Sie sind auf Seite 1von 12

Israel Journal of Chemistry

Vol. 49 2009 pp. 163–173


DOI: 10.1560/IJC.49.2.163

Minireview:
Enzymatic H Transfers: Quantum Tunneling and Coupled Motion from Kinetic
Isotope Effect Studies

Vanja Stojković and Amnon Kohen*


Department of Chemistry, University of Iowa, Iowa City, Iowa 52242, USA

(Received 18 November 2008 and in revised form 3 December 2008)

Abstract. Many hydrogen transfer processes exhibit nonclassical behavior


due to inherent quantum mechanical properties of the hydrogen. Investigation of
various enzymes under physiological conditions indicates that hydrogen transfer
processes often show significant quantum mechanical behavior. Traditionally, this
phenomenon was treated in terms of a tunneling correction to classical or semiclas-
sical models. However, more recently, it has been observed that increasing numbers
of enzymes yield data that cannot be rationalized by tunneling correction models.
Observations such as large kinetic isotope effects (KIEs ) with unusual temperature
dependence, isotope effects on Arrhenius preexponential factors with values differ-
ent from semiclassical predicted ranges, and small temperature-independent KIEs
for processes with significant energy of activation, could only be explained with full
tunneling models for H transfer. Full tunneling models presume that the solvent or
protein fluctuations generate a reactive configuration along the heavy-atom coordi-
nate, from which the hydrogen is transferred through quantum mechanical tunnel-
ing. These models are sometimes denoted as environmentally coupled tunneling
(ECT) or Marcus-like models, and they link protein dynamics to the catalyzed H
transfer. Several enzymatic systems (dihydrofolate reductase, thymidylate synthase,
and soybean lipoxygenase) are presented as case studies of proton, hydrogen, and
hydride transfer.

Introduction tion between the ground state and the TS is presumed


Many important processes that occur in biological sys- to be an equilibrium process that follows the Boltzmann
tems involve hydrogen transfer. These include processes distribution. The rate of the reaction is exponentially
ranging from proton transfers across the cell membranes proportional to the height of the energy barrier, which is
to the functional isomerization of intermediary metabo- closely related to the experimental energy of activation
lites, and H transfer between substrates and cofactors. (Ea) (eq 1).
The physical nature of the H transfer in the enzymatic
k = Ae–Ea/RT (1)
systems has been the subject of increasing interest in
the past decade. Historically, H transfer in enzymes was where R is the gas constant, T is the absolute tempera-
treated in terms of transition-state theory (TST), also ture, and A is the Arrhenius preexponential factor, which
known as absolute reaction rate theory. According to also includes the transmission coefficient from the TS
this theory the rate of a single reaction step is fully de- to product. TST has proved useful in explaining a wide
termined by the properties of the reactants (free energy range of chemical and biological reactions. However it
minimum) and the transition state (TS—a free energy cannot always be applied to electron and/or hydrogen
saddle point or a dividing surface between reactants and transfer reactions.3–6
products states), and it is independent of the path taken *Author to whom correspondence should be addressed.
between these states.1,2 The distribution of the popula- E-mail: amnon-kohen@uiowa.edu
164

Table 1. Calculated de Broglie’s wavelengths for different hy-


drogen isotopes from λ = h/(2 mE)1/2, where m is particle mass,
and E is its energy (here assumed to be 10 kJ/mol)
type of isotope de Broglie’s wavelength (Å)
protium, H 0.5
deuterium, D 0.31
tritium, T 0.25

Fig. 1. An example of the ground-state tunneling along the re- Like electron transfer reactions, hydrogen transfer
action coordinate. The picture depicts the probability of a parti- exhibits non-classical behavior due to its small mass
cle to tunnel, with a higher probability for the lighter isotope to
and consequent large uncertainty in its location and mo-
tunnel. Blue and red lines represent the probability functions for
mentum. To illustrate the importance of non-classical
lighter and heavier isotopes, respectively. Color visible online.
behavior of hydrogen, the De Broglie wavelength can
be calculated for each of the three hydrogen isotopes
(Table 1), and these can be compared to a typical dis-
Quantum mechanical (QM) effects that govern elec- tance traversed during H transfer in solution and bio-
tron transfer processes have been recognized for a long logical systems (~0.5 Å).
time, and arise from the particle’s wave-like properties.7 Consequently, the quantum nature of hydride transfer
For H transfer, QM effects at the ground state (e.g., zero has often been treated as a perturbation to TST. Howev-
point energies—ZPEs) have long been recognized as a er, several examples discussed here indicate that the tun-
major component of isotope effects (both on equilib- neling correction to TST is an oversimplified model.
rium and kinetics)8,9 and the contribution of QM tunnel- In this review, we discuss a specific methodol-
ing to TS and transmission factors has been discussed ogy used to study the nature of the H transfer and the
for gas phase and organic reaction since the 1960s.10 potential surface, viz., kinetic isotope effects (KIEs).
In general terms, QM tunneling occurs when the prob- Moreover, we present the development of theories for H
ability of finding a particle in the reactant well overlaps transfer, from the simple Bell correction to TST, to the
with the probability of finding the particle in the product full-tunneling models applicable to H transfers. Lastly,
well (Fig. 1.). A more rigorous description would be that we discuss the experimental data obtained for several
the wave function for the particle at the reactant well model systems.
(Ψr) interacts (i.e., non-zero Franck–Condon integral)
with the wave function of the particle at the product
Kinetic Isotope Effects (KIEs)—an
state (Ψp). In a ground state, the two new wave func-
experimental tool to study H transfer
tions (Ψr+Ψp and Ψr – Ψp) would lead to two energy
KIEs are widely used probes for studying chemical
levels, a phenomenon denoted as tunneling splitting. At
mechanisms due to their ability to provide information
a non-stable state (e.g., TS), this phenomenon will lower
about the reaction coordinate. The KIE is the ratio of
the effective energy barrier (effect on Ea) and critically
reaction rates of two reactants that differ only in their
affect the transmission coefficient, since the tunneling
isotopic composition (isotopologues). By using differ-
probability is governed by different factors than that of
ent isotopologues the mass of the particle of interest
the classical coefficient.
is changed without changing the electronic potential
Recognition of the quantum nature of the electron
surface, making it an exquisite tool for studying the
transfer required the development of a new model ex-
reaction coordinate. Two types of KIEs can be defined:
tending beyond the explanation given by classical TST.
primary (1º) KIEs, which refer to an isotopic substitu-
Rudolph Marcus won the Nobel Prize for proposing a
tion of an atom that is being transferred during the reac-
new model to describe electron transfer reactions, which
tion, and a secondary (2º) KIE, where the labeled atom
recognizes that the reaction coordinate is largely deter-
is proximal to the atom being transferred. Others and we
mined by the heavy-atom coordinates.3,4 This model
have used 2º KIEs extensively in studies of H-tunnel-
proposed that the free-energy barrier, ∆G*, depends
ing;11–15 however in this review, we will focus only on 1º
on λ, which is an environmental distortion that occurs
KIEs. 1º hydrogen KIEs proved to be particularly inter-
before the electron transfer (the reorganization energy),
esting because: (i) the mass ratio of its isotopes is larger
and ∆Gº, the reaction driving force (eq 2).
than for any other element, leading to large KIE values,
DG* = (DGº + l)2 / 4l (2) and (ii) the mass of hydrogen is small enough so that

