Atomic and Molecular details
PROTEINS: Structure, Function, and Genetics 43:420 – 432 (2001)
Mechanism of NAD(P)H:Quinone Reductase: Ab Initio
Studies of Reduced Flavin
German Cavelier1,2 and L. Mario Amzel1*
1
Department of Biophysics and Biophysical Chemistry, Johns Hopkins University School of Medicine, Baltimore, Maryland
2
Program in Molecular Biophysics, Johns Hopkins University, Baltimore, Maryland
ABSTRACT
NAD(P)H:quinone oxidoreductase
type 1 (QR1, NQO1, formerly DT-diaphorase; EC
1.6.99.2) is an FAD-containing enzyme that catalyzes
the nicotinamide nucleotide-dependent reduction of
quinones, quinoneimines, azo dyes, and nitro groups.
Animal cells are protected by QR1 from the toxic and
neoplastic effects of quinones and other electrophiles.
Alternatively, in tumor cells QR can activate a number of cancer chemotherapeutic agents such as mitomycins and aziridylbenzoquinones. Thus, the same
enzyme that protects the organism from the deleterious effects of quinones can activate cytotoxic chemotherapeutic prodrugs and cause cancer cell death.
The catalytic mechanism of QR includes an important
initial step in which FAD is reduced by NAD(P)H. The
unfavorable charge separation that results must be
stabilized by the protein. The details of this charge
stabilization step are inaccessible to easy experimental verification but can be studied by quantum chemistry methods. Here we report ab initio quantum mechanical calculations in and around the active site of
the enzyme that provide information about the fine
details of the contribution of the protein to the stabilization of the reduced flavin. The results show that (1)
protein interactions provide approximately 2 kcal/
mol to stabilize the planar conformation of the reduced flavin isoalloxazine ring observed in the X-ray
structure; (2) the charge separation present in the
reduced planar form of the flavin is stabilized by
interactions with groups of the protein; (3) even after
stabilization, the reduction potential of the cofactor
remains more negative than that of the free flavin,
making it a better reductant for a larger variety of
quinones; and (4) the more negative reduction potential may also result in faster kinetics for the quinone
reduction step. Proteins 2001;43:420 – 432.
effects of quinones, whose redox cycling is avoided by the
obligatory two-electron reduction carried out by QR1.
Surprisingly, the same reaction that protects against the
toxic effects of quinones can be lethal for tumor cells by
activating cytotoxic quinone anticancer drugs such as
mitomycins and aziridylbenzoquinones. Because levels of
QR1 are elevated in some tumors, the enzyme provides an
opportunity to develop improved chemotherapeutic
agents.2–7
QR1 is a homodimer (273 residues per monomer) with
two identical catalytic sites at two equivalent locations on
the dimer interface (Fig. 1). Each binding site, formed by
residues of both monomers, contains an FAD prosthetic
group that remains noncovalently bound during catalysis.
NAD(P)H and NAD(P)⫹ cycle in and out of the catalytic
site of the enzyme, providing reducing equivalents for the
reaction. The kinetics of the reaction is ping-pong: the
oxidized cofactor NAD(P)⫹ must be released before the
quinone substrate can bind.
The structure of rat QR1 (rQR1), as determined by X-ray
diffraction methods, shows that each monomer is formed
by two domains: an N-terminal catalytic domain and a
C-terminal domain involved in dimerization.5 The major
portion of each monomer, the catalytic domain, has an
overall fold characteristic of other flavoproteins: a twisted
central parallel -sheet surrounded on both sides by
connecting helices.
The structures of rQR1 determined in complexes with
NAD(P)⫹ and with a substrate and an inhibitor have
provided a detailed description of the catalytic site and
important insights into the mechanism of the enzyme.5
Recently, the structures of apo human and mouse QR1
have been reported8 as well as the structure of another
cytosolic quinone reductase, QR2.9 These structures show
that the isoalloxazine moiety of FAD interacts with resi-
© 2001 Wiley-Liss, Inc.
Key words: density functional; enzymes; flavoproteins; DT-diaphorase; isoalloxazine;
quantum chemistry
INTRODUCTION
NAD(P)H:quinone oxidoreductase type 1 (QR1, EC
1.6.99.2; for a review see Ref. 1) is an FAD-containing
protein that catalyzes the nicotinamide nucleotide-dependent reduction of quinones, quinoneimines, azo dyes, and
nitro groups.1–5 It protects against the toxic and neoplastic
©
2001 WILEY-LISS, INC.
Abbreviations: NAD(P)H, nicotinamide adenine dinucleotide (phosphate), reduced; NAD(P)⫹, nicotinamide adenine dinucleotide (phosphate), oxidized; FAD, flavin adenine dinucleotide, also oxidized form;
FADH2, flavin adenine dinucleotide, reduced; SCF, self-consistent
field method.
Grant sponsor: National Institute of General Medical Sciences;
Grant numbers: GM45540, 5T32GM-.
