem 280(36):31776–31782.
33. Dourado D, Fernandes P, Mannervik B, Ramos M (2008) Glutathione transferase: New
model for glutathione activation. Chemistry 14(31):9591–9598.
34. Adams P, et al. (2010) PHENIX: A comprehensive Python-based system for macro-
molecular structure solution. Acta Crystallogr D Biol Crystallogr 66(2):213–221.
35. Yamaguchi N, et al. (2011) Theoretical studies on model reaction pathways of pros-
taglandin H2 isomerization to prostaglandin D2/E2. Theor Chem Acc 128(2):191–206.
36. Li Y, Angelastro M, Shimshock S, Reiling S, Vaz RJ (2010) On the mechanism of mi-
crosomal prostaglandin E synthase type-2—A theoretical study of endoperoxide re-
action with MeS−. Bioorg Med Chem 20(1):338–340.
37. Tajc SG, Tolbert BS, Basavappa R, Miller BL (2004) Direct determination of thiol pKa by
isothermal titration microcalorimetry. J Am Chem Soc 126(34):10508–10509.
38. Prage EB, et al. (2011) Location of inhibitor binding sites in the human inducible
prostaglandin E synthase, MPGES1. Biochemistry 50(35):7684–7693.
39. Koeberle A, Werz O (2015) Perspective of microsomal prostaglandin E 2 synthase-1 as
drug target in inflammation-related disorders. Biochem Pharmacol 98(1):1–15.
40. Chen WJ, Boehlert CC, Rider K, Armstrong RN (1985) Synthesis and characterization of
the oxygen and desthio analogues of glutathione as dead-end inhibitors of gluta-
thione S-transferase. Biochem Biophys Res Commun 128(1):233–240.
41. Emsley P, Cowtan K (2004) Coot: Model-building tools for molecular graphics. Acta
Crystallogr D Biol Crystallogr 60(Pt 12 Pt 1):2126–2132.
42. Humphrey W, Dalke A, Schulten K (1996) VMD: Visual molecular dynamics. J Mol
Graph 14:33–38, 27–28.
43. Phillips JC, et al. (2005) Scalable molecular dynamics with NAMD. J Comput Chem
26(16):1781–1802.
44. Lomize MA, Lomize AL, Pogozheva ID, Mosberg HI (2006) OPM: Orientations of
proteins in membranes database. Bioinformatics 22(5):623–625.
Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 977
BIOCHEMISTRY

More Related Content

What's hot

Skv Enzyme Kinetics and Principles of Enzyme Inhibition
Skv Enzyme Kinetics and Principles of Enzyme InhibitionSkv Enzyme Kinetics and Principles of Enzyme Inhibition
Skv Enzyme Kinetics and Principles of Enzyme InhibitionSACHINKUMARVISHWAKAR4
 
Enzyme kinetics - Creative Enzymes
Enzyme kinetics - Creative EnzymesEnzyme kinetics - Creative Enzymes
Enzyme kinetics - Creative EnzymesCreative Enzymes
 
Inhibition of RelA-Mediated Biofilm Synthesis
Inhibition of RelA-Mediated Biofilm SynthesisInhibition of RelA-Mediated Biofilm Synthesis
Inhibition of RelA-Mediated Biofilm SynthesisJohn Cahill
 
Enzyme Inhibitors, Activation energy and MEC.
Enzyme Inhibitors, Activation energy and MEC.Enzyme Inhibitors, Activation energy and MEC.
Enzyme Inhibitors, Activation energy and MEC.pratham4012
 
Chapter 1 the physical basis of ligand binding
Chapter 1 the physical basis of ligand bindingChapter 1 the physical basis of ligand binding
Chapter 1 the physical basis of ligand bindingBenediktusMadika1
 
Enzyme kinetics and inhibition
Enzyme kinetics and inhibitionEnzyme kinetics and inhibition
Enzyme kinetics and inhibitionNRx Hemant Rathod
 
Lead Optimization of Macrolide drug
Lead Optimization of Macrolide drugLead Optimization of Macrolide drug
Lead Optimization of Macrolide drugsaurabh gupta
 
Enzyme kinetics and catalysis
Enzyme kinetics and catalysisEnzyme kinetics and catalysis
Enzyme kinetics and catalysisDr.M.Prasad Naidu
 
Pai 1 inhibitors
Pai 1 inhibitorsPai 1 inhibitors
Pai 1 inhibitorsHitesh Soni
 
Mechanisms action of enzymes
Mechanisms action of enzymesMechanisms action of enzymes
Mechanisms action of enzymesGhazwan Faisal
 
FEBS Letters 2007 Tang
FEBS Letters 2007 TangFEBS Letters 2007 Tang
FEBS Letters 2007 TangPingtao Tang
 
ThrRS Project Summary
ThrRS Project SummaryThrRS Project Summary
ThrRS Project SummaryShawn Egri
 
Building kinetic model of Trehalose
Building kinetic model of TrehaloseBuilding kinetic model of Trehalose
Building kinetic model of TrehaloseHazem Hussein
 
Pancetti et al 2004 methilene
Pancetti et al 2004 methilenePancetti et al 2004 methilene
Pancetti et al 2004 methileneJorge Parodi
 
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric Sunita Sangwan
 
Enzymes powerpoint for 27th may y12 bio
Enzymes powerpoint for 27th may y12 bioEnzymes powerpoint for 27th may y12 bio
Enzymes powerpoint for 27th may y12 bioRitchistep
 

What's hot (20)

Skv Enzyme Kinetics and Principles of Enzyme Inhibition
Skv Enzyme Kinetics and Principles of Enzyme InhibitionSkv Enzyme Kinetics and Principles of Enzyme Inhibition
Skv Enzyme Kinetics and Principles of Enzyme Inhibition
 
Enzyme kinetics - Creative Enzymes
Enzyme kinetics - Creative EnzymesEnzyme kinetics - Creative Enzymes
Enzyme kinetics - Creative Enzymes
 
Inhibition of RelA-Mediated Biofilm Synthesis
Inhibition of RelA-Mediated Biofilm SynthesisInhibition of RelA-Mediated Biofilm Synthesis
Inhibition of RelA-Mediated Biofilm Synthesis
 
Enzyme kinetics
Enzyme kineticsEnzyme kinetics
Enzyme kinetics
 
Enzyme Inhibitors, Activation energy and MEC.
Enzyme Inhibitors, Activation energy and MEC.Enzyme Inhibitors, Activation energy and MEC.
Enzyme Inhibitors, Activation energy and MEC.
 