Israel Journal of Chemistry 49 2009


165

the quantum effects are more likely to be significant.15–18


Here we will show how KIEs have been used not only
as a tool to probe the nature of the hydride transfer in
biological systems (e.g., enzymes), but also to investi-
gate the role of the whole protein and its dynamics in the
catalysis that occurs at the active site.
Interpretation of data from KIE experiments proves
to be quite challenging due to kinetic complexity, a
feature common to enzymatic systems where the iso-
topically insensitive kinetic steps (substrate binding,
conformational changes of intermediates, product re-
lease, etc.) mask the intrinsic KIE.19,20 Several methods
have been used to unmask intrinsic KIEs from observed
KIEs and to provide direct information about the chemi-
cal step in the context of kinetic cascade that involves
other steps.19–22 The magnitude of the intrinsic KIEs
compared to that expected from a non-tunneling model, Fig. 2. Graphical representation of the semiclassical model
the temperature dependence of the intrinsic KIEs com- indicating that the difference in the energy of activation (∆Ea)
bined with the difference in the activation energy be- for H, D, and T, result from their different zero-point energies
tween isotopes, and the isotope effect on the Arrhenius (ZPE) in the ground state (GS) and transition state (TS). The
GS-ZPE is constituted by all degrees of freedom but mostly
preexponential factors (AL/AH) can be used to examine
by the C–H stretching frequency and the TS-ZPE is consti-
the contribution of tunneling. The temperature depen-
tuted by all degrees of freedom orthogonal to the reaction
dence of KIEs is very useful because it can distinguish coordinate.
between data that can be explained with the tunneling
correction to TST-models, and the data that require full Table 2. Semiclassical limits for the isotope effect on Arrhe-
tunneling models. nius preexponential factors1,10
AH/AT AH/AD AD/AT
upper limit 1.7 1.4 1.2
Semiclassical kinetic isotope effects and
lower limit 0.3 0.5 0.7
tunneling corrections
Semiclassical* models predict that the isotopic rate dif-
ference for H transfer mostly arises from the difference where W represents the energy of the particle, and k is
in the zero point energies (ZPE) between the different Boltzman’s constant. Barrier penetration occurs below
isotopes at the ground state vs. transition state (Fig. 2).10 the classical transition state, and its effect is predicted
This model predicts a primary kH/kD KIE of 6.9 for C–H to be the most significant for the lightest isotope. This
cleavage at 298 K, if we assume a stretching frequency leads both to a primary kH/kD KIE which exceeds the
of ~3000 cm–1 for C–H bond and 2200 cm–1 for C–D semiclassically predicted value of 7 for C–H cleavage,
bond.1 A similar frequency of 1800 cm–1 for C–T gives and to isotopic differences in Ea and AL/AH which are
primary kH/kT = 18 at the same temperature. smaller than limits predicted by the semiclassical model
One of the first models that addressed the phenom- (Table 2).1,15,18 This approach led to a correction to the
enon of hydrogen tunneling, observed initially in or- semiclassical model, which proved useful in explaining
ganic solutions, was the Bell model.10 According to this a range of observed phenomena, such as inflated KIEs,
model, H transfer can be described in the form of the and an inverse isotope effect on Arrhenius prefactor ra-
semiclassical TST with the addition of a multiplier term tios as measured by variable-temperature KIEs.
(Qt) that indicates the probability that a particle will Another indication for the deviation from the clas-
move through an inverse parabolic barrier (eq 3). sical nature of the H transfer is the temperature depen-
kobs = Qt * kSC dence of KIEs. According to the Arrhenius equation the
W (3) rate of the reaction is exponentially proportional to Ea
Qt = kT e kT # G]W ge - kT
1 E
and the reciprocal absolute temperature (1/T) (eq 1).
Since the KIE is the ratio of the reactions’ rates, its tem-
perature dependency will follow:
*In this context, semiclassical means that quantum-mechanical
effects at ground states are taken into consideration, e.g., ZPE, kL AL DERT
a]H - Lg