*Correspondence to: L. Mario Amzel, Department of Biophysics and
Biophysical Chemistry, Johns Hopkins University School of Medicine,
725 North Wolfe Street, Baltimore, MD-. E-mail:-Received 22 May 2000; Accepted 1 February 2001
Published online 00 Month 2001
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
Fig. 1. Ribbon diagram of the X-ray structure of rQR1. The active sites
between the two domains of the homodimer show the noncovalently
bound flavin cofactor.
dues from loops of both monomers. Aromatic residues
Tyr-104, Trp-105, and Phe-106 and the main chain of
Leu-103 interact directly with the rings of the isoalloxazine, anchoring it. The two oxygen atoms of the flavin
ring (O2F1 and O4F) form hydrogen bonds with mainchain NH groups of the protein: O4 with Phe-106 and O2
with Gly-150. Nitrogens in the rings also form hydrogen
bonds with NH groups of the protein: N1F with Gly-149
and N5F with Trp-105. The two methyl groups of the
isoalloxazine ring are seated in a pocket formed by two
residues of one monomer (Tyr-104 and Trp-105) and four
residues of the other monomer (Ile-50, Tyr-67, Pro-68, and
the main chain of Glu-117). The ribitol, phosphates, and
adenine ring of FAD interact noncovalently with several
loops and helices, anchoring the FAD cofactor to the
enzyme.5
The mechanism proposed for QR1 has two distinct steps:
(1) hydride transfer from NAD(P)H to the enzyme-bound
flavin followed by the release of NAD(P)⫹ and (2) hydride
transfer from the flavin to the quinone followed by the
release of the hydroquinone. In the first half of the reaction
cycle, the nicotinamide of NAD(P)H is in an ideal position
for a direct hydride transfer to FAD: C4 of the nicotinamide is 4.0 Å from N5 of flavin.5 As mentioned previously,
the N1F is hydrogen-bonded to the NH of Gly-149. If a
negative charge develops on N1F upon FAD reduction,
there are no groups that can donate a proton to compensate the charge, so the most likely tautomer of the FAD is
the enolic (FADH2) form, with part of the negative charge
at O2F. This oxygen is already an acceptor of a hydrogen
1
The usual atom nomenclature for each cofactor is followed with a
final capital letter indicating the cofactor: F, flavin; A, adenosine of
FAD; and N, nicotinamide-ribose.
421
bond from Tyr-155, which can in turn be stabilized by the
positive charge of (or the transfer of a proton from)
His-161. This process is reversed5 when the hydride is
transferred, following the opposite direction, to the quinone that replaced the leaving NAD(P)⫹.
The structures provide a simple rationale for the observed ping-pong kinetics of the reaction; the binding of
the substrate cannot occur until NAD(P)⫹ is released
because the quinone and nicotinamide share the same site.
In this scheme, after NAD(P)⫹ leaves, the quinone arrives
and binds in an ideal orientation to receive the hydride
coming back from FADH2. The quinone is reduced by the
hydride to the singly ionized hydroquinone (hydroquinolate), and the isoalloxazine is oxidized to the quinonoid
form. The proton on O2F is transferred back to the OH of
Tyr-155. The imidazole ring of His-161 becomes fully
protonated again and can either transfer a proton to the
hydroquinolate or simply stabilize its negative charge.
Therefore, in addition to hydride transfer, the net effect of
the second half of the reaction is to transfer a proton from
O2F (accepted in the first half of the reaction) to the
hydroquinolate.5
This mechanism explains the obligatory two-electron
reduction promoted by QR1: both halves of the reaction
involve a hydride transfer (not treated in this article), first
from NAD(P)H to FAD and then from FADH2 to the
quinone. The charge relay formed by Tyr-155 and His-161
allows the reaction to proceed without unfavorable charge
separations.5 The catalytic mechanism of FAD reduction
and subsequent oxidation in its active pocket involve a
fine-tuned succession of steps. Charge separation and
subsequent charge stabilization by the enzyme are at the
core of this mechanism. For a mechanism of this kind,
theoretical quantum chemical studies of the electronic
configuration of the free FAD and the FAD in the binding
environment of the catalytic site can provide direct information about the distribution of the charges during the
reaction. In this article, we report results of quantum
mechanical ab initio calculations aimed at probing the
QR1 catalytic mechanism, particularly the electrostatic
stabilization of the FAD after reduction, and the energetic
contributions of the cofactor and the protein to the completed process. Hydride transfers are not treated in this
article; only the states before and after the reduction of the
FAD are treated.
Reaction mechanisms in enzymes have generally been
studied and rationalized on the basis of structural and
experimental data.4,10 –12 Significant advances have been
made in developing methods that can give a quantum
mechanical description of the electronic structure of entire
macromolecules.13–16 An important goal that has yet to be
accomplished is the detailed description of enzymatic
reaction mechanisms from first principles. Enzymatic
mechanisms involving electron transfer, hydride transfer,
and proton tunneling might be explained in greater detail
on the basis of precise electronic structures. For example,
recent experimental findings17–19 point to the possible
involvement of hydrogen tunneling in enzyme catalysis, a
process that would be better understood with the knowl-
422
G. CAVELIER AND L.M. AMZEL
edge of the electronic structure of reactants, products, and
intermediates calculated quantum chemically in the vicinity of the enzyme active site.
