SlideShare uma empresa Scribd logo
1 de 5
Baixar para ler offline
Probing the Catalytic Mechanism of the Insulin Receptor Kinase
with a Tetrafluorotyrosine-containing Peptide Substrate*
Received for publication, April 25, 2000, and in revised form, June 22, 2000
Published, JBC Papers in Press, June 26, 2000, DOI 10.1074/jbc.M003524200
Ararat J. Ablooglu‡§, Jeffrey H. Till§¶, Kyonghee Kimʈ, Keykavous Parangʈ, Philip A. Coleʈ,
Stevan R. Hubbard¶, and Ronald A. Kohanski‡ **
From the ‡Mount Sinai School of Medicine, Department of Biochemistry and Molecular Biology, New York, New York,
10029 ¶Skirball Institute of Biomolecular Medicine and Department of Pharmacology, New York University School of
Medicine, New York, New York 10016, and ʈThe Johns Hopkins University School of Medicine, Department of
Pharmacology and Molecular Sciences, Baltimore, Maryland 21205
The interaction of a synthetic tetrafluorotyrosyl pep-
tide substrate with the activated tyrosine kinase do-
main of the insulin receptor was studied by steady-state
kinetics and x-ray crystallography. The pH-rate profiles
indicate that the neutral phenol, rather than the chem-
ically more reactive phenoxide ion, is required for en-
zyme-catalyzed phosphorylation. The pKa of the tet-
rafluorotyrosyl hydroxyl is elevated 2 pH units on the
enzyme compared with solution, whereas the phenoxide
anion species behaves as a weak competitive inhibitor
of the tyrosine kinase. A structure of the binary enzyme-
substrate complex shows the tetrafluorotyrosyl OH
group at hydrogen bonding distances from the side
chains of Asp1132
and Arg1136
, consistent with elevation
of the pKa. These findings strongly support a reaction
mechanism favoring a dissociative transition state.
Protein kinases are enzymes that catalyze transfer of the
␥-phosphoryl group from ATP to the alcoholic side chains of
amino acid residues on their substrate proteins and peptides.
They play important roles in signaling and have been subject to
intensive cell biologic, enzymologic, structural, and kinetic
scrutiny. Elucidation of the chemical reaction mechanisms of
these enzymes is proving to be important in defining the details
of their regulation and for designing inhibitors.
The reaction mechanism of phosphoryl transfer reactions can
favor either an associative or dissociative pathway (Fig. 1). In
the extremes, these are distinguished by dependence on versus
independence from nucleophilicity of the phosphoryl acceptor,
respectively, distances between the ␥-phosphorus atom and the
attacking and leaving group oxygens, and other factors (1–7).
The associative pathway would benefit from proton abstraction
prior to formation of the transition state so that a more reactive
oxygen anion is formed, whereas that proton is retained in the
dissociative transition state. Earlier kinetic and crystallo-
graphic studies of protein kinases, especially the cAMP-de-
pendent protein kinase, demonstrated the importance of a con-
served aspartyl residue that is essential for the enzyme-
catalyzed reaction (8–10). The location of this Asp166
side chain
in the cAMP-dependent protein kinase crystal structure was
close enough to the serine hydroxyl that it could create a more
reactive nucleophile by removal of the seryl or threonyl proton,
in accordance with an associative mechanism (8, 9), but other
kinetic evidence was interpreted as favoring a dissociative
pathway (5, 11, 12, 14). Never the less, similar positioning of
the conserved aspartyl side chain in other protein kinase struc-
tures has been offered as further supporting evidence for pro-
ton abstraction, thus generating a more reactive entering oxy-
gen as a basis for catalysis (e.g. see Ref. 15). On the other hand,
as discussed by Mildvan (2), distances as short as 2.7 Å be-
tween the entering oxygen and the ␥-phosphorus could be es-
timated from modeling studies (16) and as long as 5.3 Å from
NMR (17); the latter indicates a reaction pathway with a more
dissociative character. Furthermore, the mechanism of phos-
phoryl transfer in chemical reactions of phosphate monoesters,
as models for enzyme-catalyzed reactions, appears to be disso-
ciative with relatively little contribution of nucleophilicity to
the transition state (18, 19). On the premise that chemical and
enzymatic reactions differ in stabilization but not character of
the transition state (20–26), it would seem that protein kinases
should follow a more dissociative pathway.
For protein tyrosine kinases, the ability to vary nucleophi-
licity by aromatic ring substitution on tyrosine facilitated a
chemical analysis of the phosphoryl transfer reaction catalyzed
by C-terminal Src kinase (Csk).1
Analyzing the pH-independ-
ent kinetic parameters for a series of substituted tyrosine de-
rivatives with pKas ranging from 5.2 to 10, it was shown that
kcat and kcat/Km were relatively independent of nucleophilicity
of the attacking phenolic residue, yielding a very small Bron-
sted nucleophile coefficient (0 Ͻ ␤nuc Ͻ 0.1) (6). The pH-rate
profiles also demonstrated that the chemically more reactive
phenoxide anion species was enzymatically less reactive than
the neutral phenol species. Further studies with Csk revealed
that the reverse reaction, where the fluorotyrosyl peptide is the
leaving group, gave a Bronsted leaving group coefficient (␤1g) of
Ϫ0.3 (3). This supports protonation of the phenolic O␩ in both
directions. Thus, evidence from the forward and reverse reac-
tions together suggest that the character of the transition state
for Csk is mostly dissociative (26).
Fluorotyrosyl analogs have also been used with the insulin
receptor (IR) tyrosine kinase (27, 28), but the opposite conclu-
sion had been reached in those studies; i.e. the transition state
* This work was supported by Grant DK50074 from the National
Institutes of Health (NIH) (to R. A. K.), Grant DK52916 from NIH (to
S. R. H.), by the Winston Foundation (to K. K.), and by the Burroughs
Welcome Fund and NIH Grant CA74305 (to P. A. C.). The costs of
publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked “advertisement”
in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
§ Equal contributions were made by these authors.
** To whom correspondence should be addressed. Tel.: 212-241-7288;
Fax: 212-996-7214; E-mail: Ronald.Kohanski@mssm.edu.
1
The abbreviations used are: Csk, C-terminal Src kinase; HPLC,
high performance liquid chromatography; IR, insulin receptor; IRKD,
cytoplasmic kinase domain of the IR; IRK3P, tris-phosphorylated core
kinase domain of the human insulin receptor; IRK, insulin receptor
kinase; Fmoc, N-(9-fluorenyl)methoxycarbonyl.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 275, No. 39, Issue of September 29, pp. 30394–30398, 2000
© 2000 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org30394
atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
for phosphoryl transfer seemed to be associative. Therefore, the
analysis reported here was undertaken to specifically address
this discrepancy between Csk and IR, making use of the fol-
lowing three advances since the earlier work of Graves and
co-workers (27, 28): 1) There is now a better understanding of
the substrate-specificity determinants for the insulin receptor
(29–32), which was a factor in design of the peptide substrates.
2) The recombinant cytoplasmic kinase domain of the insulin
receptor (IRKD) can be expressed as a highly purified protein
whose autophosphorylation state can be fixed experimentally
(31–34). 3) The ternary complex of the catalytic core with a
tyrosyl peptide substrate and MgATP analog has been deter-
mined, which establishes a distance between O␩ of the reactive
Tyr and the ␥-phosphorus atom of the adenine nucleotide (32).
Our results with the autophosphorylated and maximally acti-
vated IRKD support a dissociative mechanism for this protein
kinase, show an increase in pKa of the enzyme-bound fluoroty-
rosyl group compared with aqueous solution, and demonstrate
that this fluorotyrosyl side chain can form hydrogen bonds
virtually identical to those between the unmodified tyrosyl
residue of the substrate and the catalytic core of the kinase.
EXPERIMENTAL PROCEDURES
General—Dithiothreitol (Sigma Ultra), the disodium salt of ATP
(from equine muscle, catalog number A-5394), and bovine serum albu-
min (radioimmunoassay grade) were purchased from Sigma; hydroge-
nated Triton X-100 (protein grade) was from Calbiochem; EDTA was
from Fluka; Tris acetate, Tris base, Tris HCl, and electrophoresis re-
agents were from Roche Molecular Biochemicals; 1,3-bis[tris(hy-
droxymethyl)-methylamino]propane (bis-tris propane) was from Sigma;
magnesium acetate (MgAc2, enzyme grade) was from Fisher. Insect cell
culture media and fetal bovine serum were from Life Technologies, Inc.
Peptide Synthesis—L-Tetrafluorotyrosine was enzymatically pre-
pared with tyrosine phenol-lyase as described previously and converted
to the corresponding Fmoc derivative (6). Fmoc-L-tetrafluorotyrosine
was subsequently incorporated in place of tyrosine in the previously
described 18-residue IRKD substrate peptide, IRS727 (35), with the
amino acid sequence KKKLPATGDYMNMSPVGD (36) modified from
the rat sequence (residues 718–735, with Gly718
and Leu720
changed to
Lys), using solid phase peptide synthesis providing fluoro-IRS727.
IRS727 and fluoro-IRS727 were prepared by solid phase peptide syn-
thesis using the Fmoc strategy on a Rainin PS-3 instrument. After
purification by reversed phase HPLC (preparative C-18 column), pep-
tides were judged to be greater than 95% pure. The peptide substrate
IRS939 (REETGSEEYMNMDLG, residues 931–945 from the rat IRS1
sequence) was prepared by Fmoc synthesis and purified by HPLC, as
described previously (36). Peptides had free amino and carboxyl termini
and included the consensus sequence D(E)YMXM whose interactions at
the PϪ1 to Pϩ3 subsites on the catalytic core of the kinase have been
established (32); Tyr is the P0 residue. Synthetic peptides and adenine
nucleotide were dissolved in 50 mM TrisAc, pH 7.0, the pH value was
readjusted to 7.0 with the addition of dilute acetic acid and/or NaOH,
and the concentrations were adjusted to the desired values for stock
solutions, which were determined spectrophotometrically using the fol-
lowing extinction coefficients: ATP, ⑀ ϭ 15,300 cmϪ1
M
Ϫ1
at 259 nm;
IRS727 and IRS939, ⑀ ϭ 1300 cmϪ1
M
Ϫ1
at 280 nm. These stocks were
kept at Ϫ20 °C under frequent use or at Ϫ80 °C for storage. Integrity of
the reagents was checked periodically by reverse phase HPLC (pep-
tides) or by thin layer chromatography (nucleotides). The pKa of the
phenolic group in fluoro-IRS727 was determined spectrophotometri-
cally at 280 nm, at a peptide concentration of 0.25 mM.
IRKD Expression and Purification—Baculovirus-encoding amino
acid residues 953–1355 of the insulin receptor’s cytoplasmic kinase
domain2
was used to express the IRKD in High Five™ cells (Invitro-
gen). Enzyme purification, autophosphorylation, and storage were es-
sentially as described previously (38). The purified protein was quan-
tified spectrophotometrically at 280 nm (⑀ ϭ 40,200 cmϪ1
M
Ϫ1
for the
unphosphorylated IRKD, and ⑀ ϭ 35,200 cmϪ1
M
Ϫ1
for the autophos-
phorylated IRKD). The autophosphorylation stoichiometry for the