Chapter 1 the physical basis of ligand binding
Chapter 1 the physical basis of ligand bindingChapter 1 the physical basis of ligand binding
Chapter 1 the physical basis of ligand binding
 
Enzyme kinetics and inhibition
Enzyme kinetics and inhibitionEnzyme kinetics and inhibition
Enzyme kinetics and inhibition
 
Lead Optimization of Macrolide drug
Lead Optimization of Macrolide drugLead Optimization of Macrolide drug
Lead Optimization of Macrolide drug
 
Enzyme kinetics and catalysis
Enzyme kinetics and catalysisEnzyme kinetics and catalysis
Enzyme kinetics and catalysis
 
Pai 1 inhibitors
Pai 1 inhibitorsPai 1 inhibitors
Pai 1 inhibitors
 
Mechanisms action of enzymes
Mechanisms action of enzymesMechanisms action of enzymes
Mechanisms action of enzymes
 
FEBS Letters 2007 Tang
FEBS Letters 2007 TangFEBS Letters 2007 Tang
FEBS Letters 2007 Tang
 
ThrRS Project Summary
ThrRS Project SummaryThrRS Project Summary
ThrRS Project Summary
 
Building kinetic model of Trehalose
Building kinetic model of TrehaloseBuilding kinetic model of Trehalose
Building kinetic model of Trehalose
 
Enzyme
EnzymeEnzyme
Enzyme
 
Pancetti et al 2004 methilene
Pancetti et al 2004 methilenePancetti et al 2004 methilene
Pancetti et al 2004 methilene
 
Enzymes
EnzymesEnzymes
Enzymes
 
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric
Enzyme inhibition - Competitive, Non- Competitive, Uncompetitive, Allosteric
 
Enzymes powerpoint for 27th may y12 bio
Enzymes powerpoint for 27th may y12 bioEnzymes powerpoint for 27th may y12 bio
Enzymes powerpoint for 27th may y12 bio
 
Sept 5 bt202
Sept 5 bt202Sept 5 bt202
Sept 5 bt202
 

Viewers also liked

Laurens maakt werk van de geriatrische revalidatie definitief
Laurens maakt werk van de geriatrische revalidatie definitiefLaurens maakt werk van de geriatrische revalidatie definitief
Laurens maakt werk van de geriatrische revalidatie definitiefHans Stravers
 
Actividad de Aprendizaje 8,
Actividad de Aprendizaje 8,Actividad de Aprendizaje 8,
Actividad de Aprendizaje 8,DESIGN ME
 
Proceso s y e de leavell y clark
Proceso s y e de leavell y clarkProceso s y e de leavell y clark
Proceso s y e de leavell y clarkRochy Montenegro
 
'Blue Ocean Strategy' book review
'Blue Ocean Strategy' book review'Blue Ocean Strategy' book review
'Blue Ocean Strategy' book reviewWalid Saafan
 
Preguntas de investigación
Preguntas de investigaciónPreguntas de investigación
Preguntas de investigaciónSara Alarcón
 
Wikipedia Health Landscape - Top Health Conditions
Wikipedia Health Landscape - Top Health ConditionsWikipedia Health Landscape - Top Health Conditions
Wikipedia Health Landscape - Top Health ConditionsGary Monk
 
不傷脊椎的標準姿勢
不傷脊椎的標準姿勢不傷脊椎的標準姿勢
不傷脊椎的標準姿勢lys167
 
Anexo 3° bloque i
Anexo 3° bloque iAnexo 3° bloque i
Anexo 3° bloque iMary Napu
 

Viewers also liked (14)

StudyCopterBot
StudyCopterBotStudyCopterBot
StudyCopterBot
 
Laurens maakt werk van de geriatrische revalidatie definitief
Laurens maakt werk van de geriatrische revalidatie definitiefLaurens maakt werk van de geriatrische revalidatie definitief
Laurens maakt werk van de geriatrische revalidatie definitief
 
29253ip
29253ip29253ip
29253ip
 
Apresentacao casa x 2016
Apresentacao casa x 2016Apresentacao casa x 2016
Apresentacao casa x 2016
 
Actividad de Aprendizaje 8,
Actividad de Aprendizaje 8,Actividad de Aprendizaje 8,
Actividad de Aprendizaje 8,
 
DECOY-7 Slidedeck
DECOY-7 SlidedeckDECOY-7 Slidedeck
DECOY-7 Slidedeck
 
Proceso s y e de leavell y clark
Proceso s y e de leavell y clarkProceso s y e de leavell y clark
Proceso s y e de leavell y clark
 
'Blue Ocean Strategy' book review
'Blue Ocean Strategy' book review'Blue Ocean Strategy' book review
'Blue Ocean Strategy' book review
 
Preguntas de investigación
Preguntas de investigaciónPreguntas de investigación
Preguntas de investigación
 
Wikipedia Health Landscape - Top Health Conditions
Wikipedia Health Landscape - Top Health ConditionsWikipedia Health Landscape - Top Health Conditions
Wikipedia Health Landscape - Top Health Conditions
 
PROJECTE AIGUA 1r
PROJECTE AIGUA 1rPROJECTE AIGUA 1r
PROJECTE AIGUA 1r
 
不傷脊椎的標準姿勢
不傷脊椎的標準姿勢不傷脊椎的標準姿勢
不傷脊椎的標準姿勢
 
Anexo 3° bloque i
Anexo 3° bloque iAnexo 3° bloque i
Anexo 3° bloque i
 
Tema 1. bioetica
Tema 1. bioeticaTema 1. bioetica
Tema 1. bioetica
 

Similar to Dynamic Asp-Arg interaction essential for catalysis in microsomal prostaglandin E2 synthase

RAGE-Mediated Cell Signaling.pdf
RAGE-Mediated Cell Signaling.pdfRAGE-Mediated Cell Signaling.pdf
RAGE-Mediated Cell Signaling.pdfNazmunNahar479158
 
MasterLS_BMS_2013_MSOvergaauw
MasterLS_BMS_2013_MSOvergaauwMasterLS_BMS_2013_MSOvergaauw
MasterLS_BMS_2013_MSOvergaauwMaarten Overgaauw
 
PIIS1552526009012771(1)(3)
PIIS1552526009012771(1)(3)PIIS1552526009012771(1)(3)
PIIS1552526009012771(1)(3)Marco Garza
 
Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006Alex Kiselyov
 
Presentazione ii anno campana
Presentazione ii anno campanaPresentazione ii anno campana
Presentazione ii anno campanalab13unisa
 
1-s2.0-S0167488911000772-main
1-s2.0-S0167488911000772-main1-s2.0-S0167488911000772-main
1-s2.0-S0167488911000772-mainAndreas D. Song
 
Duchenne drug tested for muscular dystrophy as drug repurposing
Duchenne drug tested for muscular dystrophy as drug repurposingDuchenne drug tested for muscular dystrophy as drug repurposing
Duchenne drug tested for muscular dystrophy as drug repurposingmovvaharshavardhan
 