Quantum chemical calculations on coenzymes and other
enzyme ligands, in isolation or interacting with a limited
number of atoms, have provided additional insight into
their function and properties during enzyme catalysis.20 –24 Recently, quantum chemistry has been used to
elucidate the precise events occurring at the active site of
ribonuclease A.25 Calculations based on the extended
Hückel method have been reported for the enzyme DNA
photolyase.26 Combined quantum mechanics/molecular mechanics is a well-established method in which the quantum mechanics is applied only to a restricted part of the
molecule and the rest studied with molecular mechanics.27–29 Other approaches involve model systems to study
with ab initio methods the transition-state structure.30
Other recent methods include the use of density functional
theory (DFT) with continuum dielectric theory31 and the
application of DFT to optimize geometries of the transition
state in a model system.32
Flavin-dependent reductases and oxidases comprise an
important family of enzymes.10,33 Several mechanisms
have been proposed and studied with different experimental and theoretical methods in reactions involving this
cofactor.4,34,35 In addition, quantum mechanical calculations have been used to explore the electronic structure of
the isolated isoalloxazine ring in the reduced and oxidized
states.36 –39 However, no calculations have been reported
for flavins in a protein environment. QR1 is an excellent
system for this type of calculation. First, many of the
mechanistic questions involve understanding the distribution of charges and their stabilization, exactly the kind of
questions that quantum mechanical calculations can answer. Second, the reaction takes place in a cavity within
the protein that is not directly connected to a bulk solvent.
Thus, the calculations do not require the inclusion of
complex solvent models. In general, it can be said that
some redox enzymes, those having properties similar to
QR1, are amenable to studies such as the one presented
here.
MATERIALS AND METHODS
Programs
All quantum mechanical calculations were carried out
with the program GAUSSIAN 9840 (Gaussian Inc.) run on
J90 and SV1 CRAY computers at the Advanced Biomedical Computing Center of the National Cancer Institute
(Frederick, MD). The three-dimensional visualizations
were performed with the graphic capabilities of the software UNICHEM (Oxford Molecular, Inc.) implemented on
SGI workstations. The initial structure of the cofactor was
built on an SGI workstation with the program QUANTA
(Molecular Simulations, Inc.). The method of convergence
used in all SCF calculations was DIIS (Pulay’s Direct
Inversion in the Iterative Subspace extrapolation method)
with a tight (10⫺8) convergence limit. Additionally, all
geometry optimizations used a tight (10⫺8) convergence
limit.
Isolated Isoalloxazine Ring
The isoalloxazine ring was used with a methyl termination at N10F instead of the ribitol-adenine portion of
riboflavin. The initial structure constructed with QUANTA
was optimized with GAUSSIAN 98. Initial optimization
with the MNDO semiempirical method was followed by a
full refinement optimization at the RHF/6-31G level.40 The
optimized structures were used for all further calculations.
Initial exploratory calculations were performed with the
RHF/STO-3G model chemistry. To test the importance of
polarization effects in the calculations, we repeated several RHF calculations with the 6-31G(d) basis set. Calculations with this basis set were also performed with the
hybrid DFT method B3LYP (Becke’s41 style three-parameter DFT with LYP gradient-corrected correlation functional of Lee et al.42) to include some of the effects of
electronic correlation.
Based on the proposed mechanism, calculations were
carried out for the following configurations of the isoalloxazine ring (Fig. 2): (A) oxidized form, (B) reduced form I
(anion, charge on N1F, O2F in keto form), (C) reduced form
II (anion, charge on O2F enolate), (D) resonance form of
the configurations in A and B, (E) reduced form III
(neutral, O2F as OH), and (F) reduced form IV (neutral,
O2F in keto form, NH at N1F). The RHF/STO-3G calculations used 309 –318 basis functions. The central processing
unit (CPU) time employed for each calculation was typically 15–22 h. The B3LYP/6-31G(d) calculations used
around 310 basis functions and CPU times of 6 –7 h.
The numerical values computed comprised the energy of
each configuration, the Mulliken charge of each atom, and
the three-dimensional representations of electrostatic potential, the highest occupied molecular orbital (HOMO)
and the lowest unoccupied molecular orbital (LUMO).
Isoalloxazine Ring in the Protein Environment
We performed equivalent calculations with the isoalloxazine ring anchored in the environment of the catalytic
site of QR1 by including a number of selected residues
around the ring with the coordinates of the QR1 X-ray
structure. The electrostatic potential and charges around
the flavin molecule, as well as the hydrogen bonds the
flavin makes with active site residues, are thought to
have a predominant role in the catalytic mechanism.5,20,23,24,34,43,44 On the basis of the analysis of the
reaction mechanism previously summarized,5 several
amino acids in the environment of the flavin cofactor were
selected for the calculations because (1) they form the
hydrophobic binding site, (2) they make hydrogen bonds
with the isoalloxazine ring, or, more importantly, (3) they
are part of the proposed charge-relay system. On the basis
of these criteria, the following amino acids were selected:
Trp-105, Phe-106, Gly-149, Gly-150, Tyr-155, and His-161.
All individual amino acids and amino acid pairs (Gly-149
& Gly-150, Trp-105 & Phe-106) were terminated at their
carboxy terminus with a methyl group.