IRKD was 5.4 mol phosphate/mol enzyme, and the enzyme was deter-
mined to be maximally activated (specific activity of 800 mol phospho-
IRS939/min/mol IRKD at 1 mM ATP, 20 mM MgAc2). The absence of an
unphosphorylated or monophosphorylated activation loop, and the
presence of 20% bis and 80% tris phosphorylation of the activation loop
were established in the purified phosphoprotein by endoproteinase
Lys-C digestion and HPLC mapping (not shown).
Peptide Phosphorylation Assays—Steady-state kinetic parameters
were determined with the autophosphorylated IRKD as the enzyme,
using an HPLC-based assay (36). Quantification of the apo and phos-
phopeptide were done by peak area integration with Chrome Perfect
Software (Justice Innovations, Palo Alto, CA). Dead-end inhibition
studies were done over 0.0–0.63 mM fluoro-IRS727, constant 0.08 mM
ATP, and variable IRS939 (0.04–0.2 mM). Reactions were performed at
room temperature in 50 mM bis-tris propane, 5 mM dithiothreitol, 0.05%
bovine serum albumin (w/v), 2–8 nM phospho-IRKD, and 20 mM MgAc2
and at the pH values indicated in the figure legends. The pH of stock
solutions of peptide, ATP, dithiothreitol, and bovine serum albumin
solutions were adjusted accordingly. The reactions were initiated with
the addition of peptide substrate. Kinetic parameters were determined
from the best global fits of the data. The pH dependence of the observed
rate constant kcat,obs were fit assuming a single proton dissociation.
Plots of log(kcat) versus pH values for fluoro-IRS727 were fit to the
equation,
log(kcat,obs) ϭ log(kcat/͓1 ϩ K/H͔) (Eq. 1)
where kcat is the pH-independent rate constant, H is the proton concen-
tration, and K is the dissociation constant for the ionizable group that
is active in the protonated state (pKa ϭ log(1/K)).
Competitive inhibition kinetic analysis was done by fitting all of the
data points to the linear competitive inhibition equation of KinetAsyst
IITM
based on the algorithms of Cleland (39),
v ϭ VM*S/͓KM͑1 ϩ I/Ki͒ ϩ S͔ (Eq. 2)
using a non-linear least squares approach. The fixed substrate was
assumed to be saturating. Kinetic constants Ϯ S.E. are reported here.
X-ray Crystallographic Studies—Expression and purification of the
tris-phosphorylated core kinase domain of the human insulin receptor
(IRK3P, residues 978–1283) has been described previously (32). Crys-
tals were grown at 4 °C by vapor diffusion in hanging drops containing
2.0 ␮l of protein solution (6.5 mg/ml IRK3P and 1 mM fluoro-IRS727)
and 2.0 ␮l of reservoir buffer (18% polyethylene glycol 8000, 100 mM
Tris-HCl, pH 8.5, and 15% ethylene glycol). Small crystals (Ͻ50 ␮m)
were obtained initially and used for microseeding to grow larger crys-
tals (150–200 ␮m). Crystals belong to the trigonal space group P3221
with unit cell dimensions a ϭ b ϭ 71.8 Å, and c ϭ 125.8 Å when frozen.
There is one molecule in the asymmetric unit, and the solvent content
is 48% (assuming a partial specific volume of 0.74 cm3
/g).
Two data sets were used for structural analysis. One data set was
collected on a Rigaku RU-200 rotating anode equipped with a Rigaku
R-AXIS IIC image plate detector. The other data set was collected on
beamline X4A, National Synchrotron Light Source, equipped with an
ADSC Quantum-4 CCD detector. Crystals were flash-cooled in a dry
2
Numbering of the IR and IRKD is from Ebina et al. (37).
FIG. 1. Associative and dissociative reaction mechanisms for
phosphoryl transfer from ATP to the phenolic hydroxyl of tyro-
sine. The upper pathway depicts an associative transition state in
which the entering and leaving groups are closer to the ␥-phosphorous
atom and where the axial bond number to the entering O␩ could be as
high as one, versus a dissociative transition state, shown in the lower
pathway, in which these groups are farther apart, and the axial bond
number could be as low as zero.
Dissociative Transition State of the Insulin Receptor Kinase 30395
atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
nitrogen stream at Ϫ160 °C. Because of only ϳ80% completeness of the
synchrotron data, the data were merged with the rotating anode data,
giving an overall Rsym of 8.3%, an overall I/␴I of 24.4, and an overall
completeness of 91% for data between 30.0 and 2.6 Å. Data were
processed using DENZO and SCALEPACK (40).
A molecular replacement solution was found with AMoRE (41) using
Protein Data Bank entry 1IR3 (32) as a search model. Rigid-body,
positional, and B-factor refinement were carried out using CNS (42).
Model building was performed using O (43). The final model (2464
atoms) has a crystallographic R value of 24% (free R value of 31%) with
root mean square deviations of 0.009 Å for bonds and 1.5° for angles
(30.0–2.6 Å).
RESULTS
The insulin receptor is not a soluble kinase; it is a hetero-
meric transmembrane protein whose intracellular kinase ac-
tivity is regulated by extracellular insulin binding (44). To
perform these studies in a manner similar to those done with
the soluble enzyme Csk, we have utilized the highly purified
and activated (autophosphorylated) cytoplasmic portion of the
IR. The IRKD was previously shown to be a good model for the
holomeric receptor’s basal and activated states, and these fea-
tures are retained by the conserved catalytic core used in
crystallization (33, 45, 46). The tetrafluorotyrosyl peptide sub-
strate (fluoro-IRS727) was chosen for these studies, because
the unmodified form (IRS727) was cocrystallized with the cat-
alytic core of the IRKD (32). The peptide-inhibition studies
were done using IRS939 as the substrate because of signal to
noise considerations in the peptide phosphorylation assays.
Kinetic Analysis—Kinase assays of IRS727 using the IRKD
revealed somewhat different kinetic parameters than those
reported previously for the full-length IR (35). This was possi-
bly because of differences in assay procedures and kinase or
reagent purity. Here, global fitting of the data from our steady-
state assays gave the following kinetic parameters at pH 7: kcat
ϭ 14 Ϯ 2 sϪ1
, and Km ϭ 0.18 Ϯ 0.02 mM IRS727. Analysis of
fluoro-IRS727 showed that it was apparently a poorer sub-
strate than IRS727 at pH 7, with kcat ϭ 0.5 Ϯ 0.1 sϪ1
, and Km
ϭ 0.21 Ϯ 0.05 mM. Given the observed rate reduction, and
previous claims that no phosphorylation was detected with a
different fluorotyrosyl peptide (27, 28), structural identity of
the phosphorylated fluoro-IRS727 product was confirmed rig-
orously by mass spectrometric analysis after HPLC purifica-
tion. In addition, Ͼ95% conversion to phosphorylated fluoro-
IRS727 was observed after prolonged incubation with activated
IRKD, determined by the HPLC-based assay (not shown).
To assess further the significance of these findings, pH-rate
profiles were performed with peptide substrates for IRKD.
Preliminary experiments confirmed that IRKD enzymatic ac-
tivity versus time appeared stable between pH values of 6 and
10. In this range, phosphorylation of IRS727 showed little
change in kcat (Fig. 2) or kcat/Km (not shown)3
with increasing
pH values. In contrast, fluoro-IRS727 showed a marked 18-fold
decrease in kcat with increasing pH values; the highest activity
was observed at pH 6.0, and a pKa ϭ 7.6 was calculated for the
basic limb (Fig. 2). These results suggest that a group present
in fluoro-IRS727 but not in IRS727 requires protonation for
enzymatic processing. This group is likely to be the tetrafluo-
rotyrosine side chain. The pKa ϭ 7.6 is higher than the pKa ϭ
5.7 measured spectrophotometrically for the fluoro-IRS727 in
aqueous solution (Fig. 2, inset). This implies that the tetrafluo-
rotyrosine’s pKa is higher when the substrate is bound to the
enzyme. At pH 7.0, protonated fluoro-IRS727 in solution is only
5% of the total, calculated from the Henderson-Hasselbach
equation and the pKa of 5.7 of the free peptide. The value
kcat/Km, which reflects the protonation equilibrium of the free
substrate, should be 20-fold greater than the observed value,
because only the neutral phenolic form is the substrate. With
this correction it can be calculated that this tetrafluorotyrosyl
peptide is only 2-fold less efficient as a substrate than the
tyrosyl peptide.
To determine whether the phenolate form of the peptide
would bind to the enzyme, we measured the inhibitory proper-
ties of fluoro-IRS727 at pH 9, where the tetrafluorophenolic
group would be fully deprotonated in free solution, and the
peptide would be minimally active as substrate (kcat ϭ 0.04 sϪ1
)
compared with the varied peptide substrate IRS939 (kcat ϭ
12 Ϯ 2 sϪ1
; Km 0.040 Ϯ 0.002 mM). Fluoro-IRS727 is a linear
competitive inhibitor against IRS939, with Ki ϭ 0.59 Ϯ 0.14 mM
determined from a global fit of the data (Fig. 3). Therefore the
phenolate form of fluoro-IRS727 binds weakly to the kinase.
Crystal Structure of the Binary Complex—To gain better
insight into the interaction of the tetrafluorotyrosyl group with
active site residues in the kinase, the crystal structure of the
fluoro-IRS727 complexed with the IRKD core was determined
at 2.6-Å resolution (Fig. 4). When compared with the ternary
complex of the autophosphorylated core with the unmodified
IRS727 and Mg-5Ј-adenylyl imidodiphosphate (32), the P0 res-
idue (the reactive tyrosine) of the substrate peptide and key
residues in the active site with which it makes contact are
virtually unchanged (root mean square deviation of 0.2 Å for
the side-chain and main-chain atoms of residues Asp1132
and
Arg1136
and the P0 tetrafluorotyrosine). Importantly, the dis-
tance between the phenolic hydroxyl and Asp1132
is nearly the
same in the fluoro-IRS727 complex (2.8 Å) as in the ternary
complex (2.7 Å), indicating that a hydrogen bond is made be-
tween these groups in both cases. In addition, the distance
between the phenolic O␩ and the N␩ of Arg1136
is unperturbed,
and the hydrogen bond between N␩ of Arg1136
and O␦2 of
Asp1132
is also maintained. Together these describe a triangle
of hydrogen bonding interactions (Fig. 4). This structure is
compatible with the significant elevation of the phenolic pKa in
the complex compared with free solution.
DISCUSSION
The transition states in phosphoryl transfer reactions cata-
lyzed by enzymes have been the subject of lively discourse over
the past several years (1–7, 18–25, 47). Mounting evidence
suggests that these enzymes affect catalysis by enhancing but
not fundamentally altering the non-enzymatic mechanisms.
3
It was not possible to precisely measure the kcat/Km dependence on
pH because of inadequate signal to noise ratios and the tendency of
activated IRKD to autodephosphorylate at high enzyme concentration,
which would have been employed to increase net product formation and
improve the signal.
FIG. 2. The pH dependence of kcat for fluoro-IRS727. The log(k-
cat) was plotted versus the reaction pH values for IRS727 (squares) and
for fluoro-IRS727 (circles). The inset shows the spectrophotometric ti-
tration of fluoro-IRS727 (triangles). The dotted lines are the best theo-
retical fit for these data, assuming a single ionizable group in both
cases.
Dissociative Transition State of the Insulin Receptor Kinase30396
atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
Among the experimental approaches most consistently applied