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...Gerhard Koenig
 
Design and synthesis of novel protein kinase R (PKR) inhibitors
Design and synthesis of novel protein kinase R (PKR) inhibitorsDesign and synthesis of novel protein kinase R (PKR) inhibitors
Design and synthesis of novel protein kinase R (PKR) inhibitorssagiv weintraub
 
Receptor pharmacology Uttam & Renoo
Receptor pharmacology Uttam & RenooReceptor pharmacology Uttam & Renoo
Receptor pharmacology Uttam & Renoouttam singh
 
Environmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthEnvironmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthXunhai 郑训海
 
Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Sana Hafeez
 
Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Sana Hafeez
 

Similar to Dynamic Asp-Arg interaction essential for catalysis in microsomal prostaglandin E2 synthase (20)

J. Biol. Chem.-2015-Maganti-9812-22
J. Biol. Chem.-2015-Maganti-9812-22J. Biol. Chem.-2015-Maganti-9812-22
J. Biol. Chem.-2015-Maganti-9812-22
 
RAGE-Mediated Cell Signaling.pdf
RAGE-Mediated Cell Signaling.pdfRAGE-Mediated Cell Signaling.pdf
RAGE-Mediated Cell Signaling.pdf
 
MasterLS_BMS_2013_MSOvergaauw
MasterLS_BMS_2013_MSOvergaauwMasterLS_BMS_2013_MSOvergaauw
MasterLS_BMS_2013_MSOvergaauw
 
PIIS1552526009012771(1)(3)
PIIS1552526009012771(1)(3)PIIS1552526009012771(1)(3)
PIIS1552526009012771(1)(3)
 
Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006Nat_Chem_Biol_GPR30_2006
Nat_Chem_Biol_GPR30_2006
 
Presentazione ii anno campana
Presentazione ii anno campanaPresentazione ii anno campana
Presentazione ii anno campana
 
1-s2.0-S0167488911000772-main
1-s2.0-S0167488911000772-main1-s2.0-S0167488911000772-main
1-s2.0-S0167488911000772-main
 
D sc peptide1
D sc peptide1D sc peptide1
D sc peptide1
 
Duchenne drug tested for muscular dystrophy as drug repurposing
Duchenne drug tested for muscular dystrophy as drug repurposingDuchenne drug tested for muscular dystrophy as drug repurposing
Duchenne drug tested for muscular dystrophy as drug repurposing
 
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...
Blain_AlzResTherapy_2016_Characterization of FRM-36143 as a new γ-secretase m...
 
Design and synthesis of novel protein kinase R (PKR) inhibitors
Design and synthesis of novel protein kinase R (PKR) inhibitorsDesign and synthesis of novel protein kinase R (PKR) inhibitors
Design and synthesis of novel protein kinase R (PKR) inhibitors
 
a-FMH Poster
a-FMH Postera-FMH Poster
a-FMH Poster
 
Receptor pharmacology Uttam & Renoo
Receptor pharmacology Uttam & RenooReceptor pharmacology Uttam & Renoo
Receptor pharmacology Uttam & Renoo
 
ACS Poster
ACS PosterACS Poster
ACS Poster
 
Environmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the monthEnvironmental Factor - July 2014_ Intramural papers of the month
Environmental Factor - July 2014_ Intramural papers of the month
 
ncomms7387
ncomms7387ncomms7387
ncomms7387
 
ijms-24-07373.pdf
ijms-24-07373.pdfijms-24-07373.pdf
ijms-24-07373.pdf
 
Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)
 
Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)Phosphoglycerate_mutase (1)
Phosphoglycerate_mutase (1)
 
LSD1 - bmc-paper
LSD1 - bmc-paperLSD1 - bmc-paper
LSD1 - bmc-paper
 

Dynamic Asp-Arg interaction essential for catalysis in microsomal prostaglandin E2 synthase