= e (4)
but, QM effects close to the TS, e.g., tunneling, are ignored. kH AH

Stojković and Kohen / Enzymatic H Transfers


166

high temperature limit. It is important to note that inter-


(a) cept values close to unity do not necessarily indicate no
tunneling, because for systems with data in-between re-
gions 2 and 3, the model suggests significant tunneling
and intercept values close to unity.
TST-like models with or without tunneling correc-
tion, could, in several cases, rationalize temperature-de-
(b)
pendent KIEs,10 assuming a 1D rigid potential surface,
and large temperature-independent KIEs for systems
which showed no energy of activation for the isotopical-
ly-sensitive step. These models however fail to explain
temperature-independent small KIEs, with significant
energies of activation for the isotopically sensitive step.
One of the shortcomings of the Bell tunneling correction
Fig. 3. An Arrhenius plot of a hydrogen transfer applicable and similar TST-like models is that they exclusively fo-
for the TST-models with tunneling correction. (a) Arrhenius cus on the hydrogen reaction coordinate and ignore the
plot of reaction rates for two isotopes. (b) Arrhenius plot for contributory motions of the heavy-atom environment.
their KIEs. The KIE on the Arrhenius preexponential factors As will be presented in the following section, more
is an intercept of a tangent to a curve at the experimental tem- recent models treat H-tunneling as a multidimensional
perature ranges. Highlighted regions indicate three distinct
process.
systems: (1) no tunneling contribution, where AL/AH is close
to unity; (2) moderate tunneling contribution, where AL/AH is
During the last decade, several enzymatic systems
smaller than 1; (3) extensive tunneling contribution, where yielded data that could not be rationalized with the
AL/AH is larger than unity. Color visible online. simple Bell-correction model (Table 3). In several cases
KIEs greatly exceeded the semiclassical limit and in
most cases AL/AH surpassed unity. Moreover, some of
these systems were found to have a nonzero Ea for the
H transfer process, some with temperature-dependent
KIEs and some with temperature-independent KIEs. In
where H and L are the heavy and the light isotope, re- this review, we will specifically discuss three systems
spectively. For the majority of the enzymatic systems that required a full tunneling (Marcus-like) model in
only a narrow experimental temperature range is avail- order to explain experimental data.
able (0–80 ºC), thus plots of ln(k) vs. 1/T often appear
linear (Fig. 3). At high temperature regime, the slope
Marcus-like models
of the Arrhenius plot is exponentially proportional to
In an attempt to explain data obtained for systems where
∆Ea. At very low temperatures only tunneling contrib-
small temperature-independent KIEs with significant
utes significantly to rates due to the lack of thermal
activation energies were observed, several phenomeno-
energy (Fig. 3a, region 3).23,24 Consequently, the rates
logical models were developed. These models modified
are temperature independent, and KIEs are large (over
the Marcus theory for the electron transfer, by adding a
six orders of magnitude25) and temperature independent,
term that accounts for fluctuations of the donor–accep-
reflecting the ground-state tunneling. Moreover, AL/AH,
tor distance (DAD), and are thus referred to as Marcus-
which are intercept values of the tangents to the plot in
like models. However, various other terms have been
Fig. 3b, are expected to be inflated in comparison to the
coined for these kinds of models, such as “vibrationally
high temperature limit (Fig. 3b, region 3). Between the
enhanced tunneling” model,40 “environmentally coupled
high and low temperature extremes the Arrhenius plot
tunneling” model,41 “protein promoting vibration” mod-
of the KIEs will be curved, as the lighter isotope has a
el,42 and others. All these models suggest that the protein
higher probability to tunnel at the higher temperature
dynamics† and H-tunneling enhance the reaction rates.
than the heavier one (Fig. 3b, region 2). As a result the
Even though these different models (e.g., Borgis and
plot will appear very steep and AL/AH, at region 2, is
Hynes,43,44 Kuznetzov and Ulstrup,45 Sutcliffe and Scrut-
expected to be smaller than unity. To summarize, ac-
ton,40 Warshel,46 Caratzoulas and Schwartz,42 Nagel and
cording to this model, isotope effects on the Arrhenius
Klinman47) originate from different basic principles,
prefactors can be taken as an indication for tunneling,
where different degrees of tunneling will lead to either †
By dynamics, we refer to any motion in the system regardless
inflated or deflated intercept values in comparison to the if it is in thermal equilibrium with the environment or not.

Israel Journal of Chemistry 49 2009


167

Table 3. Enzymatic systems with properties outside the limits predicted by the “tunneling correction” models
temperature
enzyme type of transfer dependence kH/kD AH/AD ref
aromatic amine dehydrogenase proton no 55 26
methylamine dehydrogenase proton no 16.8 13.3 27
trimethylamine dehydrogenase proton (suspected) no 4.6 (for pH > 9.5) 7.8 28
sarcosine oxidase proton (suspected) no 7.3 5.8 29
pentaerythritol tetranitrate reductase hydride no ~4.1 4.1 30
ethanolamine ammonia lyase H radical 4.4–4.7 31
glutamate mutase H radical 2.4 32
soybean lipooxygenase H radical yes 81 18 33
soybean lipooxygenase, mutants H radical yes 93–112 4–0.12 33
dihydrofolate reductase (DHFR) hydride no 3.5 4.0 34
G121V-M42W DHFR hydride yes 3.5 0.04 35
primitive DHFR hydride yes 3.5 36
thymidylate synthase hydride no 3.7 37
thymidylate synthase proton yes, under 20 ºC 2.2 38
alcohol dehydrogenase (ADH) hydride yes, under 30 ºC 5.6 (kH/kT) 4.3 (>30 ºC) (AH/AT) 39