In the calculations, the X-ray coordinates of the flavin
were replaced in turn by the coordinates of each of the
optimized isoalloxazine rings described previously, main-
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
423
Fig. 2. Forms of the isoalloxazine ring used in this study: (A) oxidized form before reduction, which
represents the state of the FAD moiety at the beginning of the catalytic cycle immediately after the release of
the reduced quinone; (B) reduced form I, with a negative charge on N1F and with O2F in the keto form, which
depicts the standard way of representing the reduced FAD; (C) reduced form II, with a negative charge initially
located at the O2F atom, which corresponds to the ionized enolate form of the isoalloxazine ring; (D) resonance
form of the configurations in parts B and C (only this configuration exists for this combination of charges and
atoms); (E) reduced form III, in which the charge of the reduced FAD was neutralized by the addition of an H⫹ to
O2F, which is, therefore, protonated in the enol form; and (F) reduced form IV, with O2F in the keto form and the
proton on N1F.
taining the orientation and position of the ring in the
experimental X-ray rQR1 structure. The three structures
shown in Figure 3(A–C) for the isolated rings were used in
this portion of the work, with proper allowance for the
variations induced by the protein. Because of the size of
the analyzed structures, exploratory calculations were
carried out with a minimal basis set with restricted
Hartree–Fock theory (RHF/STO-3G; ⬃150 atoms, ⬃800
basis functions). Final calculations for selected structures
were carried out with the 6-31G(d) basis set with DFT
[B3LYP/6-31G(d)]. These calculations used 1380 basis
functions and took 40 h to 14 days of CPU time.
Effects of Geometry and Dielectric
To analyze the effect of the protein on the geometry of
the reduced isoalloxazine, we performed single-point energy calculations for bent configurations of the cofactor,
isolated and anchored in the catalytic site of QR at several
dihedral angles. The coordinates used in these calculations
were generated first with the bent optimized geometry
found for the reduced form of the free isoalloxazine ring.
The different conformations were generated with the
program QUANTA by changes in the dihedral angle in 5°
intervals around the virtual bond between N5F and N10F.
The values for these intervals were begun at 145° (a more
bent configuration than the equilibrium configuration of
the isolated ring, 151.3°) and increased past the planar
configuration (180°) to 210°. To release the strain in the
isoalloxazine ring created by the changes, we optimized
the coordinates of the substituents on N10F and N5F at
each new dihedral angle. These restrained optimizations
of the rings, isolated and in the protein environment, used
the semiempirical MNDO 94 program with the AM1
Hamiltonian as implemented in UNICHEM. Energies for
424
G. CAVELIER AND L.M. AMZEL
Fig. 3. Initial exploratory electronic structure of the RHF/STO-3G
results. The simplified, compressed representations use the following
color scheme: electrostatic potentials are represented in red (negative)
and blue (positive), LUMOs are represented by solid green surfaces, and
HOMOs are represented by solid gold. Only one phase of the molecular
orbitals is shown for clarity; the lobes with the phase of the opposite sign
are located symmetrically around the corresponding atoms. The Mulliken
charges of selected atoms are indicated as fractions of the electron
charge. The free isoalloxazine ring is shown here in three of the
configurations of Figure 2: (A) oxidized form of FAD before reduction, (B)
initial effect of the reduction by the hydride (note the replacement of
LUMOs by HOMOs and the appearance of an unfavorable charge
separation, very negative around N1F and O2F), and (C) enolic form of
the isoalloxazine ring upon protonation of O2F (note the neutralization of
the previous charge separation and the symmetrical redistribution of
HOMOs and LUMOs).
these optimized structures were calculated with the RHF/
STO-3G model chemistry. This level of theory was considered adequate for these calculations because the resulting
Fig. 4. DFT B3LYP/6-31G(d) electronic structure results. The three
free isoalloxazine rings shown correspond to the configurations of Figure
3(A–C).
energies were used only to estimate differences with
respect to the energy at the minimum. To partially account
for the effects of the solvent, we repeated calculations for
the isolated FAD with the self-consistent-reaction-field
(SCRF) method45 (as implemented in GAUSSIAN46), with
a relative dielectric constant of 80. Another way of partially accounting for the effects of the solvent is to satisfy
the hydrogen-bonding potential of N1F, O2F, N3F, O4F,
and N5F with individual water molecules. To this effect,
water molecules were placed in hydrogen-bonding positions around the isoalloxazine ring of the structure with a
proton in N1F [Fig. 2(F)]. The geometry of this hydrogenbonded isoalloxazine ring was then optimized at the
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
RHF/STO-3G level of theory, and electronic structure
calculations were carried out with the 6-31G(d) basis set as
described and shown later in Figure 7(A–B).