  • 1. A dynamic Asp–Arg interaction is essential for catalysis in microsomal prostaglandin E2 synthase Joseph S. Brocka,1 , Mats Hamberga,1 , Navisraj Balagunaseelana , Michael Goodmanb , Ralf Morgensternc , Emilia Strandbacka , Bengt Samuelssona,2 , Agnes Rinaldo-Matthisa , and Jesper Z. Haeggströma,2 a Department of Medical Biochemistry and Biophysics, Karolinska Institutet, S-171 77 Stockholm, Sweden; b Department of Chemistry, Vanderbilt University School of Medicine, Nashville, TN 37232-6304; and c Institute of Environmental Medicine, Karolinska Institutet, S-171 77 Stockholm, Sweden Contributed by Bengt Samuelsson, December 18, 2015 (sent for review October 9, 2015; reviewed by Lawrence J. Marnett, Charles N. Serhan, and Takao Shimizu) Microsomal prostaglandin E2 synthase type 1 (mPGES-1) is responsible for the formation of the potent lipid mediator prostaglandin E2 under proinflammatory conditions, and this enzyme has received consider- able attention as a drug target. Recently, a high-resolution crystal struc- ture of human mPGES-1 was presented, with Ser-127 being proposed as the hydrogen-bond donor stabilizing thiolate anion formation within the cofactor, glutathione (GSH). We have combined site-directed mutagenesis and activity assays with a structural dynamics analysis to probe the functional roles of such putative catalytic residues. We found that Ser-127 is not required for activity, whereas an interaction be- tween Arg-126 and Asp-49 is essential for catalysis. We postulate that both residues, in addition to a crystallographic water, serve critical roles within the enzymatic mechanism. After characterizing the size or charge conservative mutations Arg-126–Gln, Asp-49–Asn, and Arg- 126–Lys, we inferred that a crystallographic water acts as a general base during GSH thiolate formation, stabilized by interaction with Arg-126, which is itself modulated by its respective interaction with Asp-49. We subsequently found hidden conformational ensembles within the crystal structure that correlate well with our biochemical data. The resulting contact signaling network connects Asp-49 to distal residues involved in GSH binding and is ligand dependent. Our work has broad implications for development of efficient mPGES-1 inhibitors, potential anti-inflammatory and anticancer agents. inflammation | prostaglandin | mPGES-1 | MAPEG | mechanism Prostaglandin E2 (PGE2) is an abundant lipid mediator that signals via four receptors (EP1–4) to induce an array of important biological actions in physiology as well as pathophysiology (1). Under proinflammatory conditions, biosynthesis of PGE2 proceeds from arachidonic acid, which is converted to the unstable endoper- oxide PGH2 by cyclooxygenase type 2 (COX-2). PGH2 is further isomerized into PGE2 by microsomal PGE synthase type 1 (mPGES- 1) (2, 3). mPGES-1 is encoded by PTGES and is up-regulated by mitogens and cytokines in a pathway that is functionally coupled to COX-2 (2, 4). Because of its key role in PGE2 synthesis, mPGES-1 has attracted attention as a potential drug target in the areas of inflammation, pain, fever, and cancer (5). mPGES-1 is a member of the MAPEG (Membrane-Associated Proteins in Eicosanoid and Glutathione metabolism) superfamily of enzymes (6), which also includes two key proteins in the leukotriene (LT) cascade, viz. 5-lipoxygenase activating protein and LT C4 syn- thase (LTC4S). All MAPEG members are integral, homotrimeric membrane proteins, and structural information on this family has been scarce. However, significant progress has recently been made in this area with several high-resolution structures being solved by X-ray crystallography (7–9). In particular, the crystal structures of human LTC4S provided detailed structural information, including an argi- nine residue that was later shown to activate the glutathione (GSH) thiolate (10, 11). We have previously proposed that this conserved arginine residue is also essential for enzymatic activity in mPGES-1 as Arg-126 (12). The recent structural determination of mPGES-1, however, at an exceptionally high resolution of 1.16 Å, uncovered several unanticipated structural features (13). The active sites, found at the three monomeric interfaces, show that Ser-127 is positioned near the GSH thiol group, indicating that it may act as a hydrogen- bond donor to assist in thiolate formation during catalysis. Further- more, Arg-126 and Asp-49 participate in a charge interaction that could also contribute to catalysis. This structural information is supported by a mesophase crystal structure of an engineered version of mPGES-1 (14) and several inhibitor complexes that have recently been published (15). Here, we initially confirmed the necessity for GSH thiolate during catalysis via incubations with the analogous tripeptide γ-Glu–Ser–Gly (GOH). We then used site-directed mutagenesis to analyze the functional roles of active site residues. We found that Ser-127 is nonessential for catalysis, whereas Arg-126 and Asp-49 are crucial and mutually dependent for native isomerase activity of the enzyme. Because the latter codependence of activity could be rationalized by a dynamic functional role of these residues, we turned to the high- resolution X-ray data (13, 15) deposited in the Protein Data Bank (PDB; www.rcsb.org) (16) to provide evidence of their dynamic motion within the crystal structure. Several recent studies have shown that the information present in such data is often under- estimated and that it is possible to refine multiple conformations of residues simultaneously, each with individually refined occupancy and B factors, without overfitting the data (17–22). By such sampling of low-level electron density, discrete, “hidden” conformations are revealed, facilitating a more quantitative representation of dynamic Significance Microsomal prostaglandin E2 synthase type 1 (mPGES-1) is an integral membrane protein that produces prostaglandin E2 (PGE2), a mediator of inflammation, fever, pain, and tumorigenesis. Here we show that a serine residue implicated by the crystal structure is not required for function, whereas an arginine and aspartate residue in the active site, observed to be interacting within the crystal structure, are essential and mutually dependent during catalysis. We also demonstrate that a contact signaling network can interrupt the arginine–asparagine interaction and facilitate their participation in the chemical mechanism. Our work has broad implications for development of effective mPGES-1 inhib- itors, potential drugs with clinical application in treatment of in- flammatory diseases and cancer. Author contributions: M.H., B.S., A.R.-M., and J.Z.H. designed research; J.S.B., M.H., N.B., and E.S. performed research; M.G. and R.M. contributed new reagents/analytic tools; J.S.B., M.H., N.B., A.R.-M., and J.Z.H. analyzed data; and J.S.B. and J.Z.H. wrote the paper. Reviewers: L.J.M., Vanderbilt University Medical Center; C.N.S., Brigham and Women’s Hospital/Harvard Medical School; and T.S., National Center for Global Health and Medicine. The authors declare no conflict of interest. Freely available online through the PNAS open access option. 1 J.S.B. and M.H. contributed equally to this work. 2 To whom correspondence may be addressed. Email: jesper.haeggstrom@ki.se or bengt. samuelsson@ki.se. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1522891113/-/DCSupplemental. 972–977 | PNAS | January 26, 2016 | vol. 113 | no. 4 www.pnas.org/cgi/doi/10.1073/pnas.1522891113