to chemical and enzymatic reactions has been linear free en-
ergy relationships to probe the mechanistic role of nucleophi-
licity (1, 18). Measurements of ␤nuc Յ 0.3 for non-enzymatic
phosphoryl transfer reactions involving phosphate monoesters
suggested only a small impact of nucleophilicity in formation of
the transition state and therefore favored a dissociative mech-
anism (1, 13, 18, 48–50). Extending this conclusion to protein
kinases would suggest that proton abstraction, to generate the
chemically more reactive anionic oxygen, should not be a sig-
nificant feature of enzyme-catalyzed phosphoryl transfer.
This hypothesis can be tested directly using substrates for
protein tyrosine kinases, which offer the unusual opportunity
to alter nucleophilicity of the entering oxygen in an enzymatic
phosphoryl transfer reaction by modification of the aromatic
ring in the substrate (6, 25, 26). Systematic substitution of
hydrogens with fluorine confers a broad range of pKa effects
and allows Bronsted coefficients to be determined. This ap-
proach has provided compelling evidence to support a dissocia-
tive transition state for the reaction catalyzed by Csk (6, 26)
but an associative transition state for the reaction catalyzed by
IR (27, 28). There were two important concerns not addressed
specifically in the study on Csk that are addressed here and
that allowed us to re-evaluate the basis for an apparent dis-
crepancy between Csk and IRK catalytic mechanisms. First,
because the fluorine atomic radius is only 0.2 Å greater than
the hydrogen atomic radius, these substitutions should not
impose steric constraints on orientation of the P0 residue in the
active site. That has been confirmed by comparison of two
structures, showing the virtual absence of any positional dif-
ference between the enzyme-bound P0 tetrafluorotyrosine, re-
ported here, and the enzyme-bound P0 tyrosine reported pre-
viously by Hubbard (32). Indeed, this shows that positioning of
the phosphoryl-acceptor oxygen occurs even in the absence of
MgATP (or analog), at least in a ground state binary complex.
This structural observation (Fig. 4) allows greater confidence in
the significance of kinetic studies using fluorine-substituted
tyrosyl peptides with respect to the phosphoryl transfer mech-
anism of the natural tyrosine-containing substrates.
Second, it was presumed that the pKa of the fluorotyrosyl
derivatives may be altered in the enzyme-bound state (27). The
kcat versus pH profile likely represents the pKa of the phenolic
group when the substrate is bound to the enzyme (in the
ternary complex). It is apparently elevated 1.5–2 pH units
when compared with aqueous solution (Fig. 2). The increased
pKa in the bound state is also supported by the x-ray structure,
which shows a hydrogen bond present between Asp1132
and the
tetrafluorotyrosine phenol. These crystals were grown at pH
8.5, and in order for the neutral phenol to be the major species
in the complex at this pH we would expect a higher pKa in this
complex (perhaps 2–3 units greater than for the free peptide).
These crystals lack an ATP analog that could affect the pKa of
the phenolic group. Additionally, the pKa in the solid state may
be harder to assess given the presence of large quantities of
polyethylene glycol and other factors that might govern crys-
tallization. However, the physical basis for pKa elevation ob-
served in these kinetic experiments could be the electrostatic
repulsion that would result from simultaneous deprotonation
of the substrate-hydroxyl and the conserved catalytic aspartyl
side chain, whose proximity has been established for IRK (see
Ref. 32 and Fig. 4).
The elevated pKa may appear surprising at first glance,
because the enzymatic reaction ultimately catalyzes proton
removal concomitant with phosphoryl transfer. However, we
can estimate ␤nuc Ϸ 0.1 by taking into account the fact that
kcat/Km should represent all steps prior to the first irreversible
step and that kcat/Km of fluoro-IRS727 (solution pKa ϭ 5.7) is
only 2-fold less than IRS727 (the tyrosyl group in solution will
have a pKa ϭ 10). This small value is in good agreement with
the ␤nuc determined for protein tyrosine kinase Csk using a
broader range of fluorotyrosyl derivatives (6, 26). Therefore,
the reaction mechanism for IRK favors a dissociative pathway
such that facilitation of proton removal early in the reaction
coordinate is unnecessary, and apparently unfavorable, as dis-
cussed above. Our finding that the phenolate form of fluoro-
IRS727 was a weak competitive inhibitor and not a substrate
further supports this view. It seems clear, therefore, that a
neutral phenolic group of the P0 residue is required for enzy-
matic processing by IRK, as with Csk, and the transition states
for IRK and Csk are likely to be similar.
This and other published evidence (1–3) suggests the pro-
tein-tyrosine and serine-threonine kinases, as do other en-
zymes involved in phosphoryl transfer of phosphate mo-
noesters, likely favor dissociative transition states.
FIG. 3. Competitive inhibition by fluoro-IRS727. Inhibition of
IRS939 phosphorylation was determined at pH 9.0, and the data are
presented as a double-reciprocal plot. Reaction conditions are given
under “Experimental Procedures.” The concentrations of fluoro-IRS727
were 0 mM (solid circles), 0.32 mM (open circles), and 0.63 mM (inverted
solid triangles).
FIG. 4. Electron density map in the active site of the binary
complex between IRK3P and the tetrafluorotyrosine-contain-
ing peptide. Shown is an Fo Ϫ Fc omit map (30.0–2.6 Å, 3␴ contour) in
which the side chains of Asp1132
, Arg1136
, and the tetrafluorotyrosine
(Tyr(4F)) were omitted from the map calculation. Carbon atoms are
colored yellow, oxygen atoms are red, nitrogen atoms are blue, and
fluorine atoms are magenta. Hydrogen bond (black dashes) distances
are indicated.
Dissociative Transition State of the Insulin Receptor Kinase 30397
atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
Establishing the nature of the protein kinase transition states
provides geometric and functional constraints for understand-
ing the roles of individual residues in regulation of kinase
activity, the factors that govern protein substrate selectivity,
and the design of transition state analogs as selective
inhibitors.
REFERENCES
1. Herschlag, D., and Jencks, W. P. (1989) J. Am. Chem. Soc. 111, 7587–7596
2. Mildvan, A. S. (1997) Proteins 29, 401–416
3. Cole, P. A., Sondhi, D., and Kim, K. (1999) Pharmacol. Ther. 82, 219–229
4. Cole, P. A., Grace, M. R., Phillips, R. S., Burn, P., and Walsh, C. T. (1995)
J. Biol. Chem. 270, 22105–22108
5. Zhou, J., and Adams, J. A. (1997) Biochemistry 36, 2977–2984
6. Kim, K., and Cole, P. A. (1997) J. Am. Chem. Soc. 119, 11096–11097
7. Grace, M., Walsh, C. T., and Cole, P. A. (1997) Biochemistry 36, 1874–1881
8. Knighton, D. R., Zheng, J., Ten Eyck, L. F., Ashford, V. A., Xuong, N.-H.,
Taylor, S. S., and Sowadski, J. M. (1991) Science 253, 407–413
9. Bossemeyer, D., Engh, R. A., Kinzel, V., Ponstingl, H., and Huber, R. (1993)
EMBO J. 12, 849–859
10. Gibbs, C. S., and Zoller, M. J. (1991) J. Biol. Chem. 266, 8923–8931
11. Adams, J. A., and Taylor, S. S. (1992) Biochemistry 31, 8516–8522
12. Adams, J. A. (1996) Biochemistry 20, 10949–10956
13. Epstein, J., Plapinger, R. E., Michel, H. O., Cable, J. R., Stephani, R. A.,
Billington, C. A., and West, G. R. (1964) J. Am. Chem. Soc. 86, 3075–3084
14. Grant, B. D., and Adams, J. A. (1996) Biochemistry 35, 2022–2029
15. Lowe, E. D., Noble, M. E., Skamneli, V. T., Oikonomahos, N. G., Owen, D. J.,
and Johnson, L. N. (1997) EMBO J. 16, 6646–6658
16. Madhusudan, Trafny, E. A., Xuong, N. H., Adams, J. A., Ten Eyck, L. F.,
Taylor, S. S., and Sowadski, J. M. (1994) Protein. Sci. 3, 176–187
17. Granot, J., Mildvan, A. S., Bramson, H. N., and Kaiser, E. T. (1980) Biochem-
istry 19, 3537–3543
18. Herschlag, D., and Jencks, W. P. (1989) J. Am. Chem. Soc. 111, 7597–7606
19. Admiraal, S. J., and Herschlag, D. (2000) J. Am. Chem. Soc. 122, 2145–2148
20. Knowels, J. R. (1980) Annu. Rev. Biochem. 49, 877–919
21. Ho, M., Bramson, H. N., Hansen, D. E., Knowels, J. R., and Kaiser, E. T. (1988)
J. Am. Chem. Soc. 110, 2680–2688
22. Admiraal, S. J., and Herschlag, D. (1995) Chem. Biol. 2, 729–739
23. Florian, J., and Warshel, A. (1997) J. Am. Chem. Soc. 119, 5473–5474
24. Jones, J. P., Weiss, P. M., and Cleland, W. W. (1991) Biochemistry 30,
3624–3629
25. Cole, P. A., Burn, P., Takacs, B., and Walsh, C. T. (1994) J. Biol. Chem. 269,
30880–30887
26. Kim, K., and Cole, P. A. (1998) J. Am. Chem. Soc. 120, 6851–6858
27. Martin, B. L., Wu, D., Jakes, S., and Graves, D. J. (1990) J. Biol. Chem. 265,
7108–7111
28. Yuan, C.-J., Jakes, S., Elliott, S., and Graves, D. J. (1990) J. Biol. Chem. 265,
16205–16209
29. Songyang, Z., Carraway, K. L., Eck, M. J., Harrison, S. C., Feldman, R. A.,
Mohammadi, M., Schlessinger, J., Hubbard, S. R., Smith, D. P., Eng, C.,
Lorenzo, M. J., Ponder, B. A. J., Mayer, B. J., and Cantley, L. C. (1995)
Nature 373, 536–539
30. Feener, E. P., Badur, J. M., King, G. L., Wilden, P. A., Sun, X., Kahn, C. R., and
White, M. F. (1993) J. Biol. Chem. 268, 11256–11264
31. Wei, L., Hubbard, S. R., Hendrickson, W. A., and Ellis, L. (1995) J. Biol. Chem.
270, 8122–8130
32. Hubbard, S. R. (1997) EMBO J. 16, 5572–5581
33. Kohanski, R. A. (1993) Biochemistry 32, 5766–5772
34. Cann, A. D., and Kohanski, R. A. (1997) Biochemistry 36, 7681–7689
35. Shoelson, S. E., Chatterjee, S., Chaudhuri, M., and White, M. F. (1992) Proc.
Natl. Acad. Sci. U. S. A. 89, 2027–2031
36. Cann, A. D., Wolf, I., and Kohanski, R. A. (1997) Anal. Biochem. 247, 327–332
37. Ebina, Y., Ellis, L., Jarnagin, K., Edery, M., Graf, L., Clauser, E., Ou, J.,
Masiarz, F., Kan, Y. W., Goldfine, I. D., Roth, R. A., and Rutter, W. J. (1985)
Cell 40, 747–758
38. Bishop, S. M., Ross, J. B. A., and Kohanski, R. A. (1999) Biochemistry 38,
3079–3089
39. Cleland, W. W. (1979) Methods Enzymol. 63, 103–128
40. Otwinowski, Z. (1993) in Proceedings of the CCP4 Study Weekend (Sawyer, L.,
Isaacs, N., and Bailey, S., eds) pp. 56–62, SERC Daresbury Laboratory,
Daresbury, United Kingdom
41. Navaza, J. (1994) Acta Crystallogr. Sect. A 50, 157–163
42. Bru¨nger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse-
Kunstleve, R. W., Jiang, J. S., Kuszewski, J., Nigles, M., Pannu, N. S., Read,
R. J., Rice, L. M., Simonson, T., and Warren, G. L. (1998) Acta Crystallogr.
Sect. D Biol. Crystallogr. 54, 905–921
43. Jones, T. A., Zou, J. Y., Cowan, S. W., and Kjeldgaard, M. (1991) Acta
Crystallogr. Sect. A 47, 110–119
44. Lane, M. D., Ronnett, G., Slieker, L. J., Kohanski, R. A., and Olson, T. L. (1985)
Biochimie (Paris) 67, 1069–1080
45. Kohanski, R. A., and Schenker, E. (1991) Biochemistry 30, 2406–2414
46. Kohanski, R. A. (1993) Biochemistry 32, 5773–5780
47. Maegley, K. A., Admiraal, S. J., and Herschlag, D. (1996) Proc. Natl. Acad. Sci.
U. S. A. 93, 8160–8166
48. Kirby, A. J., and Varvogolis, A. G. (1968) J. Chem. Soc. B 135–141
49. Herschlag, D., and Jencks, W. P. (1987) J. Am. Chem. Soc. 109, 4665–4674
50. Khan, S. A., and Kirby, A. J. (1970) J. Chem. Soc. B 1172–1182
Dissociative Transition State of the Insulin Receptor Kinase30398
atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom

Mais conteúdo relacionado

Mais procurados

Omar Quintero Resume early Sept 2016 linkedin
Omar Quintero Resume early Sept 2016 linkedinOmar Quintero Resume early Sept 2016 linkedin
Omar Quintero Resume early Sept 2016 linkedin
Omar Quintero-Monzon
 
Ic0506815
Ic0506815Ic0506815
Ic0506815
inscore
 
Sub-optimal phenotypes of double-knockout of E.coli
Sub-optimal phenotypes of double-knockout of E.coliSub-optimal phenotypes of double-knockout of E.coli
Sub-optimal phenotypes of double-knockout of E.coli
Dr Fatumina Abukar
 
------- Enzymology --------
 -------  Enzymology    -------- -------  Enzymology    --------
------- Enzymology --------
aqeel Hadithe
 
Ben Kelty Summer Research Poster Presentation
Ben Kelty Summer Research Poster Presentation Ben Kelty Summer Research Poster Presentation
Ben Kelty Summer Research Poster Presentation
Benjamin Kelty
 
acs.jmedchem.5b00495
acs.jmedchem.5b00495acs.jmedchem.5b00495
acs.jmedchem.5b00495
Justin Murray
 

Mais procurados (20)

SHMT_ABB_1996
SHMT_ABB_1996SHMT_ABB_1996
SHMT_ABB_1996
 
JB REU Report
JB REU ReportJB REU Report
JB REU Report
 
BBL2
BBL2BBL2
BBL2
 
J. Biol. Chem.-2006
J. Biol. Chem.-2006J. Biol. Chem.-2006
J. Biol. Chem.-2006
 
Omar Quintero Resume early Sept 2016 linkedin
Omar Quintero Resume early Sept 2016 linkedinOmar Quintero Resume early Sept 2016 linkedin
Omar Quintero Resume early Sept 2016 linkedin
 
Altered substrate-specificity-of-the-pterygoplichthys-sp.-loricariidae-cyp1 a...
Altered substrate-specificity-of-the-pterygoplichthys-sp.-loricariidae-cyp1 a...Altered substrate-specificity-of-the-pterygoplichthys-sp.-loricariidae-cyp1 a...
Altered substrate-specificity-of-the-pterygoplichthys-sp.-loricariidae-cyp1 a...
 
Following the Evolution of New Protein Folds via Protodomains
Following the Evolution of New Protein Folds via ProtodomainsFollowing the Evolution of New Protein Folds via Protodomains
Following the Evolution of New Protein Folds via Protodomains
 
WBC 2016 Poster
WBC 2016 PosterWBC 2016 Poster
WBC 2016 Poster
 
Smith_PNAS_2016
Smith_PNAS_2016Smith_PNAS_2016
Smith_PNAS_2016
 
8.1 metabolism
8.1 metabolism8.1 metabolism
8.1 metabolism
 
Study on In Vitro Kinetic Characterization of Transporter-Mediated Permeability
Study on In Vitro Kinetic Characterization of Transporter-Mediated PermeabilityStudy on In Vitro Kinetic Characterization of Transporter-Mediated Permeability
Study on In Vitro Kinetic Characterization of Transporter-Mediated Permeability
 
Ic0506815
Ic0506815Ic0506815
Ic0506815
 
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
 
Sub-optimal phenotypes of double-knockout of E.coli
Sub-optimal phenotypes of double-knockout of E.coliSub-optimal phenotypes of double-knockout of E.coli
Sub-optimal phenotypes of double-knockout of E.coli
 
2012Brookins_TeQuion
2012Brookins_TeQuion2012Brookins_TeQuion
2012Brookins_TeQuion
 
------- Enzymology --------
 -------  Enzymology    -------- -------  Enzymology    --------
------- Enzymology --------
 
Ben Kelty Summer Research Poster Presentation
Ben Kelty Summer Research Poster Presentation Ben Kelty Summer Research Poster Presentation
Ben Kelty Summer Research Poster Presentation
 
acs.jmedchem.5b00495
acs.jmedchem.5b00495acs.jmedchem.5b00495
acs.jmedchem.5b00495
 
95.full
95.full95.full
95.full
 
CcP_Cytc_PNAS_2004
CcP_Cytc_PNAS_2004CcP_Cytc_PNAS_2004
CcP_Cytc_PNAS_2004
 

Destaque

Avatardesign ved Tatiana
Avatardesign ved TatianaAvatardesign ved Tatiana
Avatardesign ved Tatiana
henrikjuel
 
Segurança no trabalho
Segurança no trabalhoSegurança no trabalho
Segurança no trabalho
micaelaaveiro
 
Præsentation til Crach courses
Præsentation til Crach coursesPræsentation til Crach courses
Præsentation til Crach courses
henrikjuel
 
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagarMedia blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
Daytona
 
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
Daytona
 

Destaque (17)

McKonly & Asbury Webinar - Fraud Not If, But When
McKonly & Asbury Webinar - Fraud Not If, But WhenMcKonly & Asbury Webinar - Fraud Not If, But When
McKonly & Asbury Webinar - Fraud Not If, But When
 
Flipster for Chaminade University
Flipster for Chaminade UniversityFlipster for Chaminade University
Flipster for Chaminade University
 