they all address a similar phenomenological model,


and successfully differentiate between the temperature
dependence on the rate and temperature dependence of
the KIEs.5,6 Even though these models often use dif-
ferent terminology, they all (i) treat hydrogen quantum
mechanically throughout the reaction coordinate and (ii)
consider the reaction coordinate to mostly represent the
dynamics of the environment surrounding the hydrogen
that is being transferred.48 Moreover, they all consider
two requirements for efficient tunneling: degeneracy of
the reactant and product energy levels and narrow bar-
rier width.5,6 A general description for these models can
be expressed by the following equation: Fig. 4. Illustration of the Marcus-like model as expressed
r1 in eq 5, where the two orthogonal coordinates represent:
“Marcus term”, the environmental energy parabolas for the
k = Ce -]DGc + mg
2/]4mRT g
# e F]mg e - E F]mg
/k b T
dDAD (5)
reactant (R) and product (P) states, and the “gating term”,
fluctuation of the DAD along the hydrogen potential surface.
r0

An illustration of a “Marcus-like” model is presented The Franck–Condon nuclear overlap is presented as a red cur-
in Fig. 4. vature indicating the probability of finding the particle at the
In this model, the rate of H transfer is determined reactant and product state. Figure was modified from ref 36.
by the isotope-independent preexponential term (C = Color visible online.
constant), a Marcus term (first exponential), a Franck–
Condon term (F(m)), and a “gating term” (EF(m)) that cur before significant tunneling can take place (i.e.,
describes the temperature and mass dependence of the the reorganization of the donor and acceptor toward a
DAD fluctuations. The last two terms are integrated “tunneling ready” conformation).47,49 The Franck–Con-
for all the conformations sampled by the DAD. The don nuclear overlap along the hydrogen coordinate,
Marcus term is isotopically insensitive, and represents which is the integrated hydrogen tunneling probability
the heavy-atom reorganization‡ that is required to oc- (Fig. 1), is an isotopically sensitive but temperature-in-
dependent term. It depends on the DAD, which is both
isotopically and temperature dependent. In eq 5, the

Reorganization refers to heavy atom motions that bring the
system from a ground state of the chemical step to a tunneling last exponential term describes the relative fluctuations
ready conformation. The term “preorganization” is sometimes of DAD, and is graphically illustrated in Fig. 5 as the
used to describe the protein motions, of both distal and “gating term” coordinate. According to this model, the
proximal residues, that bring the system to the ground state of temperature-independent KIEs indicate a system with
the H transfer reaction. a perfectly reorganized reaction coordinate, where the

Stojković and Kohen / Enzymatic H Transfers


168

of the preferred systems for structure–dynamics–func-


tion studies due to its small size, flexibility, and the
simple chemical transformation it catalyzes. Moreover,
due to its pivotal role in nucleotide biosynthesis in many
organisms, DHFR is the target for many antibiotic and
chemotherapeutic agents51 and an excellent platform
for genetic and evolutionary analysis.52 Consequently,
DHFR has been studied by different experimental meth-
ods in order to gain understanding of its structure,53 dy-
namics,54 and kinetics.55
The coupling of the enzyme’s structure and dynamics
to the chemical step has been studied in wild type (wt)
DHFR as well as for the three remote mutants (G121V,
M42W, and their double mutant G121V-M42W).34,35,56,57
Fig. 5. Comparison of the Arrhenius plots of intrinsic H/T These residues are about 17 Å from the active site and
KIEs of wt (red), G121V (green), M42W (blue), and G121V- 25 Å from each other, yet, QM/MM simulation sug-
M42W (black). Figure was reproduced with permission from gested that they are dynamically coupled to each other
ref 35. Color visible online. and to the reaction coordinate.58 The intrinsic 1º H/D
KIE and H/T KIE have been obtained for these enzymes
average DAD is ideal for tunneling so thermal activa- over a temperature range of 5–45 ºC (Fig. 5). Compara-
tion of the fluctuations of the DAD does not change tive isotope effects on Arrhenius preexponential factors
the Franck–Condon integral. In other cases, where the together with ∆Ea are presented in Table 4. The tem-
reorganization is not perfect, sampling of DAD results perature-independent KIEs, with large AH/AD, and non-
in temperature-dependent KIEs. zero Ea values for the wtDHFR (Fig. 5, red line, color
Several enzymatic systems (dihydrofolate reductase, visible online) have been rationalized in the context of
thymidylate synthase, and soybean lipoxygenase) and the Marcus-like models with ideal rearrangement of the
their mutants are examples of proton, hydrogen, and hy- potential surface.34 The average DAD in this system ap-
dride transfer systems, where Marcus-like models were pears to be ideal for tunneling.
implemented in order to explain experimental data. Two single mutants (G121V and M42W) showed
slightly inflated KIEs with weak temperature depen-
dences (Fig. 5, green and blue lines, respectively. Color
Experimental examples using intrinsic
visible online.) These observations suggest that both
1º KIE, isotope effects on Arrhenius
mutants have less perfectly reorganized reaction coordi-
exponential factors, and temperature
nates in comparison to wtDHFR. Therefore, weak tem-
dependence of KIEs
perature-dependent KIEs suggest that single mutants
Dihydrofolate Reductase required some thermally activated fluctuations in order
Dihydrofolate reductase (DHFR) catalyzes the re- to reduce the DAD and to promote tunneling. In con-
duction of 7,8-dihydrofolate (DHF) to 5,6,7,8-tetrahy- trast to both single mutants, the double mutant (G121V-
drofolate (THF), with the stereospecific transfer of a M42W) had a very steep temperature dependence of the
hydride from the pro-R C4 position of NADPH to the si KIEs (Fig. 5, black line, color visible online), implying
face of C6 of the pterin ring50 (Scheme 1). DHFR is one poor rearrangement, and an average DAD that is too
long to enable efficient tunneling. Consequently, essen-
tial thermally activated fluctuations of the DAD in the
double mutant lead to a large temperature dependence.
These findings indicated that the remote single mutants
mostly affected the prearrangement of the system, re-
sulting in the observed slower reaction (Table 4, first
row). However, each single mutation had a small effect
on the rearrangement of the system (weak temperature
dependence of KIEs). The double mutation, on the other
Scheme 1. Mechanism for the hydride transfer in DHFR, hand, resulted in a significant change in the nature of
where R = adenine dinucleotide 2¢phosphate and R¢ = (p-ami- the hydride transfer leading to a substantial temperature
nobenozoyl) glutamate. dependence of KIEs, which indicates a poor reorganiza-