  • 2. motion within the crystal lattice. Furthermore, it has been shown that this information is often essential for understanding enzymatic function (23) and mechanism (24, 25) and successfully achieving structure-based drug design (26). We quantified the dynamic conformations of active site residues using the software qFit (20) and CONTACT (27). This method generated significant improvements in the quality indicators for PDB ID codes 4AL0 and 4AL1 [1.16 and 1.95 Å, respectively (13)] and revealed a ligand-dependent contact network that corroborates the mechanism suggested from biochemical data. These van der Waals interactions within the binary complex with GSH (PDB ID code 4AL0) reveal an extensive network of correlated side-chain motions within the cytoplasmic “C-domain” that forms the bottom of the active site and confirm a dynamic role of Asp-49 in catalysis. In comparison, the much smaller networks found in PDB ID code 4AL1, with a bisphenyl–GSH analog and detergent molecule bound in the active site, indicate that ligand binding can influence network signaling. This finding is also supported by our analysis of recently published inhibitor complexes (15). Our results suggest that the positively charged Arg-126 stabilizes transient thiolate formation and that its dynamic interaction with Asp-49 is essential for catalysis. We also observed a crystallo- graphic water molecule that is ideally situated to act as a proton acceptor during this process. Furthermore, we found a striking contact signaling network within the active site that effects the conformation of residues in a ligand-dependent fashion. Results and Discussion GSH Thiolate Is Essential for Catalysis. Fig. 1 depicts a substrate- limited assay that measured absolute product formation (nano- grams) as described in Materials and Methods. Incubations with native enzyme in the presence of GSH resulted in near total con- version of PGH2 to PGE2 and are concurrent with the specific activity previously reported in the literature (∼4 μmol·min−1 ·mg−1 at 0 °C) (28). A negative control involved microsomal preparations of WT mPGES-1 resuspended in buffer containing the GSH analog, GOH, that differs from the native cofactor only by the replacement of the thiol moiety by a hydroxyl group and did not produce product above background levels. This finding provides strong evidence that thiolate anion is the chemical species of the cofactor essential for catalysis. Ser-127 Is Nonessential for Catalysis. Judging from the orientation of Ser-127 in the recently published crystal structure (13), the authors’ hypothesis that its hydroxyl group may act as a hydrogen- bond donor to stabilize a GSH thiolate is apt, because this mech- anism of thiol activation is a common theme within the evolution of soluble GSH transferases (29). However, because we had pre- viously proposed Arg-126 as a strong candidate for this role (12), we investigated the function of Ser-127 in the conversion of PGH2 into PGE2. To detail the role of Ser-127 in mPGES-1, we exchanged this residue for an alanine by site-directed mutagenesis. After expres- sion in Pichia pastoris and purification, aliquots of recombinant protein were incubated with PGH2. Formation of PGE2 was ana- lyzed by GC-MS. The combined measurements obtained from at least three different preparations of enzyme are depicted in Fig. 1 and show that Ser-127–Ala exhibits the same level of PGE2 syn- thase activity as WT mPGES-1. This finding was true for both purified and microsomal preparations of the enzyme (Fig. 1). In addition, the dual conformations observed for this residue in the Fig. 1. Mutagenic analysis of mPGES-1 active site residues. Aliquots of WT, S127A, D49N, R126Q, and R126K mPGES-1 were incubated with 12 μM PGH2 and analyzed for PGE2 formation by GC-MS, as described in Materials and Methods. The total amount (nanograms) of PGE2 formed is shown from both purified and microsomal preparations of enzyme. PGE2 formation was also monitored for microsomal preparations of WT enzyme incubated with GOH. Values represent the combined measurements from at least three different preparations of en- zyme (n = 3), with error bars representing their SD. The levels of PGF2α formed were also measured by this method as shown in Fig. S1. Fig. 2. The active site architecture of mPGES-1. The active sites of PDB ID codes 4AL0 (A) and 4AL1 (B) are compared with post-qFit conformational fitting and re- finement. The coordinates of PDB depositions are overlaid in translucent over the refined qFit ensembles shown as opaque conformers in stick representation. The 2mFo- DFc electron density corresponding to a mechanistically relevant solvent molecule is shown as blue mesh contoured at 1 rmsd, and the rotation plane of the R126 guanidinium is shown relative to the carboxylate of D49 (44.7°) (A). This water is absent within the qFit-refined bis-phenyl complex (B), potentially due to a reduced capacity for GSH thiol interaction. A molecule of octyl glucoside bound at the C-domain and low occupancy GSH (∼13%) within the active site of B have been omitted for clarity (cf. Fig. S3). Polar interactions are shown as dashed lines with distances given in Å. Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 973 BIOCHEMISTRY
  • 3. crystal structure indicate the absence of a strong hydrogen-bonding interaction. Conversely, Arg-126 is observed in a single confor- mation with an Nη-GSH thiol distance of 3.4 Å. We believe that this active site geometry also substantiates strong evidence for a mechanism of GSH thiol activation by an Arg-126 guanidinium interaction (30). Hence, despite compelling structural evidence, Ser-127 does not play a critical role in mPGES-1 catalysis. Mutation of Arg-126 and Asp-49 Compromises PGE2 Synthase Activity, but Allows PGH2 Reduction to PGF2α. In light of the new structural data (13), we wanted to reexamine the functional role of Arg-126 and mutated this residue into both a glutamine and a lysine residue using site-directed mutagenesis. According to the crystal structure of mPGES-1 (13), Arg-126 and Asp-49 participate in an inter- monomeric charge interaction. Therefore, we also mutated the negatively charged counterpart, Asp-49, into an asparagine residue. We anticipated that the size and charge conservative mutations of these residues could serve in probing their role in the enzymatic mechanism, while minimizing steric and electrostatic repulsion effects, such as disruption of the monomer interface. Although we also attempted to create the charge conservative mutant Asp-49– Glu, the resulting transformed construct failed to express, pre- sumably because it resulted in an unstable quaternary structure. After solubilization with detergent and purification via Ni-affinity chromatography, these mutants were assayed for PGE2 synthase activity as described above. For three different purifications of each isoform, we found that the mutated enzymes did not convert PGH2 into PGE2 above background levels. After preparations of micro- somal fractions, however, we found that the charge conservative mutation Arg-126–Lys still retained a low level of isomerase activity, indicating that a native membrane environment and a formal posi- tive charge at position 126 are important factors for catalysis (Fig. 1). From these results, we conclude that both Arg-126 and Asp-49 are key to the PGE2 synthase activity of mPGES-1. That both of these residues are essential for catalysis is intriguing, because one could expect Arg-126 to be precluded from participating Fig. 3. Contact signaling within mPGES-1. (A) The van der Waals contact network identified by qFit conformational fitting and subsequent CONTACT analysis of mPGES-1 PDB entries are shown with translucent molecular surface representations over alternate conformers and correspondingly colored node diagrams. The nodes are connected by edges whose width is weighted according to the number of networks involving the pair they connect. The network observed within PDB ID code 4AL0, which includes the active site residue D49, is shown in red, both from the perspective of the membrane plane (Left) and the cy- toplasm (Right). (B) The corresponding contact networks identified from the qFit ensemble complex within the bis-phenyl GSH complex (PDB ID code 4AL1) are much smaller and are shown in cyan and red. The latter contains the active site residue R126, now observed in dual conformations, possibly due to negation of GSH thiol interaction. One possible pathway is show in more detail within Movie S1. 974 | www.pnas.org/cgi/doi/10.1073/pnas.1522891113 Brock et al.