RESULTS AND DISCUSSION
Isolated Isoalloxazine Ring
Quantum mechanical calculations for the chosen configurations of the isoalloxazine ring [Fig. 2(A–F)] were carried
out as described previously. The oxidized form [Fig. 2(A)]
represents the state of the FAD moiety at the beginning of
the catalytic cycle, immediately after the release of the
reduced quinone. Structural optimization shows that in
this form the isoalloxazine ring adopts a planar conformation. A comparison of the calculated optimized structure
with the experimental structure from the Cambridge
Crystallographic Database shows an excellent agreement
(root-mean-square deviation ⫽ 0.08 Å). Reduced form I
[Fig. 2(B)], with a negative charge on N1F and with O2F in
the keto form, depicts the standard way of representing
the reduced FAD. Reduced form II [Fig. 2(C)] corresponds
to the form with the O2F in the enolate form (with a
negative charge). Upon optimization, however, the two
configurations give almost exactly the same coordinates
(root-mean-square deviation ⫽ 0.0009 Å) and atomic Mulliken charges, as expected for true resonance forms. Therefore, only one configuration exists for such a combination
of charges and atoms [Fig. 2(D]. The optimized structure
shows that the isoalloxazine adopts a bent, butterfly-like
conformation with a dihedral angle of 151.3° between the
two halves of the molecule. Two forms were calculated in
which the charge of the reduced FAD was neutralized by
the addition of an H⫹: reduced form III [Fig. 2(E)] with
O2F protonated in the enol form and reduced form IV [Fig.
2(F)] with O2F in the keto form and the proton on N1F.
The dihedral angles of these conformations are 155.3 and
148.4° respectively. This conformation is in excellent agreement with the experimentally determined structure of
lumiflavin47 (dihedral angle ⫽ 144.5°). A previous quantum chemical study39 also found a bent conformation, but
the bend angle (164°) was significantly farther from the
experimental value.
Initial exploratory RHF/STO-3G results of the quantum
mechanical calculations performed after geometry optimization are shown in Figure 3(A–C). The simplified, compressed representation is described in the figure. The
Mulliken charges of selected atoms are indicated as fractions of the electron charge. Only three of the configurations in Figure 2(A–F) are shown [Fig. 3(A–C)]: (1) the
optimized oxidized FAD, corresponding to the form in
Figure 2(A); (2) the optimized reduced FADH⫺, corresponding to the resonance form in Figure 2(D) that arises from
the optimizations of the forms in Figure 2(B,C); and (3) the
optimized reduced FADH2, corresponding to the form in
Figure 2(E). The optimized reduced form IV [Fig. 2(F)] is
not shown, as it was not used in the comparisons done in
the protein environment because it had a higher energy
and a less favorable charge separation than the one in
Figure 3(C) (discussed later).
425
The initial exploratory RHF/STO-3G calculations for the
oxidized isoalloxazine ring [Fig. 3(A)] show a positive
electrostatic potential almost everywhere in the molecule,
except for slightly electronegative regions around the O2F
and O4F oxygens and to a lesser extent around the N1F
and N5F nitrogens, all of which have partial negative
charges. The most negatively charged atoms are N1F
(Mulliken charge ⫽ ⫺0.333) and N3F (charge ⫽ ⫺0.388; it
is compensated by a covalently bound hydrogen with a
charge of ⫹0.220). The oxygens are charged more or less
equally, OO2F with ⫺0.263 and O4F with ⫺0.243. The
electrostatic potential is nonetheless more negative around
the oxygens than around the nitrogens because the oxygens, in addition to being more electronegative, are bound
to only one atom. By contrast, the nitrogens are more
closely surrounded by carbon and hydrogen atoms. The
dimethylbenzene ring has an almost uniform charge distribution when hydrogens are taken into account. The unoccupied electron-acceptor orbital (LUMO) in the hydride
acceptor N5F has high overlap with the LUMO in C4F,
C4aF, and C10aF, which in turn overlaps the LUMO in
N1F, C2F, and N3F [Fig. 3(A)], providing a molecular
orbital path for the delocalization of the electrons after
reduction. Therefore, after reduction of the isolated isoalloxazine ring to the reduced anionic form [Fig. 3(B)], the most
important changes in atomic Mulliken charges occur in
N1F (from ⫺0.333 to ⫺0.423), O2F (from ⫺0.263 to
⫺0.383), and O4F (from ⫺0.243 to ⫺0.392). The rest of the
atoms of the isoalloxazine ring experience smaller changes
in charge. Accordingly, the electrostatic potential becomes
much more electronegative in the vicinity of N1F, O2F,
and O4F [Fig. 3(B)], particularly in the region around O2F.
In the absence of a neutralizing proton, the negatively
charged N1F shares its negative electrostatic potential
with O2F [Fig. 3(B)]. N5F is also more electronegative
than before reduction, but the hydride hydrogen neutralizes its negative potential. The HOMOs in the reduced
isoalloxazine ring [Fig. 3(B)] are in the regions where the
LUMOs were in the oxidized ring [Fig. 3(A)]: the most
significant HOMOs occur on N5F, C4F, C4aF, and C10aF;
on N1F, C2F and N3F; and on O2F and O4F. The most
important difference, however, is that after reduction the
planar conformation of the oxidized isoalloxazine ring [Fig.
3(A)] changes to a bent conformation (dihedral angle ⫽
151.3°) when it becomes reduced [because of the orientation of the drawing, the bent structure is not very apparent
in Fig. 3(B)].