Israel Journal of Chemistry 49 2009


169

Table 4. Comparative KIEs on Arrhenius preexponential factors for wtDHFR and its mutants.
Reproduced using data published in ref 35
parameters wt G121V M42W G121V-M42W
kH* (s )
–1
228 ± 8 1.4 ± 0.2 5.6 ± 0.4 0.03 ± 0.005
AH/AT 7.0 ± 1.5 7.4 ± 1.6 2.8 ± 0.2 0.1 ± 0.1
AH/AD 3.5 ± 0.5 4.7 ± 1.5 2.1 ± 0.2 0.04 ± 0.03
AD/AT 1.70 ± 0.14 1.70 ± 0.07 1.35 ± 0.05 0.25 ± 0.09
∆Ea(T–H), kcal/mol –0.1 ± 0.2 0.23 ± 0.03 0.58 ± 0.04 3.6 ± 0.3
*Pre-steady-state rates of H transfer at 25 ºC and pH 7.

tion of the reaction coordinate. The synergistic effect of of evolution in optimizing enzyme dynamics to enhance
these two remote residues on the active site chemistry tunneling. The comparative data indicated that unlike
supports the idea of the long-range dynamically coupled cDHFR, for which intrinsic KIEs are nearly temperature
network previously suggested by hybrid quantum/clas- independent (∆Ea = –0.1 ± 0.3 kcal mol–1)34, the R67-
sical molecular dynamics simulations.58 DHFR has temperature-dependent KIEs (∆Ea = 0.87 ±
Another aspect of DHFR catalysis that was investi- 0.03 kcal mol–1). Moreover, AH/AT for cDHFR is signifi-
gated was the possible role of evolution in tuning en- cantly larger than the semiclassical value (AH/AT = 7.4 ±
zyme dynamics to enhance tunneling.36 The nature of 4.0)34, while the isotope effect on AH/AT for R67-DHFR
the H transfer in primitive R67 DHFR was compared to is much smaller and falls in the semiclassical limit (AH/
that of a mature, highly evolved enzyme, chromosomal AT = 1.36 ± 0.07 kcal mol–1).
DHFR (cDHFR). The R67 DHFR is genetically unrelat- Interpretation of these results depends on the model
ed to the cDHFR; it is an R-plasmid-encoded DHFR and used. According to TST-models with tunneling correc-
is present in some bacterial strains59 that are resistant to tion, the above data would suggest extensive tunneling
the antibiotic drug trimethoprim (TMP—a cDHFR pi- in cDHFR, and no tunneling in the primitive DHFR.
comolar inhibitor).59 Even though it catalyzes the same Alternatively, Marcus-like models would suggest that
reaction as the cDHFR, they share neither sequence nor the reaction coordinate in the highly developed enzyme
structural similarities (Fig. 6). Furthermore, these two is perfectly reorganized (temperature-independent
enzymes exhibit different degrees of flexibility. The ma- KIEs), and that the primitive enzyme exhibits a poor
ture cDHFR is flexible and exhibits motions on different reorganization of the reaction coordinate (temperature-
timescales that are crucial for its proper function.60,61 On dependent KIEs). These findings are in accordance with
the other hand, R67-DHFR shows rigidity in its motions the hypothesis that enzymes evolved to optimize their
on all the timescales examined.62 Therefore, these two reaction coordinate for efficient tunneling.
systems are a good platform for examination of the role
Thymidylate Synthase (TSase)
Thymidylate synthase catalyzes the reductive meth-
ylation of 2¢-deoxyuridine-5¢-monophosphate (dUMP)
to 2¢-deoxythymidine-5¢-monophosphate (dTMP)
(Scheme 2). The cofactor N5,N10-methylene-5,6,7,8-
tetrahydrofolate (MTHF) serves both as a methy-
lene and hydride donor, yielding 7,8-dihydrofolate
(DHF).63 The hydride transfer step (Step 5 in Scheme
2) was studied using 1º H/T and D/T KIEs, and tem-
perature-independent intrinsic H/T KIEs of 7 were
obtained.37 For the wtTSase, the isotope effects on Ar-
rhenius preexponential factors were much higher than
the semiclassical limit (AH/AT= 6.8 ± 2.8 and AD/AT =
Fig. 6. Top: Structures of cDHFR (left) and R67-DHFR
1.9 ± 0.25), and the reaction exhibited nonzero energy
(right). Bottom: the reverse images for each active site em- of activation (Ea = 4.0 ± 0.01 kcal/mol). Temperature-
phasizing the different orientation of the donor (magenta) and independent KIEs with a significant energy of activation
acceptor (green).59 Figure was reproduced with permission suggest a perfectly reorganized potential surface for the
from Wong et al., PNAS 2006, National Academy of Sciences, hydride transfer, which is in accordance with the envi-
U.S.A. Color visible online. ronmentally coupled tunneling model.