  • 4. in thiolate stabilization if it was already engaged in a stable salt bridge interaction with Asp-49. Analysis of the relative torsional angles, however, shows the out-of-plane angle of the Asp-49 carboxylate relative to the Arg-126 guanidinium to be 44.7° (Fig. 2A). This value is far in excess of the ∼8–10° found to be typical of bidentate inter- actions for structures of a similar resolution as reported in a recent comprehensive review (31). Therefore, the Asp-49–Arg-126 in- teraction cannot be classified as the energetically stable, bidentate interaction of a formal salt bridge and implies that the energetic barrier for its disruption would be low. This finding provides evidence for the capacity of these residues to dynamically participate in active site chemistry and is corroborated by conformational fitting with qFit (27), which reveals hidden conformations of both residues depending on the identity of the adjacent ligand (Fig. 2). As we had previously observed for Arg-126 mutants (12), we found that other catalytically inactive mutants assayed in this study displayed a promiscuous reductase activity, converting PGH2 into PGF2α. Notably, the most pronounced activity in this respect was again observed for microsomal preparations of the Arg-126–Lys mutant (Fig. S1). We confirmed that all mutants possessed the same tertiary fold as native enzyme via comparison of circular-dichroism spectra (Fig. S2). A Crystallographic Water Molecule Is Ideally Situated to Participate in the Mechanism. Analysis of the active site architecture also sug- gests that the α-carboxylate of GSH is involved in thiolate for- mation, via a tightly bound crystallographic water (2Fo − Fc peak of ∼5 rmsd, ADP = 21.9 Å2 ) within the active site (Fig. 2A). By forming a hydrogen-bonding network from the α-carboxylate moiety of GSH to its thiol group, it is ideally placed to assist in deprotonation of the latter during catalysis. The pKa of the α-carboxylate, in turn, is undoubtedly lowered by the side-on, out-of-plane interaction with the guanidinium of Arg-38 (torsion angle 45.9°), which is itself engaged in solvent-mediated inter- actions with the main-chain carboxyl groups of Ala-43 and Arg- 60. This architecture is highly reminiscent of the “electron-sharing network” that is functionally conserved in all classes of soluble GSTs for the same purpose (32), and the use of a bridging water molecule to transfer the thiol proton to the α-carboxylate of GSH has been shown to be energetically favorable within an alpha class soluble GST (33). Crucially, density for this water is absent for the Phenix (34) refined bis-phenyl GSH complex (PDB ID code 4AL1), in which the relative occupancies to GSH were refined as 0.87:0.13, respectively. After conformational change of Asp-49, we hypothesize that Arg-126 can further decrease the GSH thiol pKa via charge stabilization. The crystallographic water molecule could then function as the yet-unidentified base that accepts a proton from GSH during thiolate formation, concurrently forming a transient hydronium ion or shuttling the proton to the α-carboxylate. Once formed and stabilized by in- teraction with Arg-126, we expect attack of GSH thiolate upon the endoperoxide ring at the C-9 position, resulting in O–O bond cleavage and proton donation via the hydronium ion. Asp-49, liberated from its interaction with Arg-126, would now be free to function as a base within the resulting transition state, facilitating a decrease of the C-9 proton pKa, and spontaneous decomposition to yield the product PGE2 and regenerated GSH (Fig. 4). Although an alternative mechanism in which thiolate would act as a general base abstracting the C-9 proton has been suggested to be more energetically favorable in model systems (35, 36), the probability of either pathway would ultimately be determined by the precise orientation of substrate relative to cofactor within the enzymatic active site. Although the apparent ability of the active site mu- tants characterized here to produce PGF2α via reduction of a putative sulphenic acid ester intermediate speaks in favor of the former (Fig. S1), this alternative mechanism is shown in Fig. S4. A Contact Signaling Network Modulates Active Site Residues in mPGES-1. We submitted the PDB entries associated with the recently published crystal structure of mPGES-1 (PDB ID codes 4AL0 and 4AL1) (13) to the qFit server (smb.slac.stanford.edu/qFitServer/) (20). This software automatically samples conformational heterogeneity that is interpretable by fitting partial occupancy conformational ensembles into low-level electron density. The CONTACT algorithm was then used to calculate resulting van der Waals contact networks that in- dicate a probable correlation of conformations at each site (27). Post-qFit conformational fitting and subsequent refinement by Phenix (34) of the 1.16-Å binary complex with GSH (PDB ID code 4AL0) resulted in a small, but significant, improvement of structure quality indicators, including the decrease of R/Rfree values from 12.2/13.0% to 11.6/12.8%, respectively. Subsequent analysis of the structure with CONTACT revealed an extensive network of cor- related side chain interactions centered upon the short, cytoplasmic helix separating transmembrane helices I and II that was referred to by Sjögren et al. (13), and will be hereafter, as the C-domain. Of most interest is that the network facilitates signal transduction from residues involved in the recognition of GSH to the active site res- idue Asp-49. This process could facilitate the disruption of its interaction with Arg-126, facilitating the latter’s role in thiolate stabilization on a time scale specific to catalysis. Thr-34 and Leu-69, located on helices I and II, respectively, make hydrophobic contacts with the γ-glutamyl moiety of GSH and initiate series of correlated van der Waals overlaps that ultimately affect Asp-49, e.g., Leu-69 → Thr-34 → Cys-68 → Asp-64 → Lys-41 → Arg-40 → Leu-39 → His- 53 → K42 → H53 → Asp-49 (Fig. 3A and Movie S1). Conversely, Arg-126, with which it forms an intermonomeric interaction, is fitted as the single conformation observed within the crystal structure (13) (Fig. 2A). As discussed above, we believe this active site geometry is strong evidence of a GSH thiol–Arg-126 interaction. Although the pKa of GSH thiol has been measured as 9.42 in solution (37), the dynamic interaction of GSH thiol with Arg-126, combined with the solvent restricted electrostatics of the active site, may allow GSH to transiently form thiolate during ca- talysis via a mechanism of charge redistribution. Specifically, the hydrogen-bonding network formed by a crystallographic water molecule between the α-carboxylate and thiol moieties of GSH may be crucial in this respect (Figs. 2A and 4). This finding is corroborated by comparison with the qFit- generated structural ensembles of the bis-phenyl complex (PDB ID code 4AL1), in which Arg-126, now with a reduced potential for interaction with thiol, is observed to be in dynamic motion (Figs. 2B and 3B) (see below). Fig. 4. Proposed mechanism of mPGES-1. For details, please see Results and Discussion. Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 975 BIOCHEMISTRY