In the mechanism proposed for QR1, the reduced isoalloxazine ring is in the enolic form protonated at O2F. For this
protonated form, the isolated isoalloxazine ring has the
charges redistributed in a symmetrical and nonlocalized
manner [Fig. 3(C)]. The negative electrostatic potential
produced by charge separation at N1F and O2F in the
unprotonated form is almost absent in the uncharged
form. Although the two electrons of the reduction are still
there, their distribution is less polarized. The overall effect
of the protonation of O2F is, therefore, a smoothing of the
charge separation and a decrease in the charge asymmetry
caused by the reduction. Protonation also has a major
426
G. CAVELIER AND L.M. AMZEL
effect on the molecular orbitals: the HOMOs are redistributed toward the central ring in a symmetrical fashion,
whereas the LUMOs are redistributed more evenly along
the whole isoalloxazine ring.
Finally, another possible structure is one protonated at
N1F [Fig. 2(F)]. The effect of this protonation (not shown in
Figs. 3– 4) is similar to that of protonation at O2F, but the
potential at O2F and in O4F is more negative, and the
charges remain somewhat separated and more polarized.
That is, the charge imbalance is not as uniformly distributed as for the protonation of O2F. There is also some
strain in the molecule because the methyl in the N10F
bends toward the back of the isoalloxazine ring. For these
reasons and because N1F cannot be protonated when
bound to rQR1, this form was not used in the comparisons
with the ring in the protein environment.
Calculations with DFT methods and a polarized basis
set that includes d functions for the heavy atoms [B3LYP/
6-31G(d); Fig. 4(A–C), which parallels Fig. 3(A–C)] give
highly similar results. The main difference of including d
functions is evidenced as expected in the Mulliken charges
that are all larger in the calculation with B3LYP/6-31G(d).
Although this is a more energetically accurate calculation,
the Mulliken charges calculated this way are probably
unrealistic because the higher delocalization afforded by
the use of a polarized basis set makes it less appropriate to
use the resulting electron density to estimate point charges.
Isoalloxazine Ring in the Protein Environment
The structure of the isoalloxazine ring refined as part of
the crystal structure of QR1 has slight differences in bond
lengths and bond angles compared with the structure of
the optimized free flavins calculated previously [Fig. 2(A–
F)]. Small errors in the experimental coordinates of the
QR1 structure, determined at a 2.1-Å resolution, can
account for these differences. Therefore, for these calculations hybrid coordinate sets were used: the coordinates of
the protein atoms were directly those of the X-ray structure, and the coordinates of the flavin ring were replaced
in turn by those of each of the optimized isoalloxazine
rings described previously. The orientation and position of
the ring were adjusted to be those in the X-ray structure.
From energetic analysis (shown later in Fig. 8) and
restrained optimizations of the reduced ring in the protein
environment, it is clear that the reduced ring inside the
protein is more planar than the free reduced ring, adopting
a dihedral of around 171–175° in the protein environment
instead of the dihedral of 151.3° found for the free ring.
The planar form (dihedral angle ⫽ 171–175°) was then
used for the final calculations for the reduced isoalloxazine
ring.
Calculations carried out with the optimized structures
of the isoalloxazine ring anchored in the protein environment gave the results shown in Figure 5(A–C). They
correspond to the isolated ring structures in Figure 3(A–
C). Exploratory RHF/STO-3G calculations show that
charges in the isoalloxazine ring have tendencies similar
to those found in the isolated ring, but the influence of the
protein environment on the values of the charges and the
electrostatic potential is significant. Placing the oxidized
isoalloxazine ring in the protein environment results in
small changes in the Mulliken charges [Fig. 5(A)]. The
general trend is toward a higher polarization of the
charges [compared with the isolated ring in Fig. 3(A)]: the
negative charges of the oxygens and nitrogens increase,
and the positive charges of the carbon atoms adjacent to
them also increase. This effect helps to anchor the isoalloxazine ring in its place by intensifying the strength of the
corresponding hydrogen bonds. The electrostatic potential
at and around the atoms of the oxidized isoalloxazine ring
is only slightly negative with respect to the surrounding
protein. The LUMOs are very similar to those observed in
the free isoalloxazine. There is an unoccupied overlapping
path from the LUMO on N5F to the LUMO on C4F, C4aF,
and C10aF. This second LUMO in turn overlaps the
LUMO on N1F, which approaches that of O2F. Therefore,
the same unoccupied path is present for the arriving
reducing electrons as for the isolated oxidized isoalloxazine ring, from N5F through C4aF and C10aF to N1F and
from there to O2F. The influence of the surrounding amino
acids on the flavin leaves the oxidized isoalloxazine ring in
the protein environment mainly with acceptor (LUMO)
orbitals, an optimal situation for receiving the reducing
electrons through hydride transfer.