Региональное интернет-издание. Опыт неполумиллионника
Региональное интернет-издание. Опыт неполумиллионника Региональное интернет-издание. Опыт неполумиллионника
Региональное интернет-издание. Опыт неполумиллионника
 
Monarchy of counterfeit drugs towards global health care terrorism
Monarchy of counterfeit drugs towards global health care terrorismMonarchy of counterfeit drugs towards global health care terrorism
Monarchy of counterfeit drugs towards global health care terrorism
 
Dinesh Puri Resume
Dinesh Puri ResumeDinesh Puri Resume
Dinesh Puri Resume
 
biteck
biteckbiteck
biteck
 
Begrafenisondernemer in rotterdam
Begrafenisondernemer in rotterdamBegrafenisondernemer in rotterdam
Begrafenisondernemer in rotterdam
 
SANSKAR
SANSKARSANSKAR
SANSKAR
 
Avatardesign ved Tatiana
Avatardesign ved TatianaAvatardesign ved Tatiana
Avatardesign ved Tatiana
 
Segurança no trabalho
Segurança no trabalhoSegurança no trabalho
Segurança no trabalho
 
Congreso Nacional de Educación en Rosario
Congreso Nacional de Educación en RosarioCongreso Nacional de Educación en Rosario
Congreso Nacional de Educación en Rosario
 
Præsentation til Crach courses
Præsentation til Crach coursesPræsentation til Crach courses
Præsentation til Crach courses
 
Nature recall save water- a life heritage
Nature recall  save water- a life heritageNature recall  save water- a life heritage
Nature recall save water- a life heritage
 
EXPRESSIONS UN-DEFINED
EXPRESSIONS UN-DEFINEDEXPRESSIONS UN-DEFINED
EXPRESSIONS UN-DEFINED
 
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagarMedia blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
Media blir mobilt 2009 - Joel Bergqvist, Daytona - 3644 dagar
 
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
Topp 10 webbtrender för 2010 - Webbfredag hos Daytona 2009-04-17
 
the-key-of-david
the-key-of-davidthe-key-of-david
the-key-of-david
 

Semelhante a Ablooglu et al JBC 2000

Ablooglu et al 2001 Nat Str Biol
Ablooglu et al 2001 Nat Str BiolAblooglu et al 2001 Nat Str Biol
Ablooglu et al 2001 Nat Str Biol
Ararat Ablooglu
 
05-11-16 Thesis Progress
05-11-16 Thesis Progress05-11-16 Thesis Progress
05-11-16 Thesis Progress
Shelby Kratt
 
Spectroscopic and ITC studies of binding of Ferulic acid with BSA
Spectroscopic and ITC studies of binding of Ferulic acid with BSASpectroscopic and ITC studies of binding of Ferulic acid with BSA
Spectroscopic and ITC studies of binding of Ferulic acid with BSA
Dr Himanshu Ojha
 
EUPHYTICA_Minh Luan NGUYEN_2013
EUPHYTICA_Minh Luan NGUYEN_2013EUPHYTICA_Minh Luan NGUYEN_2013
EUPHYTICA_Minh Luan NGUYEN_2013
nguyen luan
 
Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011
Alex Kiselyov
 
khairia-m-youssef-future-university-egypt.ppt
khairia-m-youssef-future-university-egypt.pptkhairia-m-youssef-future-university-egypt.ppt
khairia-m-youssef-future-university-egypt.ppt
taoufikakabli1
 
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
rajmaha9
 
The Thiazide-sensitive NaCl
The Thiazide-sensitive NaClThe Thiazide-sensitive NaCl
The Thiazide-sensitive NaCl
Avin Snyder
 

Semelhante a Ablooglu et al JBC 2000 (20)

Ablooglu et al 2001 Nat Str Biol
Ablooglu et al 2001 Nat Str BiolAblooglu et al 2001 Nat Str Biol
Ablooglu et al 2001 Nat Str Biol
 
9693128
96931289693128
9693128
 
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
 
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
 
05-11-16 Thesis Progress
05-11-16 Thesis Progress05-11-16 Thesis Progress
05-11-16 Thesis Progress
 
4
44
4
 
Lysine Crotonylation.pptx
Lysine Crotonylation.pptxLysine Crotonylation.pptx
Lysine Crotonylation.pptx
 
Spectroscopic and ITC studies of binding of Ferulic acid with BSA
Spectroscopic and ITC studies of binding of Ferulic acid with BSASpectroscopic and ITC studies of binding of Ferulic acid with BSA
Spectroscopic and ITC studies of binding of Ferulic acid with BSA
 
OBC- 2013
OBC- 2013OBC- 2013
OBC- 2013
 
EUPHYTICA_Minh Luan NGUYEN_2013
EUPHYTICA_Minh Luan NGUYEN_2013EUPHYTICA_Minh Luan NGUYEN_2013
EUPHYTICA_Minh Luan NGUYEN_2013
 
PROTEOMICS.pptx
PROTEOMICS.pptxPROTEOMICS.pptx
PROTEOMICS.pptx
 
Vaillencourt
VaillencourtVaillencourt
Vaillencourt
 
Thermodynamic and biophysical characterization of cytochrome p450 bio i from ...
Thermodynamic and biophysical characterization of cytochrome p450 bio i from ...Thermodynamic and biophysical characterization of cytochrome p450 bio i from ...
Thermodynamic and biophysical characterization of cytochrome p450 bio i from ...
 
StAR
StARStAR
StAR
 
Crystal
CrystalCrystal
Crystal
 
Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011Polyalkoxybenzenes JMC2011
Polyalkoxybenzenes JMC2011
 
biomedicina_Dolzan_TransporterVariability.pdf
biomedicina_Dolzan_TransporterVariability.pdfbiomedicina_Dolzan_TransporterVariability.pdf
biomedicina_Dolzan_TransporterVariability.pdf
 
khairia-m-youssef-future-university-egypt.ppt
khairia-m-youssef-future-university-egypt.pptkhairia-m-youssef-future-university-egypt.ppt
khairia-m-youssef-future-university-egypt.ppt
 
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
Simplified receptor based pharmacophore approach to retrieve potent ptp lar i...
 
The Thiazide-sensitive NaCl
The Thiazide-sensitive NaClThe Thiazide-sensitive NaCl
The Thiazide-sensitive NaCl
 