  • 5. Contact Signaling Is Ligand-Dependent. After an iterative fitting of alternate conformations with qFit (20), building of N-terminal residues into density, and subsequent refinement with Phenix (34) (described in Materials and Methods), a significant im- provement of quality indicators was achieved for the 1.95-Å resolution mPGES-1 complex with bisphenyl–GSH (PDB ID code 4AL1), with a reduction of R/Rfree values from 16.3/17.2% to 13.7/16.6%, respectively. Subsequent CONTACT analysis lacked the extensive signaling network found within the GSH complex, however, which were instead focused on opposing sides of the active site. The ensemble structure was found to contain two networks of four and five residues, respectively, the latter of which occurs in the cytoplasmic loop between helices III and IV and contains alternate conformations of Arg-126 (Fig. 3B). This finding suggests that the activation of dynamic contact networks in mPGES-1 may be dependent upon the identity of the ligand bound at the active site. Although the difference in structural information inherent in the two datasets (1.16 Å cf. 1.95 Å) should be considered when drawing comparisons between the qFit-generated ensemble structures, the resolution of the 4AL1 dataset is still significantly higher than the upper limit of 2.1 Å suggested by the software developers (smb.slac.stanford.edu/ qFitServer/) (20). In addition, we performed a qFit/CONTACT analysis of high-resolution (1.41–1.52 Å) mPGES-1 inhibitor complexes recently published (15) (PDB ID codes 4YK5, 4YL0, 4YL1, and 4YL3). These four compounds are also observed to bind in the intermonomeric active site, making extensive contact with the C-domain. Although the four inhibitors are varied in structure and binding modes, they all share a common interaction with the C-domain and lack the extensive networks found in the holoenzyme complex with GSH (PDB ID code 4AL0). Intriguingly, the same interaction is also fulfilled by an octyl glucoside (n-octyl-β-D- glucoside) detergent molecule (not shown in Figs. 2 and 3 for clarity; cf. Fig. S3) within the bis-phenyl complex (PDB ID code 4AL1), whose polar head group also makes contact with the turn/helix C-domain motif and whose hydrophobic tail stacks against the bis- phenyl moiety of the GSH analog (13) (Fig. S3). This finding indi- cates that stabilizing contacts within this region may disrupt potential for signal transduction (Fig. S3). As shown in Fig. 3A and Movie S1, the dynamic conformations of Lys-41, Arg-40, Leu-39, and His-53 are essential to the transmission of the contact network within the C-domain, and ultimately to the active site residue, Asp-49. There- fore, it is possible that their mode of inhibition is mediated by fa- voring certain conformations of these residues from the structural ensemble and subsequent interruption of signaling (26). This mechanism could be a common theme of potent mPGES-1 inhibitors. In a recent analysis of binding sites via mass spectrometry hydrogen/deuterium exchange experiments (38), the authors found that the greatest differences common to the two most potent inhibitors were observed in residues 37–54, corresponding to the C-domain. Conclusions The combined results of site-directed mutagenesis, functional assays, structural ensemble, and contact network analysis pre- sented herein provide strong evidence for a mechanism of PGE2 synthesis by mPGES-1 that features an activation of GSH thio- late by Arg-126, modulated via its respective interaction with Asp-49. Furthermore, we show that conformations of the latter can be affected by a ligand-dependent contact signaling, con- necting it to distal residues involved in GSH recognition, with the potential to dynamically alter the Asp-49–Arg-126 interaction during catalysis (Fig. 3). We propose a previously unidentified mechanism of PGH2 isom- erization by mPGES-1 that features a prominent role of a water- mediated interaction with the α-carboxylate of GSH and an Asp- 49–mediated thiolate stabilization by Arg-126 (Fig. 4). We hypothesize that the active site of mPGES-1 lowers the pKa of GSH thiol and the C-9 proton of PGH2 concurrently via respective interac- tions with Arg-126 and Asp-49, facilitated by their dynamic con- formational change in response to contact network signaling. Charge conservation in this solvent-restricted environment could thus be achieved via proton shuffling by the crystallographic water/α-carboxylate hydrogen-bonding network (Fig. 4). This work has broad implications for the pharmacological efforts to inhibit this enzyme, which are a current topic of discussion within the literature (39). Materials and Methods Protein Expression and Purification. Recombinant wild-type (WT) and active-site mutants of human mPGES-1 were overexpressed in P. pastoris and purified by Ni-affinity chromatography before exchanging buffer to 0.1 M phosphate buffer, 0.03% dodecyl maltoside, and 2.5 mM GSH, pH 7.4. Microsomal prep- arations were prepared via ultracentrifugation of lysed cell supernatant and homogenization of the microsomal pellets in assay buffer (20 mM Tris·HCl, pH 7.8, 2.5 mM GSH). For further details, please refer to SI Materials and Methods. Synthesis of GOH. The oxygen analog of GSH, GOH, was synthesized in a three- step procedure based on a published method (40). For further details, please see SI Materials and Methods. Enzyme Activity Assay. Conversion of PGH2 to PGE2 by WT or mutated mPGES-1 were quantified by using GC-MS as described (12). For further details, please refer to SI Materials and Methods. Analysis of Dynamic Contact Networks. The qFit Web server (smb.slac. stanford.edu/qFitServer/) and the CONTACT algorithm (27) were used for the quantification of conformational ensembles and functional contact networks, respectively. Before analysis, the physiological trimer was gener- ated from the asymmetric unit via crystallographic symmetry using the program COOT (41). For PDB ID codes 4AL0, 4YL0, 4YL1, 4YL3, and 4YK5, the coordinates were submitted to the qFit server and refined and prepared for CONTACT as described (27), by using Phenix-1.9-1692 (34) without manual intervention. For PDB ID code 4AL1, significant density improvement at the amino terminus allowed residues 4–9 to be built into density after qFit conformer fitting and refinement. After a second round of refinement, the resulting improvement in quality indicators such as the Rfree value were significant, such that the improved phase estimates were anticipated to affect the conformational ensemble fitting. Hence, the improved coordi- nates were resubmitted to the qFit server before being refined and pre- pared for analysis with CONTACT as above. Settings for all CONTACT analyses were as follows: Tstress (percentile) = 0.4, max_path_length = 100, sc_only_flag = f (all atom), relief_threshold = 0.90. ACKNOWLEDGMENTS. We thank Gunvor Hamberg for technical assistance and gratefully acknowledge the late Richard Armstrong, who provided the GOH GSH analogue. Part of this work was performed at the Karolinska Institutet Protein Science Facility. Some computations were performed on resources provided by the Swedish National Infrastructure for Computing at Linköping University. This work was supported by Swedish Research Coun- cil Grant 10350 and CERIC Linnaeus Grant; the Stockholm County Council (Cardiovascular Program, Thematic Center Inflammation); and NovoNordisk Foundation Grant NNF15CC0018346. J.Z.H. is the recipient of a Distinguished Professor Award from Karolinska Institutet. 