The configuration of the anionic form of the reduced
isoalloxazine ring in the protein environment is shown in
Figure 5(B). The geometry of the free, isolated reduced
isoalloxazine ring is the bent conformation described previously (dihedral angle ⫽ 151.3°). However, all X-ray structures of rQR1, as well as most flavoproteins, show a rather
planar isoalloxazine ring in the reduced form.48 This
conformation must be a consequence of the protein environment, which imposes a planar structure on the reduced
isoalloxazine ring. Quantum mechanical calculations confirm this explanation. A restrained optimization (with
MNDO 94, AM1 Hamiltonian; see the Materials and
Methods section) of the dihedral of this form of the reduced
isoalloxazine ring in the protein environment produced a
configuration with a dihedral of 171.8°. In addition, an
analysis of the bend angle based on single-point energy
calculations (shown later in Fig. 8) also shows that a
conformation with a dihedral of 171.1° is the most energetically favorable for this form of the reduced flavin in the
protein environment, partly because it leads to more
favorable hydrogen bonds. Probably, the bent reduced
conformation (dihedral angle ⫽ 151.3°) of the free isoalloxazine ring never occurs in the protein environment. The
protein environment seems to be ideally suited to adjust
the charges, whereas the geometry and the electrostatic
potential seem to stabilize a more planar conformation of
the isoalloxazine ring.48
Figure 5(B) shows that upon reduction the negative
charges of the N1F, O2F, and O4F atoms is increased by
the charge of the two reducing electrons migrating through
the unoccupied path present in the oxidized FAD [LUMOs
in Fig. 5(A)]. The occupied molecular orbitals in the
reduced form [HOMOs in Fig. 5(B)] however, have a
similar configuration as in the isolated reduced isoallox-
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
427
Fig. 5. Initial exploratory RHF/STO-3G results for the electronic configurations of the isoalloxazine ring in the protein environment (cf. Fig. 3): (A)
oxidized isoalloxazine ring in the enzyme catalytic site; (B) reduced isoalloxazine ring in the enzyme catalytic site, the initial step of the charge relay [cf.
Fig. 3(B)]; (Bⴕ) reduced isoalloxazine ring in the enzyme catalytic site, the intermediate step in the charge-relay mechanism [cf. Fig. 3(B,C)], with a proton
transferred from Tyr-155 to O2F; and (C) reduced isoalloxazine ring in the enzyme catalytic site, the completion of the charge relay [cf. parts B and B⬘ and
Fig. 3(C)].
azine ring, with occupancy around the whole molecule.
However, the negative charges and potential show a
tendency to increase with respect to the oxidized form,
mainly in N1F, O2F, and O4F [Fig. 5(B)]. The large charge
separation that would exist if this configuration remained
as shown (i.e., no charge relay) can be seen in a comparison
of the charges and the negative electrostatic potential of
this form with those of the reduced ring in Figure 5(C).
At this stage, the result of the reduction is the migration
of the charge of the reducing electrons from N5F to N1F,
428
G. CAVELIER AND L.M. AMZEL
Fig. 6. DFT B3LYP/6-31G(d) electronic structure results for the isoalloxazine ring in the protein environment. The three configurations shown
correspond to those of Figure 5(A–C).
O2F, and O4F. Despite the ring planarity imposed by the
protein, there is still a charge separation, but it is only
temporary, and further compensating mechanisms are put
in place by the protein environment. Note in Figure 5(B)
that a LUMO is present at residue His-161, a residue
postulated to cooperate with Tyr-155 in the establishment
of a charge-relay system that helps to neutralize the initial
charge separation.5 The charge stabilization is brought
about by the migration of the proton of the OH of Tyr-155
to O2F; stabilization of the negative charge left in Tyr-155
is accomplished by either proton migration from or direct
charge stabilization by His-161.
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
Fig. 7. Effects of geometry and dielectric: (A) hydrogen-bonded
waters partially account for the effect of the solvent around the isoalloxazine ring and (B) the dielectric effect is taken into account with the SCRF
method. Compare the effect of the waters or the SCRF with the effect of
the protein in Figure 5(C).
An intermediate step in the charge-relay mechanism
can be observed in Figure 5(B⬘). This result correspond to
the case in which the proton from Tyr-155 has been
transferred to the O2F in the ring, but the proton of
His-161 has not been transferred to the Tyr-155 phenolic
oxygen, which remains negatively charged. Note in Figure
5(B⬘) that there is an additional charge in O2F, N3F, O4F,
and related carbons and hydrogens, with respect to the full
charge relay shown in Figure 5(C). However, this intermediate step shows that the sole transfer of a proton from
Tyr-155 to O2F is enough to cause a neutralization of the
initial charge separation in the ring [see the corresponding
charges in Fig. 5(B)]. Moreover, occupied HOMOs migrate
from the isoalloxazine ring to Tyr-155, with a corresponding negative potential around the side chain of this residue: its oxygen has a charge of ⫺0.531.
As a result of the charge relay [Fig. 5(C)], a proton is
transferred from Tyr-155 to O2F and stabilizes the enol
form. In turn, His-161 stabilizes the charge in Tyr-155.
Partial charges of the isoalloxazine atoms [Fig. 5(C)] are
changed with respect to those of the oxidized, planar form
[Fig. 5(A)] and much more changed with respect to those of
the configuration obtained before the charge relay [Fig.