Ablooglu et al JBC 2000

  • 1. Probing the Catalytic Mechanism of the Insulin Receptor Kinase with a Tetrafluorotyrosine-containing Peptide Substrate* Received for publication, April 25, 2000, and in revised form, June 22, 2000 Published, JBC Papers in Press, June 26, 2000, DOI 10.1074/jbc.M003524200 Ararat J. Ablooglu‡§, Jeffrey H. Till§¶, Kyonghee Kimʈ, Keykavous Parangʈ, Philip A. Coleʈ, Stevan R. Hubbard¶, and Ronald A. Kohanski‡ ** From the ‡Mount Sinai School of Medicine, Department of Biochemistry and Molecular Biology, New York, New York, 10029 ¶Skirball Institute of Biomolecular Medicine and Department of Pharmacology, New York University School of Medicine, New York, New York 10016, and ʈThe Johns Hopkins University School of Medicine, Department of Pharmacology and Molecular Sciences, Baltimore, Maryland 21205 The interaction of a synthetic tetrafluorotyrosyl pep- tide substrate with the activated tyrosine kinase do- main of the insulin receptor was studied by steady-state kinetics and x-ray crystallography. The pH-rate profiles indicate that the neutral phenol, rather than the chem- ically more reactive phenoxide ion, is required for en- zyme-catalyzed phosphorylation. The pKa of the tet- rafluorotyrosyl hydroxyl is elevated 2 pH units on the enzyme compared with solution, whereas the phenoxide anion species behaves as a weak competitive inhibitor of the tyrosine kinase. A structure of the binary enzyme- substrate complex shows the tetrafluorotyrosyl OH group at hydrogen bonding distances from the side chains of Asp1132 and Arg1136 , consistent with elevation of the pKa. These findings strongly support a reaction mechanism favoring a dissociative transition state. Protein kinases are enzymes that catalyze transfer of the ␥-phosphoryl group from ATP to the alcoholic side chains of amino acid residues on their substrate proteins and peptides. They play important roles in signaling and have been subject to intensive cell biologic, enzymologic, structural, and kinetic scrutiny. Elucidation of the chemical reaction mechanisms of these enzymes is proving to be important in defining the details of their regulation and for designing inhibitors. The reaction mechanism of phosphoryl transfer reactions can favor either an associative or dissociative pathway (Fig. 1). In the extremes, these are distinguished by dependence on versus independence from nucleophilicity of the phosphoryl acceptor, respectively, distances between the ␥-phosphorus atom and the attacking and leaving group oxygens, and other factors (1–7). The associative pathway would benefit from proton abstraction prior to formation of the transition state so that a more reactive oxygen anion is formed, whereas that proton is retained in the dissociative transition state. Earlier kinetic and crystallo- graphic studies of protein kinases, especially the cAMP-de- pendent protein kinase, demonstrated the importance of a con- served aspartyl residue that is essential for the enzyme- catalyzed reaction (8–10). The location of this Asp166 side chain in the cAMP-dependent protein kinase crystal structure was close enough to the serine hydroxyl that it could create a more reactive nucleophile by removal of the seryl or threonyl proton, in accordance with an associative mechanism (8, 9), but other kinetic evidence was interpreted as favoring a dissociative pathway (5, 11, 12, 14). Never the less, similar positioning of the conserved aspartyl side chain in other protein kinase struc- tures has been offered as further supporting evidence for pro- ton abstraction, thus generating a more reactive entering oxy- gen as a basis for catalysis (e.g. see Ref. 15). On the other hand, as discussed by Mildvan (2), distances as short as 2.7 Å be- tween the entering oxygen and the ␥-phosphorus could be es- timated from modeling studies (16) and as long as 5.3 Å from NMR (17); the latter indicates a reaction pathway with a more dissociative character. Furthermore, the mechanism of phos- phoryl transfer in chemical reactions of phosphate monoesters, as models for enzyme-catalyzed reactions, appears to be disso- ciative with relatively little contribution of nucleophilicity to the transition state (18, 19). On the premise that chemical and enzymatic reactions differ in stabilization but not character of the transition state (20–26), it would seem that protein kinases should follow a more dissociative pathway. For protein tyrosine kinases, the ability to vary nucleophi- licity by aromatic ring substitution on tyrosine facilitated a chemical analysis of the phosphoryl transfer reaction catalyzed by C-terminal Src kinase (Csk).1 Analyzing the pH-independ- ent kinetic parameters for a series of substituted tyrosine de- rivatives with pKas ranging from 5.2 to 10, it was shown that kcat and kcat/Km were relatively independent of nucleophilicity of the attacking phenolic residue, yielding a very small Bron- sted nucleophile coefficient (0 Ͻ ␤nuc Ͻ 0.1) (6). The pH-rate profiles also demonstrated that the chemically more reactive phenoxide anion species was enzymatically less reactive than the neutral phenol species. Further studies with Csk revealed that the reverse reaction, where the fluorotyrosyl peptide is the leaving group, gave a Bronsted leaving group coefficient (␤1g) of Ϫ0.3 (3). This supports protonation of the phenolic O␩ in both directions. Thus, evidence from the forward and reverse reac- tions together suggest that the character of the transition state for Csk is mostly dissociative (26). Fluorotyrosyl analogs have also been used with the insulin receptor (IR) tyrosine kinase (27, 28), but the opposite conclu- sion had been reached in those studies; i.e. the transition state * This work was supported by Grant DK50074 from the National Institutes of Health (NIH) (to R. A. K.), Grant DK52916 from NIH (to S. R. H.), by the Winston Foundation (to K. K.), and by the Burroughs Welcome Fund and NIH Grant CA74305 (to P. A. C.). The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. § Equal contributions were made by these authors. ** To whom correspondence should be addressed. Tel.: 212-241-7288; Fax: 212-996-7214; E-mail: Ronald.Kohanski@mssm.edu. 1 The abbreviations used are: Csk, C-terminal Src kinase; HPLC, high performance liquid chromatography; IR, insulin receptor; IRKD, cytoplasmic kinase domain of the IR; IRK3P, tris-phosphorylated core kinase domain of the human insulin receptor; IRK, insulin receptor kinase; Fmoc, N-(9-fluorenyl)methoxycarbonyl. THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 275, No. 39, Issue of September 29, pp. 30394–30398, 2000 © 2000 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A. This paper is available on line at http://www.jbc.org30394 atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
  • 2. for phosphoryl transfer seemed to be associative. Therefore, the analysis reported here was undertaken to specifically address this discrepancy between Csk and IR, making use of the fol- lowing three advances since the earlier work of Graves and co-workers (27, 28): 1) There is now a better understanding of the substrate-specificity determinants for the insulin receptor (29–32), which was a factor in design of the peptide substrates. 2) The recombinant cytoplasmic kinase domain of the insulin receptor (IRKD) can be expressed as a highly purified protein whose autophosphorylation state can be fixed experimentally (31–34). 3) The ternary complex of the catalytic core with a tyrosyl peptide substrate and MgATP analog has been deter- mined, which establishes a distance between O␩ of the reactive Tyr and the ␥-phosphorus atom of the adenine nucleotide (32). Our results with the autophosphorylated and maximally acti- vated IRKD support a dissociative mechanism for this protein kinase, show an increase in pKa of the enzyme-bound fluoroty- rosyl group compared with aqueous solution, and demonstrate that this fluorotyrosyl side chain can form hydrogen bonds virtually identical to those between the unmodified tyrosyl residue of the substrate and the catalytic core of the kinase. EXPERIMENTAL PROCEDURES General—Dithiothreitol (Sigma Ultra), the disodium salt of ATP (from equine muscle, catalog number A-5394), and bovine serum albu- min (radioimmunoassay grade) were purchased from Sigma; hydroge- nated Triton X-100 (protein grade) was from Calbiochem; EDTA was from Fluka; Tris acetate, Tris base, Tris HCl, and electrophoresis re- agents were from Roche Molecular Biochemicals; 1,3-bis[tris(hy- droxymethyl)-methylamino]propane (bis-tris propane) was from Sigma; magnesium acetate (MgAc2, enzyme grade) was from Fisher. Insect cell culture media and fetal bovine serum were from Life Technologies, Inc. Peptide Synthesis—L-Tetrafluorotyrosine was enzymatically pre- pared with tyrosine phenol-lyase as described previously and converted to the corresponding Fmoc derivative (6). Fmoc-L-tetrafluorotyrosine was subsequently incorporated in place of tyrosine in the previously described 18-residue IRKD substrate peptide, IRS727 (35), with the amino acid sequence KKKLPATGDYMNMSPVGD (36) modified from the rat sequence (residues 718–735, with Gly718 and Leu720 changed to Lys), using solid phase peptide synthesis providing fluoro-IRS727. IRS727 and fluoro-IRS727 were prepared by solid phase peptide syn- thesis using the Fmoc strategy on a Rainin PS-3 instrument. After purification by reversed phase HPLC (preparative C-18 column), pep- tides were judged to be greater than 95% pure. The peptide substrate IRS939 (REETGSEEYMNMDLG, residues 931–945 from the rat IRS1 sequence) was prepared by Fmoc synthesis and purified by HPLC, as described previously (36). Peptides had free amino and carboxyl termini and included the consensus sequence D(E)YMXM whose interactions at the PϪ1 to Pϩ3 subsites on the catalytic core of the kinase have been established (32); Tyr is the P0 residue. Synthetic peptides and adenine nucleotide were dissolved in 50 mM TrisAc, pH 7.0, the pH value was readjusted to 7.0 with the addition of dilute acetic acid and/or NaOH, and the concentrations were adjusted to the desired values for stock solutions, which were determined spectrophotometrically using the fol- lowing extinction coefficients: ATP, ⑀ ϭ 15,300 cmϪ1 M Ϫ1 at 259 nm; IRS727 and IRS939, ⑀ ϭ 1300 cmϪ1 M Ϫ1 at 280 nm. These stocks were kept at Ϫ20 °C under frequent use or at Ϫ80 °C for storage. Integrity of the reagents was checked periodically by reverse phase HPLC (pep- tides) or by thin layer chromatography (nucleotides). The pKa of the phenolic group in fluoro-IRS727 was determined spectrophotometri- cally at 280 nm, at a peptide concentration of 0.25 mM. IRKD Expression and Purification—Baculovirus-encoding amino acid residues 953–1355 of the insulin receptor’s cytoplasmic kinase domain2 was used to express the IRKD in High Five™ cells (Invitro- gen). Enzyme purification, autophosphorylation, and storage were es- sentially as described previously (38). The purified protein was quan- tified spectrophotometrically at 280 nm (⑀ ϭ 40,200 cmϪ1 M Ϫ1 for the unphosphorylated IRKD, and ⑀ ϭ 35,200 cmϪ1 M Ϫ1 for the autophos- phorylated IRKD). The autophosphorylation stoichiometry for the IRKD was 5.4 mol phosphate/mol enzyme, and the