Stojković and Kohen / Enzymatic H Transfers


170

Scheme 2. The chemical mechanism for the TSase catalyzed reaction. The transferred methylene group is in purple, the nu-
cleophilic Cys in yellow, and the hydride in green. R = 2¢-deoxyribose-5¢-phosphate. R¢ = (p-aminobenzoyl)glutamate. Color
visible online.
The method established for the wtTSase was then
used to resolve a controversy regarding the mechanism
of TSase. Until recently, two distinct mechanisms were
proposed for step 5 (Scheme 2). Schultz and cowork-
ers64 studied unnatural amino acid mutations at the
enzyme’s active site, and suggested a two-step radical
mechanism where an electron is first transferred from
THF to form a tryptophan (W80)-stabilized cation radi-
cal, and then an H-radical is transferred from this cation
radical to form product dTMP (Scheme 3, bottom path).
Stroud and coworkers,65,66 on the other hand, studied
crystal structure of wtTSase and several W80 mutants,
and suggested a one-step hydride transfer mechanism
(Scheme 3, top path), which implied that W80 could
not stabilize the H4folate cation radical. Examination
of these mechanisms was performed by comparing the
nature of the H transfer between wtTSase and W80M-
TSase mutant, which cannot stabilize a cation radical.67
It was found that the nature of the H transfer and its
coupling to the environment did not change upon muta- Scheme 3. Two proposed mechanisms for the H transfer from
tion, and that the three-orders-of-magnitude rate reduc- C6 of THF to the exocyclic methylene. (Top) A one-step
tion of this mutant was a result of a smaller percentage hydride transfer mechanism. (Bottom) The two-step radical
of reactive conformations. These findings supported mechanism. Full arrows symbolize transfer of a pair of elec-
the hydride transfer mechanism67 and demonstrated the trons, and half arrows symbolize transfer of a single electron.
ability of these methods to examine the physical nature Figure was reproduced with permission from ref 67.
of H transfer, its coupling to dynamics, and the role of
active site residues in catalysis.
coupled electron transfer is greater than 80 at 298 K.
Soybean Lipoxygenase-1 (SLO) The reaction has a fairly small activation energy (Ea =
Soybean lipooxygenase-1 is a non-heme Fe-depen- 2.1 kcal/mol), and the isotope effect on Arrhenius pref-
dent enzyme that catalyzes the conversion of linoleic actors is close to 20 (Table 2).68–70 An Eyring analysis of
acid to 13(S)-hydroperoxy-9(Z),11(E)-octadecanoic the kinetic data for the wtSLO implies that the reaction
acid (Scheme 4). It exhibits one of the largest ever barrier is largely entropic (∆H‡ = 1.5 kcal/mol, –T∆S‡ =
measured KIEs, where the H/D KIE for the proton 12.8 kcal/mol).70 Significant KIEs indicate a very large

Israel Journal of Chemistry 49 2009


171

Scheme 4. Consensus mechanism for SLO-1, in which active site Fe(III)-OH abstracts a proton and electron from the C-11 po-
sition of substrate. The structures 2a, 2b, and 2c indicate the three resonance forms, in which the single electron density resides
at positions 11, 13, and 9, respectively (where R = C5H11, and R¢ = C7H14COOH). Figure was reproduced with permission from
Meyer et al., PNAS 2008, National Academy of Sciences, U.S.A.

inherent chemical barrier, suggesting that the transfer of ics of the protein contribute to the enhancement of the
H-radical occurs mainly through tunneling. However, catalyzed chemical transformation. One of the working
semiclassical theory cannot account for such an enor- hypotheses in this field is that tunneling in enzymatic
mous KIE with inflated AH/AD. reactions occurs from a better pre-/re-arranged con-
Interestingly, a distal mutant, I553A (I553 is one he- formation along the reaction coordinate than a similar
lix turn away from the active site residue L546), has a reaction in solution or in primitive enzymes. Therefore,
1º /D KIE and Ea parameters similar to wtSLO, however H-tunneling is used as a probe for the organization of
its AH/AD ratio becomes inverted and its KIEs become the system, and to investigate the dynamic contribution
temperature dependent.70 These parameters (kH/kD, Ea, to catalysis.
and AH/AD) could only be explained with the Marcus- In this review, we presented several examples of the
like models suggesting that wtSLO has an active site way the nature of H-tunneling has been examined. Data
structure that is well reorganized to facilitate hydrogen obtained through examination of temperature depen-
tunneling, and that the remote I553A mutation perturbs dence of KIEs for enzymatic systems such as DHFR,
the structural and dynamic elements that support that TSase, and SLO could not be rationalized using the
process. The mutation seems to increase the average semiclassical models with or without tunneling correc-
DAD. Consequently, the mutant requires thermally ac- tion. Better analysis and rationale was achieved with
tivated motion in order to facilitate tunneling leading to Marcus-like models. With the development of these new
temperature-dependent KIEs. Klinman and coworkers theoretical models we are at a point where we can ex-
further supported this conclusion while studying a series amine questions like whether or not enzymes evolved to
of I553 mutants.71 optimize their reaction coordinate for tunneling, as well
as other questions that could have a significant impact
on biomimetic catalysis design.
Summary Acknowledgments. This work was supported by NIH R01
The focus of this field has drastically changed in the past GM65368 and NSF CHE- 0715448.
ten years. The most frequent question in the late 1990s
concerned the prevalence of H-tunneling in C–H acti-
vation reactions, and if it contributes more to enzyme References and notes
(1) Melander, L.; Saunders, W.H. Reaction Rates of Iso-
catalyzed H transfers relative to uncatalyzed ones. In
topic Molecules, 4th ed.; Krieger, R.E.: Malabar, FL,
recent years it was recognized that H-tunneling contrib- 1987.
utes to both catalyzed and uncatalyzed reactions. The (2) Kraut, J. Science 1988, 242, 533–540.
current assessment is that H-tunneling is always part (3) Marcus, R.A.; Sutin, N. Biochem. Biophys. Acta 1985,
of the H transfer reactions, whether reactions occur in 811, 265–322.
the gas phase or the solid phase, solutions or enzymes. (4) Marcus, R.A. J. Phys. Chem. B. 2007, 111, 6643–6654.
Today most investigators are examining if the dynam- (5) Kohen, A. In Isotope Effects in Chemistry and Biology;