1. Samuelsson B, et al. (1978) Prostaglandins and thromboxanes. Annu Rev Biochem 47(1):997–1029. 2. Jakobsson PJ, Thorén S, Morgenstern R, Samuelsson B (1999) Identification of human prostaglandin E synthase: A microsomal, glutathione-dependent, inducible enzyme, constituting a potential novel drug target. Proc Natl Acad Sci USA 96(13):7220–7225. 3. Watanabe K, Kurihara K, Suzuki T (1999) Purification and characterization of mem- brane-bound prostaglandin E synthase from bovine heart. Biochim Biophys Acta 1439(3):406–414. 4. Murakami M, et al. (2000) Regulation of prostaglandin E2 biosynthesis by inducible membrane-associated prostaglandin E2 synthase that acts in concert with cyclo- oxygenase-2. J Biol Chem 275(42):32783–32792. 5. Samuelsson B, Morgenstern R, Jakobsson P-J (2007) Membrane prostaglandin E syn- thase-1: A novel therapeutic target. Pharmacol Rev 59(3):207–224. 6. Jakobsson PJ, Morgenstern R, Mancini J, Ford-Hutchinson A, Persson B (2000) Mem- brane-associated proteins in eicosanoid and glutathione metabolism (MAPEG). A widespread protein superfamily. Am J Respir Crit Care Med 161(2 Pt 2):S20–S24. 976 | www.pnas.org/cgi/doi/10.1073/pnas.1522891113 Brock et al.
  • 6. 7. Ferguson AD, et al. (2007) Crystal structure of inhibitor-bound human 5-lipoxygenase- activating protein. Science 317(5837):510–512. 8. Ago H, et al. (2007) Crystal structure of a human membrane protein involved in cysteinyl leukotriene biosynthesis. Nature 448(7153):609–612. 9. Martinez Molina D, et al. (2007) Structural basis for synthesis of inflammatory me- diators by human leukotriene C4 synthase. Nature 448(7153):613–616. 10. Rinaldo-Matthis A, et al. (2010) Arginine 104 is a key catalytic residue in leukotriene C4 synthase. J Biol Chem 285(52):40771–40776. 11. Saino H, et al. (2011) The catalytic architecture of leukotriene C4 synthase with two arginine residues. J Biol Chem 286(18):16392–16401. 12. Hammarberg T, et al. (2009) Mutation of a critical arginine in microsomal prosta- glandin E synthase-1 shifts the isomerase activity to a reductase activity that converts prostaglandin H2 into prostaglandin F2alpha. J Biol Chem 284(1):301–305. 13. Sjögren T, et al. (2013) Crystal structure of microsomal prostaglandin E2 synthase provides insight into diversity in the MAPEG superfamily. Proc Natl Acad Sci USA 110(10):3806–3811. 14. Li D, et al. (2014) Crystallizing membrane proteins in the lipidic mesophase. experi- ence with human prostaglandin E2 synthase 1 and an evolving strategy. Cryst Growth Des 14(4):2034–2047. 15. Luz JG, et al. (2015) Crystal structures of mPGES-1 inhibitor complexes form a basis for the rational design of potent analgesic and anti-inflammatory therapeutics. J Med Chem 58(11):4727–4737. 16. Berman HM, et al. (2000) The Protein Data Bank. Nucleic Acids Res 28(1):235–242. 17. Kuzmanic A, Pannu NS, Zagrovic B (2014) X-ray refinement significantly underesti- mates the level of microscopic heterogeneity in biomolecular crystals. Nat Commun 5:3220. 18. Lang PT, Holton JM, Fraser JS, Alber T (2014) Protein structural ensembles are re- vealed by redefining X-ray electron density noise. Proc Natl Acad Sci USA 111(1): 237–242. 19. Burnley BT, et al. (2012) Modelling dynamics in protein crystal structures by ensemble refinement. eLife 1:e00311. 20. Fraser JS, et al. (2011) Accessing protein conformational ensembles using room- temperature X-ray crystallography. Proc Natl Acad Sci USA 108(39):16247–16252. 21. Lang PT, et al. (2010) Automated electron-density sampling reveals widespread conformational polymorphism in proteins. Protein Sci 19(7):1420–1431. 22. Henzler-Wildman KA, et al. (2007) Intrinsic motions along an enzymatic reaction trajectory. Nature 450(7171):838–844. 23. Yang L-Q, et al. (2014) Protein dynamics and motions in relation to their functions: Several case studies and the underlying mechanisms. J Biomol Struct Dyn 32(3): 372–393. 24. Fraser JS, et al. (2009) Hidden alternative structures of proline isomerase essential for catalysis. Nature 462(7273):669–673. 25. Klinman JP (2015) Dynamically achieved active site precision in enzyme catalysis. Acc Chem Res 48(2):449–456. 26. Fischer M, Coleman RG, Fraser JS, Shoichet BK (2014) Incorporation of protein flexi- bility and conformational energy penalties in docking screens to improve ligand discovery. Nat Chem 6(7):575–583. 27. van den Bedem H, Bhabha G, Yang K, Wright PE, Fraser JS (2013) Automated iden- tification of functional dynamic contact networks from X-ray crystallography. Nat Methods 10(9):896–902. 28. Ouellet M, et al. (2002) Purification and characterization of recombinant microsomal prostaglandin E synthase-1. Protein Expr Purif 26(3):489–495. 29. Oakley A (2011) Glutathione transferases: A structural perspective. Drug Metab Rev 43(2):138–151. 30. Zhou P, Tian F, Lv F, Shang Z (2009) Geometric characteristics of hydrogen bonds involving sulfur atoms in proteins. Proteins 76(1):151–163. 31. Donald JE, Kulp DW, DeGrado WF (2011) Salt bridges: Geometrically specific, de- signable interactions. Proteins 79(3):898–915. 32. Winayanuwattikun P, Ketterman AJ (2005) An electron-sharing network involved in the catalytic mechanism is functionally conserved in different glutathione transferase classes. J Biol Chem 280(36):31776–31782. 33. Dourado D, Fernandes P, Mannervik B, Ramos M (2008) Glutathione transferase: New model for glutathione activation. Chemistry 14(31):9591–9598. 34. Adams P, et al. (2010) PHENIX: A comprehensive Python-based system for macro- molecular structure solution. Acta Crystallogr D Biol Crystallogr 66(2):213–221. 35. Yamaguchi N, et al. (2011) Theoretical studies on model reaction pathways of pros- taglandin H2 isomerization to prostaglandin D2/E2. Theor Chem Acc 128(2):191–206. 36. Li Y, Angelastro M, Shimshock S, Reiling S, Vaz RJ (2010) On the mechanism of mi- crosomal prostaglandin E synthase type-2—A theoretical study of endoperoxide re- action with MeS−. Bioorg Med Chem 20(1):338–340. 37. Tajc SG, Tolbert BS, Basavappa R, Miller BL (2004) Direct determination of thiol pKa by isothermal titration microcalorimetry. J Am Chem Soc 126(34):10508–10509. 38. Prage EB, et al. (2011) Location of inhibitor binding sites in the human inducible prostaglandin E synthase, MPGES1. Biochemistry 50(35):7684–7693. 39. Koeberle A, Werz O (2015) Perspective of microsomal prostaglandin E 2 synthase-1 as drug target in inflammation-related disorders. Biochem Pharmacol 98(1):1–15. 40. Chen WJ, Boehlert CC, Rider K, Armstrong RN (1985) Synthesis and characterization of the oxygen and desthio analogues of glutathione as dead-end inhibitors of gluta- thione S-transferase. Biochem Biophys Res Commun 128(1):233–240. 41. Emsley P, Cowtan K (2004) Coot: Model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr 60(Pt 12 Pt 1):2126–2132. 42. Humphrey W, Dalke A, Schulten K (1996) VMD: Visual molecular dynamics. J Mol Graph 14:33–38, 27–28. 43. Phillips JC, et al. (2005) Scalable molecular dynamics with NAMD. J Comput Chem 26(16):1781–1802. 44. Lomize MA, Lomize AL, Pogozheva ID, Mosberg HI (2006) OPM: Orientations of proteins in membranes database. Bioinformatics 22(5):623–625. Brock et al. PNAS | January 26, 2016 | vol. 113 | no. 4 | 977 BIOCHEMISTRY