enzyme was deter- mined to be maximally activated (specific activity of 800 mol phospho- IRS939/min/mol IRKD at 1 mM ATP, 20 mM MgAc2). The absence of an unphosphorylated or monophosphorylated activation loop, and the presence of 20% bis and 80% tris phosphorylation of the activation loop were established in the purified phosphoprotein by endoproteinase Lys-C digestion and HPLC mapping (not shown). Peptide Phosphorylation Assays—Steady-state kinetic parameters were determined with the autophosphorylated IRKD as the enzyme, using an HPLC-based assay (36). Quantification of the apo and phos- phopeptide were done by peak area integration with Chrome Perfect Software (Justice Innovations, Palo Alto, CA). Dead-end inhibition studies were done over 0.0–0.63 mM fluoro-IRS727, constant 0.08 mM ATP, and variable IRS939 (0.04–0.2 mM). Reactions were performed at room temperature in 50 mM bis-tris propane, 5 mM dithiothreitol, 0.05% bovine serum albumin (w/v), 2–8 nM phospho-IRKD, and 20 mM MgAc2 and at the pH values indicated in the figure legends. The pH of stock solutions of peptide, ATP, dithiothreitol, and bovine serum albumin solutions were adjusted accordingly. The reactions were initiated with the addition of peptide substrate. Kinetic parameters were determined from the best global fits of the data. The pH dependence of the observed rate constant kcat,obs were fit assuming a single proton dissociation. Plots of log(kcat) versus pH values for fluoro-IRS727 were fit to the equation, log(kcat,obs) ϭ log(kcat/͓1 ϩ K/H͔) (Eq. 1) where kcat is the pH-independent rate constant, H is the proton concen- tration, and K is the dissociation constant for the ionizable group that is active in the protonated state (pKa ϭ log(1/K)). Competitive inhibition kinetic analysis was done by fitting all of the data points to the linear competitive inhibition equation of KinetAsyst IITM based on the algorithms of Cleland (39), v ϭ VM*S/͓KM͑1 ϩ I/Ki͒ ϩ S͔ (Eq. 2) using a non-linear least squares approach. The fixed substrate was assumed to be saturating. Kinetic constants Ϯ S.E. are reported here. X-ray Crystallographic Studies—Expression and purification of the tris-phosphorylated core kinase domain of the human insulin receptor (IRK3P, residues 978–1283) has been described previously (32). Crys- tals were grown at 4 °C by vapor diffusion in hanging drops containing 2.0 ␮l of protein solution (6.5 mg/ml IRK3P and 1 mM fluoro-IRS727) and 2.0 ␮l of reservoir buffer (18% polyethylene glycol 8000, 100 mM Tris-HCl, pH 8.5, and 15% ethylene glycol). Small crystals (Ͻ50 ␮m) were obtained initially and used for microseeding to grow larger crys- tals (150–200 ␮m). Crystals belong to the trigonal space group P3221 with unit cell dimensions a ϭ b ϭ 71.8 Å, and c ϭ 125.8 Å when frozen. There is one molecule in the asymmetric unit, and the solvent content is 48% (assuming a partial specific volume of 0.74 cm3 /g). Two data sets were used for structural analysis. One data set was collected on a Rigaku RU-200 rotating anode equipped with a Rigaku R-AXIS IIC image plate detector. The other data set was collected on beamline X4A, National Synchrotron Light Source, equipped with an ADSC Quantum-4 CCD detector. Crystals were flash-cooled in a dry 2 Numbering of the IR and IRKD is from Ebina et al. (37). FIG. 1. Associative and dissociative reaction mechanisms for phosphoryl transfer from ATP to the phenolic hydroxyl of tyro- sine. The upper pathway depicts an associative transition state in which the entering and leaving groups are closer to the ␥-phosphorous atom and where the axial bond number to the entering O␩ could be as high as one, versus a dissociative transition state, shown in the lower pathway, in which these groups are farther apart, and the axial bond number could be as low as zero. Dissociative Transition State of the Insulin Receptor Kinase 30395 atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
  • 3. nitrogen stream at Ϫ160 °C. Because of only ϳ80% completeness of the synchrotron data, the data were merged with the rotating anode data, giving an overall Rsym of 8.3%, an overall I/␴I of 24.4, and an overall completeness of 91% for data between 30.0 and 2.6 Å. Data were processed using DENZO and SCALEPACK (40). A molecular replacement solution was found with AMoRE (41) using Protein Data Bank entry 1IR3 (32) as a search model. Rigid-body, positional, and B-factor refinement were carried out using CNS (42). Model building was performed using O (43). The final model (2464 atoms) has a crystallographic R value of 24% (free R value of 31%) with root mean square deviations of 0.009 Å for bonds and 1.5° for angles (30.0–2.6 Å). RESULTS The insulin receptor is not a soluble kinase; it is a hetero- meric transmembrane protein whose intracellular kinase ac- tivity is regulated by extracellular insulin binding (44). To perform these studies in a manner similar to those done with the soluble enzyme Csk, we have utilized the highly purified and activated (autophosphorylated) cytoplasmic portion of the IR. The IRKD was previously shown to be a good model for the holomeric receptor’s basal and activated states, and these fea- tures are retained by the conserved catalytic core used in crystallization (33, 45, 46). The tetrafluorotyrosyl peptide sub- strate (fluoro-IRS727) was chosen for these studies, because the unmodified form (IRS727) was cocrystallized with the cat- alytic core of the IRKD (32). The peptide-inhibition studies were done using IRS939 as the substrate because of signal to noise considerations in the peptide phosphorylation assays. Kinetic Analysis—Kinase assays of IRS727 using the IRKD revealed somewhat different kinetic parameters than those reported previously for the full-length IR (35). This was possi- bly because of differences in assay procedures and kinase or reagent purity. Here, global fitting of the data from our steady- state assays gave the following kinetic parameters at pH 7: kcat ϭ 14 Ϯ 2 sϪ1 , and Km ϭ 0.18 Ϯ 0.02 mM IRS727. Analysis of fluoro-IRS727 showed that it was apparently a poorer sub- strate than IRS727 at pH 7, with kcat ϭ 0.5 Ϯ 0.1 sϪ1 , and Km ϭ 0.21 Ϯ 0.05 mM. Given the observed rate reduction, and previous claims that no phosphorylation was detected with a different fluorotyrosyl peptide (27, 28), structural identity of the phosphorylated fluoro-IRS727 product was confirmed rig- orously by mass spectrometric analysis after HPLC purifica- tion. In addition, Ͼ95% conversion to phosphorylated fluoro- IRS727 was observed after prolonged incubation with activated IRKD, determined by the HPLC-based assay (not shown). To assess further the significance of these findings, pH-rate profiles were performed with peptide substrates for IRKD. Preliminary experiments confirmed that IRKD enzymatic ac- tivity versus time appeared stable between pH values of 6 and 10. In this range, phosphorylation of IRS727 showed little change in kcat (Fig. 2) or kcat/Km (not shown)3 with increasing pH values. In contrast, fluoro-IRS727 showed a marked 18-fold decrease in kcat with increasing pH values; the highest activity was observed at pH 6.0, and a pKa ϭ 7.6 was calculated for the basic limb (Fig. 2). These results suggest that a group present in fluoro-IRS727 but not in IRS727 requires protonation for enzymatic processing. This group is likely to be the tetrafluo- rotyrosine side chain. The pKa ϭ 7.6 is higher than the pKa ϭ 5.7 measured spectrophotometrically for the fluoro-IRS727 in aqueous solution (Fig. 2, inset). This implies that the tetrafluo- rotyrosine’s pKa is higher when the substrate is bound to the enzyme. At pH 7.0, protonated fluoro-IRS727 in solution is only 5% of the total, calculated from the Henderson-Hasselbach equation and the pKa of 5.7 of the free peptide. The value kcat/Km, which reflects the protonation equilibrium of the free substrate, should be 20-fold greater than the observed value, because only the neutral phenolic form is the substrate. With this correction it can be calculated that this tetrafluorotyrosyl peptide is only 2-fold less efficient as a substrate than the tyrosyl peptide. To determine whether the phenolate form of the peptide would bind to the enzyme, we measured the inhibitory proper- ties of fluoro-IRS727 at pH 9, where the tetrafluorophenolic group would be fully deprotonated in free solution, and the peptide would be minimally active as substrate (kcat ϭ 0.04 sϪ1 ) compared with the varied peptide substrate IRS939 (kcat ϭ 12 Ϯ 2 sϪ1 ; Km 0.040 Ϯ 0.002 mM). Fluoro-IRS727 is a linear competitive inhibitor against IRS939, with Ki ϭ 0.59 Ϯ 0.14 mM determined from a global fit of the data (Fig. 3). Therefore the phenolate form of fluoro-IRS727 binds weakly to the kinase. Crystal Structure of the Binary Complex—To gain better insight into the interaction of the tetrafluorotyrosyl group with active site residues in the kinase, the crystal structure of the fluoro-IRS727 complexed with the IRKD core was determined at 2.6-Å resolution (Fig. 4). When compared with the ternary complex of the autophosphorylated core with the unmodified IRS727 and Mg-5Ј-adenylyl imidodiphosphate (32), the P0 res- idue (the reactive tyrosine) of the substrate peptide and key residues in the active site with which it makes contact are virtually unchanged (root mean square deviation of 0.2 Å for the side-chain and main-chain atoms of residues Asp1132 and Arg1136 and the P0 tetrafluorotyrosine). Importantly, the dis- tance between the phenolic hydroxyl and Asp1132 is nearly the same in the fluoro-IRS727 complex (2.8 Å) as in the ternary complex (2.7 Å), indicating that a hydrogen bond is made be- tween these groups in both cases. In addition, the distance between the phenolic O␩ and the N␩ of Arg1136 is unperturbed, and the hydrogen bond between N␩ of Arg1136 and O␦2 of Asp1132 is also maintained. Together these describe a triangle of hydrogen bonding interactions (Fig. 4). This structure is compatible with the significant elevation of the phenolic pKa in the complex compared with free solution. DISCUSSION The transition states in phosphoryl transfer reactions cata- lyzed by enzymes have been the subject of lively discourse over the past several years (1–7, 18–25, 47). Mounting evidence suggests that these enzymes affect catalysis by enhancing but not fundamentally altering the non-enzymatic mechanisms. 3 It was not possible to precisely measure the kcat/Km dependence on pH because of inadequate signal to noise ratios and the tendency of activated IRKD to autodephosphorylate at high enzyme concentration, which would have been employed to increase net product formation and improve the signal. FIG. 2. The pH dependence of kcat for fluoro-IRS727. The log(k- cat) was plotted versus the reaction pH values for IRS727 (squares) and for fluoro-IRS727 (circles). The inset shows the spectrophotometric ti- tration of fluoro-IRS727 (triangles). The dotted lines are the best theo- retical fit for these data, assuming a single ionizable group in both cases. Dissociative Transition State of the Insulin Receptor Kinase30396 atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