Stojković and Kohen / Enzymatic H Transfers


172

Kohen, A.; Limbach, H.H., Eds.; Taylor & Francis, (31) Bandarian, V.; Reed, G.H. Biochemistry 2000, 39,
CRC Press: Boca Raton, FL, 2006; Ch. 28, pp 743– 12069–12075.
764. (32) Cheng, M.C.; Marsh, E.N. Biochemistry 2005, 44,
(6) Kohen, A. In Hydrogen Transfer Reactions; Hynes, 2868–2691.
J.T.; Limbach, H.H.; Klinman, J.P.; Schowen, R.L., (33) Knapp, M.J.; Klinman, J.P. J. Am. Chem. Soc. 2002,
Eds.; Wiley-VCH: Weinheim, 2006; Vol. 4, Ch. II-12, 124, 3865–3874.
pp 1311–1340. (34) Sikorski, R.S.; Wang, L.; Markham, K.A.; Rajagopalan,
(7) Steinfeld, J.I.; Francisco, J.S.; Hase, W.L. Chemical P.T.R.; Benkovic, S.J.; Kohen, A. J. Am. Chem. Soc.
Kinetics and Dynamics, 2nd ed.; Prentice Hall: Upper 2004, 126, 4778–4779.
Saddle River, NJ, 1998. (35) Wang, L.; Goodey, N.M.; Benkovic, S.J.; Kohen, A. Proc.
(8) Bigeleisen, J.; Mayer, M.G. J. Chem. Phys. 1947, 15, Natl. Acad. Sci. U.S.A. 2006, 103, 15753–15758.
261–267. (36) Yahashiri, A.; Howell, E.E.; Kohen, A. ChemPhysChem
(9) Bigeleisen, J. J. Chem. Phys. 1955, 23, 2264–2267. 2008, 9, 980–982.
(10) Bell, R.P. The Tunnel Effect in Chemistry; Chapman & (37) Agrawal, N.; Hong, B.; Mihai, C.; Kohen, A. Biochemistry
Hall: London & New York, 1980. 2004, 43, 1998–2006.
(11) Cleland, W.W. Methods Enzymol. 1980, 64, 104–125. (38) Hong, B.; Maley, F.; Kohen, A. Biochemistry 2007, 46,
(12) Huskey, W.P.; Schowen, R.L. J. Am. Chem. Soc. 1983, 14188–14197.
105, 5704–5706. (39) Kohen, A.; Cannio, R.; Bartolucci, S.; Klinman, J.P.
(13) Kohen, A.; Jensen, J.H. J. Am. Chem. Soc. 2002, 124, Nature 1999, 399, 496–499.
3858–3864. (40) Sutcliffe, M.J.; Scrutton, N.S. Eur. J. Biochem. 2002,
(14) Cha, Y.; Murray, C.J.; Klinman, J.P. Science 1989, 243, 269, 3096–3102.
1325–1330. (41) Francisco, W.A.; Knapp, M.J.; Blackburn, N.J.; Klin-
(15) Kohen, A. Prog. React. Kinet. Mech. 2003, 28, 119– man, J.P. J. Am. Chem. Soc. 2002, 124, 8194–8195.
156. (42) Caratzoulas, S.; Schwartz, S.C. J. Chem. Phys. 2001,
(16) Klinman, J.P. Trends Biochem. Sci. 1989, 14, 368. 114, 2910–2918.
(17) Kohen, A.; Klinman, J.P. Acc. Chem. Res. 1998, 31, (43) Borgis, D.; Hynes, J.T. J. Chem. Phys. 1991, 94, 3619–
397–404. 3628.
(18) Kohen, A.; Klinman, J.P. Chem. Biol. 1999, 6, R191– (44) Borgis, D.; Hynes, J.T. Chem. Phys. 1993, 170, 315–
198. 346.
(19) Cleland, W.W. In Enzyme Mechanism from Isotope (45) Kuznetsov, A.M.; Ulstrup, J. Can. J. Chem. 1999, 77,
Effects; Cook, P.F., Ed.; CRC Press: Boca Raton, FL, 1085–1096.
1991, pp 247–268. (46) Warshel, A. Proc. Natl. Acad. Sci. USA 1984, 81, 444–
(20) Northrop, D.B. In Enzyme Mechanism from Isotope Ef- 448.
fects; Cook, P.F., Ed.; CRC Press: Boca Raton, FL, 1991, (47) Nagel, Z.D.; Klinman, J.P. Chem. Rev. 2006, 106,
pp 181–202. 3095–3118.
(21) Northrop, D.B. Annu. Rev. Biochem. 1981, 50, 103–131. (48) Kiefer, P.M.; Hynes, J.T. In Isotope Effects in Chemistry
(22) Cleland, W.W. In Isotope Effects in Chemistry and and Biology; Kohen, A.; Limbach, H.H., Eds.; Taylor &
Biology; Kohen, A.; Limbach, H.H., Eds.; Taylor & Francis, CRC Press: Boca Raton, FL, 2006; Ch. 21, pp
Francis, CRC Press: Boca Raton, FL, 2006; Ch. 37, pp 549–578.
915–930. (49) Marcus, R.A. J. Phys. Chem. B 2007, 111, 6643–6654.
(23) Kim, T.; Kreevoy, M.M. J. Am. Chem. Soc. 1992, 114, (50) Miller, G.P.; Benkovic, S.J. Chem. Biol. 1998, 5, R105–
7116. R113.
(24) Jonsson, T.; Glickman, M.H.; Sun, S.J.; Klinman, J.P. J. (51) Warlick, C.A.; Sweeney, C.L.; McIvor, R.S. Biochem.
Am. Chem. Soc. 1996, 118, 10319–10320. Pharmacol. 2000, 59, 141–151.
(25) Moiseyev, N.; Rucker, J.; Glickman, M.H. J. Am. Chem. (52) Lee, J.; Natarajan, M.; Nashine, V.C.; Socolich, M.; Vo,
Soc. 1997, 119, 3853–3860. T.; Russ, W.P.; Benkovic, S.J.; Ranganathan, R. Science
(26) Masgrau, L.; Roujeinikova, A.; Johannissen, L.O.; Hothi, 2008, 322, 438–442.
P.; Basran, J.; Ranaghan, K.E.; Mulholland, A.J.; Sut- (53) Sawaya, M.R.; Kraut, J. Biochemistry 1997, 36, 586–603.
cliffe, M.J.; Scrutton, N.S.; Leys, D. Science 2006, 312, (54) Boehr D.D.; M.D.; Dyson H.J.; Wright, P.E. Science
237–241. 2006, 313, 1638–1642.
(27) Basran, J.; Sutcliffe, M.J.; Scrutton, N.S. Biochemistry (55) Fierke, C.A.; Johnson, K.A.; Benkovic, S.J. Biochemistry
1999, 38, 3218–3222. 1987, 26, 4085–4092.
(28) Basran, J.; Sutcliffe, M.J.; Scrutton, N.S. J. Biol. Chem. (56) Wang, L.; Goodey, N.M.; Benkovic, S.J.; Kohen, A. Phi-
2001, 276, 24581–24587. los. Trans. R. Soc. London B 2006, 361, 1307–1315.
(29) Harris, R.J.; Meskys, R.; Sutcliffe, M.J.; Scrutton, N.S. (57) Wang, L.; Tharp, S.; Selzer, T.; Benkovic, S.J.; Kohen, A.
Biochemistry 2000, 39, 1189–1198. Biochemistry 2006, 45, 1383–1392.
(30) Basran, J.; Harris, R.J.; Sutcliffe, M.J.; Scrutton, N.S. J. (58) Wong, K.F.; Selzer, T.; Benkovic, S.J.; Hammes-Schiffer,
Biol. Chem. 2003, 278, 43973–43982. S. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 6807–6812.