  • 4. Among the experimental approaches most consistently applied to chemical and enzymatic reactions has been linear free en- ergy relationships to probe the mechanistic role of nucleophi- licity (1, 18). Measurements of ␤nuc Յ 0.3 for non-enzymatic phosphoryl transfer reactions involving phosphate monoesters suggested only a small impact of nucleophilicity in formation of the transition state and therefore favored a dissociative mech- anism (1, 13, 18, 48–50). Extending this conclusion to protein kinases would suggest that proton abstraction, to generate the chemically more reactive anionic oxygen, should not be a sig- nificant feature of enzyme-catalyzed phosphoryl transfer. This hypothesis can be tested directly using substrates for protein tyrosine kinases, which offer the unusual opportunity to alter nucleophilicity of the entering oxygen in an enzymatic phosphoryl transfer reaction by modification of the aromatic ring in the substrate (6, 25, 26). Systematic substitution of hydrogens with fluorine confers a broad range of pKa effects and allows Bronsted coefficients to be determined. This ap- proach has provided compelling evidence to support a dissocia- tive transition state for the reaction catalyzed by Csk (6, 26) but an associative transition state for the reaction catalyzed by IR (27, 28). There were two important concerns not addressed specifically in the study on Csk that are addressed here and that allowed us to re-evaluate the basis for an apparent dis- crepancy between Csk and IRK catalytic mechanisms. First, because the fluorine atomic radius is only 0.2 Å greater than the hydrogen atomic radius, these substitutions should not impose steric constraints on orientation of the P0 residue in the active site. That has been confirmed by comparison of two structures, showing the virtual absence of any positional dif- ference between the enzyme-bound P0 tetrafluorotyrosine, re- ported here, and the enzyme-bound P0 tyrosine reported pre- viously by Hubbard (32). Indeed, this shows that positioning of the phosphoryl-acceptor oxygen occurs even in the absence of MgATP (or analog), at least in a ground state binary complex. This structural observation (Fig. 4) allows greater confidence in the significance of kinetic studies using fluorine-substituted tyrosyl peptides with respect to the phosphoryl transfer mech- anism of the natural tyrosine-containing substrates. Second, it was presumed that the pKa of the fluorotyrosyl derivatives may be altered in the enzyme-bound state (27). The kcat versus pH profile likely represents the pKa of the phenolic group when the substrate is bound to the enzyme (in the ternary complex). It is apparently elevated 1.5–2 pH units when compared with aqueous solution (Fig. 2). The increased pKa in the bound state is also supported by the x-ray structure, which shows a hydrogen bond present between Asp1132 and the tetrafluorotyrosine phenol. These crystals were grown at pH 8.5, and in order for the neutral phenol to be the major species in the complex at this pH we would expect a higher pKa in this complex (perhaps 2–3 units greater than for the free peptide). These crystals lack an ATP analog that could affect the pKa of the phenolic group. Additionally, the pKa in the solid state may be harder to assess given the presence of large quantities of polyethylene glycol and other factors that might govern crys- tallization. However, the physical basis for pKa elevation ob- served in these kinetic experiments could be the electrostatic repulsion that would result from simultaneous deprotonation of the substrate-hydroxyl and the conserved catalytic aspartyl side chain, whose proximity has been established for IRK (see Ref. 32 and Fig. 4). The elevated pKa may appear surprising at first glance, because the enzymatic reaction ultimately catalyzes proton removal concomitant with phosphoryl transfer. However, we can estimate ␤nuc Ϸ 0.1 by taking into account the fact that kcat/Km should represent all steps prior to the first irreversible step and that kcat/Km of fluoro-IRS727 (solution pKa ϭ 5.7) is only 2-fold less than IRS727 (the tyrosyl group in solution will have a pKa ϭ 10). This small value is in good agreement with the ␤nuc determined for protein tyrosine kinase Csk using a broader range of fluorotyrosyl derivatives (6, 26). Therefore, the reaction mechanism for IRK favors a dissociative pathway such that facilitation of proton removal early in the reaction coordinate is unnecessary, and apparently unfavorable, as dis- cussed above. Our finding that the phenolate form of fluoro- IRS727 was a weak competitive inhibitor and not a substrate further supports this view. It seems clear, therefore, that a neutral phenolic group of the P0 residue is required for enzy- matic processing by IRK, as with Csk, and the transition states for IRK and Csk are likely to be similar. This and other published evidence (1–3) suggests the pro- tein-tyrosine and serine-threonine kinases, as do other en- zymes involved in phosphoryl transfer of phosphate mo- noesters, likely favor dissociative transition states. FIG. 3. Competitive inhibition by fluoro-IRS727. Inhibition of IRS939 phosphorylation was determined at pH 9.0, and the data are presented as a double-reciprocal plot. Reaction conditions are given under “Experimental Procedures.” The concentrations of fluoro-IRS727 were 0 mM (solid circles), 0.32 mM (open circles), and 0.63 mM (inverted solid triangles). FIG. 4. Electron density map in the active site of the binary complex between IRK3P and the tetrafluorotyrosine-contain- ing peptide. Shown is an Fo Ϫ Fc omit map (30.0–2.6 Å, 3␴ contour) in which the side chains of Asp1132 , Arg1136 , and the tetrafluorotyrosine (Tyr(4F)) were omitted from the map calculation. Carbon atoms are colored yellow, oxygen atoms are red, nitrogen atoms are blue, and fluorine atoms are magenta. Hydrogen bond (black dashes) distances are indicated. Dissociative Transition State of the Insulin Receptor Kinase 30397 atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom
  • 5. Establishing the nature of the protein kinase transition states provides geometric and functional constraints for understand- ing the roles of individual residues in regulation of kinase activity, the factors that govern protein substrate selectivity, and the design of transition state analogs as selective inhibitors. REFERENCES 1. Herschlag, D., and Jencks, W. P. (1989) J. Am. Chem. Soc. 111, 7587–7596 2. Mildvan, A. S. (1997) Proteins 29, 401–416 3. Cole, P. A., Sondhi, D., and Kim, K. (1999) Pharmacol. Ther. 82, 219–229 4. Cole, P. A., Grace, M. R., Phillips, R. S., Burn, P., and Walsh, C. T. (1995) J. Biol. Chem. 270, 22105–22108 5. Zhou, J., and Adams, J. A. (1997) Biochemistry 36, 2977–2984 6. Kim, K., and Cole, P. A. (1997) J. Am. Chem. Soc. 119, 11096–11097 7. Grace, M., Walsh, C. T., and Cole, P. A. (1997) Biochemistry 36, 1874–1881 8. Knighton, D. R., Zheng, J., Ten Eyck, L. F., Ashford, V. A., Xuong, N.-H., Taylor, S. S., and Sowadski, J. M. (1991) Science 253, 407–413 9. Bossemeyer, D., Engh, R. A., Kinzel, V., Ponstingl, H., and Huber, R. (1993) EMBO J. 12, 849–859 10. Gibbs, C. S., and Zoller, M. J. (1991) J. Biol. Chem. 266, 8923–8931 11. Adams, J. A., and Taylor, S. S. (1992) Biochemistry 31, 8516–8522 12. Adams, J. A. (1996) Biochemistry 20, 10949–10956 13. Epstein, J., Plapinger, R. E., Michel, H. O., Cable, J. R., Stephani, R. A., Billington, C. A., and West, G. R. (1964) J. Am. Chem. Soc. 86, 3075–3084 14. Grant, B. D., and Adams, J. A. (1996) Biochemistry 35, 2022–2029 15. Lowe, E. D., Noble, M. E., Skamneli, V. T., Oikonomahos, N. G., Owen, D. J., and Johnson, L. N. (1997) EMBO J. 16, 6646–6658 16. Madhusudan, Trafny, E. A., Xuong, N. H., Adams, J. A., Ten Eyck, L. F., Taylor, S. S., and Sowadski, J. M. (1994) Protein. Sci. 3, 176–187 17. Granot, J., Mildvan, A. S., Bramson, H. N., and Kaiser, E. T. (1980) Biochem- istry 19, 3537–3543 18. Herschlag, D., and Jencks, W. P. (1989) J. Am. Chem. Soc. 111, 7597–7606 19. Admiraal, S. J., and Herschlag, D. (2000) J. Am. Chem. Soc. 122, 2145–2148 20. Knowels, J. R. (1980) Annu. Rev. Biochem. 49, 877–919 21. Ho, M., Bramson, H. N., Hansen, D. E., Knowels, J. R., and Kaiser, E. T. (1988) J. Am. Chem. Soc. 110, 2680–2688 22. Admiraal, S. J., and Herschlag, D. (1995) Chem. Biol. 2, 729–739 23. Florian, J., and Warshel, A. (1997) J. Am. Chem. Soc. 119, 5473–5474 24. Jones, J. P., Weiss, P. M., and Cleland, W. W. (1991) Biochemistry 30, 3624–3629 25. Cole, P. A., Burn, P., Takacs, B., and Walsh, C. T. (1994) J. Biol. Chem. 269, 30880–30887 26. Kim, K., and Cole, P. A. (1998) J. Am. Chem. Soc. 120, 6851–6858 27. Martin, B. L., Wu, D., Jakes, S., and Graves, D. J. (1990) J. Biol. Chem. 265, 7108–7111 28. Yuan, C.-J., Jakes, S., Elliott, S., and Graves, D. J. (1990) J. Biol. Chem. 265, 16205–16209 29. Songyang, Z., Carraway, K. L., Eck, M. J., Harrison, S. C., Feldman, R. A., Mohammadi, M., Schlessinger, J., Hubbard, S. R., Smith, D. P., Eng, C., Lorenzo, M. J., Ponder, B. A. J., Mayer, B. J., and Cantley, L. C. (1995) Nature 373, 536–539 30. Feener, E. P., Badur, J. M., King, G. L., Wilden, P. A., Sun, X., Kahn, C. R., and White, M. F. (1993) J. Biol. Chem. 268, 11256–11264 31. Wei, L., Hubbard, S. R., Hendrickson, W. A., and Ellis, L. (1995) J. Biol. Chem. 270, 8122–8130 32. Hubbard, S. R. (1997) EMBO J. 16, 5572–5581 33. Kohanski, R. A. (1993) Biochemistry 32, 5766–5772 34. Cann, A. D., and Kohanski, R. A. (1997) Biochemistry 36, 7681–7689 35. Shoelson, S. E., Chatterjee, S., Chaudhuri, M., and White, M. F. (1992) Proc. Natl. Acad. Sci. U. S. A. 89, 2027–2031 36. Cann, A. D., Wolf, I., and Kohanski, R. A. (1997) Anal. Biochem. 247, 327–332 37. Ebina, Y., Ellis, L., Jarnagin, K., Edery, M., Graf, L., Clauser, E., Ou, J., Masiarz, F., Kan, Y. W., Goldfine, I. D., Roth, R. A., and Rutter, W. J. (1985) Cell 40, 747–758 38. Bishop, S. M., Ross, J. B. A., and Kohanski, R. A. (1999) Biochemistry 38, 3079–3089 39. Cleland, W. W. (1979) Methods Enzymol. 63, 103–128 40. Otwinowski, Z. (1993) in Proceedings of the CCP4 Study Weekend (Sawyer, L., Isaacs, N., and Bailey, S., eds) pp. 56–62, SERC Daresbury Laboratory, Daresbury, United Kingdom 41. Navaza, J. (1994) Acta Crystallogr. Sect. A 50, 157–163 42. Bru¨nger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse- Kunstleve, R. W., Jiang, J. S., Kuszewski, J., Nigles, M., Pannu, N. S., Read, R. J., Rice, L. M., Simonson, T., and Warren, G. L. (1998) Acta Crystallogr. Sect. D Biol. Crystallogr. 54, 905–921 43. Jones, T. A., Zou, J. Y., Cowan, S. W., and Kjeldgaard, M. (1991) Acta Crystallogr. Sect. A 47, 110–119 44. Lane, M. D., Ronnett, G., Slieker, L. J., Kohanski, R. A., and Olson, T. L. (1985) Biochimie (Paris) 67, 1069–1080 45. Kohanski, R. A., and Schenker, E. (1991) Biochemistry 30, 2406–2414 46. Kohanski, R. A. (1993) Biochemistry 32, 5773–5780 47. Maegley, K. A., Admiraal, S. J., and Herschlag, D. (1996) Proc. Natl. Acad. Sci. U. S. A. 93, 8160–8166 48. Kirby, A. J., and Varvogolis, A. G. (1968) J. Chem. Soc. B 135–141 49. Herschlag, D., and Jencks, W. P. (1987) J. Am. Chem. Soc. 109, 4665–4674 50. Khan, S. A., and Kirby, A. J. (1970) J. Chem. Soc. B 1172–1182 Dissociative Transition State of the Insulin Receptor Kinase30398 atBiomedicalLibrary,UCSDonFebruary27,2007www.jbc.orgDownloadedfrom