Israel Journal of Chemistry 49 2009


173

(59) Howell, E.E. ChemBioChem 2005, 6, 590–600. (66) Finer-Moore, J.S.; Santi, D.V.; Stroud, R.M. Biochem-
(60) Sawaya, M.R.; Kraut, J. Biochemistry 1997, 36, 586– istry 2003, 42, 248–256.
603. (67) Hong, B.; Haddad, M.; Maley, F.; Jensen, J.H.; Kohen,
(61) Boehr, D.D.; McElheny, D.; Dyson, H.J.; Wright, P.E. A. J. Am. Chem. Soc. 2006, 128, 5636–5637.
Science 2006, 313, 1638–1642. (68) Jonsson, T.; Glickman, M.H.; Sun, S.; Klinman, J.P. J.
(62) Pitcher, W.H.R.; DeRose, E.F.; Mueller, G.A.; Howell, Am. Chem. Soc. 1996, 118, 10319–10320.
E.E.; London, R.E. Biochemistry 2003, 42, 11150–11160.
(69) Rickert, K.W.; Klinman, J.P. Biochemistry 1999, 38,
(63) Carreras, C.W.; Santi, D.V. Annu. Rev. Biochem. 1995,
12218–12228.
64, 721–762.
(64) Barrett, J.E.; Lucero, C.M.; Schultz, P.G. J. Am. Chem. (70) Knapp, M.J., Rickert, K.; Klinman, J.P. J. Am. Chem.
Soc. 1999, 121, 7965–7966. Soc. 2002, 124, 3865–3874.
(65) Stroud, R.M.; Finer-Moore, J.S. Biochemistry 2003, 42, (71) Meyer, M.P.; Tomchick, D.R.; Klinman, J.P. Proc.
239–247. Natl. Acad. Sci., USA 2008, 105, 1146–1151.

Stojković and Kohen / Enzymatic H Transfers

Das könnte Ihnen auch gefallen