5(B)]. The difference in the calculated energies between
the structure before [Fig. 5(B)] and after [Fig. 5(C)] the
charge relay is ⫺9.9 kcal/mol, indicating that the charge
429
relay has a large stabilizing effect on the reduced state. A
comparison of Figure 5(C) with the case of the free reduced
isoalloxazine ring [Fig. 3(C)] shows greater negative
charges in N1F, O2F, N3F, and O4F that may result in
strong hydrogen bonds. Similar charges to the ones found
in the isolated ring of Figure 3(C) are found when the
electronic structure is calculated for the unstable, planar
configuration of Figure 5(C) taken outside of the protein
(not shown). Therefore, the increased charges of the bound
cofactor probably reflect the planar configuration of the
isoalloxazine ring stabilized by the binding energy provided by the protein [Fig. 5(C), dihedral angle ⫽ 175°].
This planar configuration may in turn contribute to the
enhanced reducing power of the FADH2 in the protein and
the rate of the second step of the QR1 reaction (FADH2
reducing the quinones; see the Stabilization Energy section later).
Comparing Figure 5(A,B) with Figure 5(C) suggests that
the energetically favorable interactions result from the
electrostatic potential around atoms N5F, N1F, O2F, and
O4F. As a result of the charge-relay mechanism, a corresponding equalization (spreading) of the HOMO in the
isoalloxazine ring is observed. The negative charges of
N1F, O2F, and O4F are lower in magnitude than before
the changes associated with the charge relay [Fig. 5(B)],
and the adjoining carbon atoms are more positive, such
that the cofactor shows mainly a weakly positive electrostatic potential relative to the rest of the environment [Fig.
5(C)]. The HOMOs are distributed evenly throughout the
isoalloxazine; in particular, there is a strong HOMO at the
N5F atom overlapping with the other HOMOs of the ring.
These HOMOs hold the electronic charge for the quinone
reduction that can occur after NAD(P)⫹ leaves and the
quinone arrives in its place. The LUMOs are no longer at
His-161; they are now in the isoalloxazine ring.
As in the free isoalloxazine ring, calculations at the
B3LYP/6-31G(d) level of theory give very similar results
[Fig. 6(A–C)] to those in Figure 5(A–C). Like the other
cases discussed previously (Figs. 3 and 4), the main
difference is evidenced in the Mulliken charges.
Effects of Geometry and Dielectric
To study the energetic influence of the protein environment on the conformation of the isoalloxazine ring, singlepoint energy calculations were made for different bent
configurations of the ring at several dihedral angles. They
were performed at the RHF/STO-3G level for the isolated
ring and for the ring anchored in the catalytic site of the
enzyme. Although this is a minimal basis set and thus the
computed energies may have a bias, it has been found that
energy differences from RHF/STO-3G calculations may
have the required accuracy.46 Therefore, in this analysis
only differences in energy with respect to a minimum were
used. To represent more closely the aqueous environment
for the free isoalloxazine ring (not in the protein environment), we performed calculations with an approximate
method that employs a bulk dielectric constant to represent the solvent: the SCRF method.46 A relative dielectric
constant of 80 was used in calculations with the SCRF
430
G. CAVELIER AND L.M. AMZEL
Fig. 8. Stabilization energy. The bent configuration of the free isoalloxazine ring (circles) is destabilized in
the protein environment, and the planar configuration of the reduced isoalloxazine ring (black squares) is
stabilized by the protein environment. Finally, the more planar configuration of the protonated enol form in the
catalytic site (black triangles) is stabilized by the protein.
method. The relaxed optimized configurations of the rings
found before were the basis for these single-point energy
calculations at the RHF/STO-3G level with the SCRF
method. These results are shown in Figure 7(A–B) and are
further discussed later.
Calculations with hydrogen-bonded waters, as described
in the Materials and Methods section, were also performed
as an alternative method to partially account for the effect
of the solvent around the free isoalloxazine ring outside
the protein. The results with hydrogen-bonded waters
[corresponding to the reduced structure of Figure 2(F) with
a proton in N1F, the most likely protonated solution form]
are shown in Figure 7(A) and can be compared to the
visualizations obtained with SCRF electronic structure
calculations on the isolated reduced isoalloxazine ring
[Fig. 7(B)]. The hydrogen-bonded waters provide an environment similar to the hydrogen bonds of the protein
environment: the charges, molecular orbitals, and electrostatic potential are similar to those of Figure 5(C), the
result of the charge relay inside of the protein, where the
protonation is only possible at O2F, not at N1F. Some
charges and negative potentials are much higher than
inside the protein because of the negative charge of the
oxygens in the waters [Fig. 7(A)]. In general, the trend
from in vacuo (not shown) to the SCRF calculation [Fig.
7(B)] to the water hydrogen-bonded structure [Fig. 7(A)] is
toward an increase in the charges of negative atoms such
as N1F, N5F, O2F, and O4F, together with an increase in
the positive charge of positive atoms surrounding them,
such as the hydrogen in N1F, the carbon in C2F, and the
hydrogen at N5F. The molecular orbitals in the three cases
are very similar to the one in Figure 5(C) (protein environment, protonated at O2F). The calculations performed
with surrounding waters and those performed with SCRF
show similar trends and magnitudes of the changes in the
charges.
Stabilization Energy
The energies associated with a number of conformations
of the reduced isoalloxazine ring were calculated for the
free ring and the ring in the protein environment before
[Figs. 3(B)–5(B)] and after the charge relay [Figs. 3(C)–
5(C)]. The total energy was calculated for different degrees
of bending (the angle between the planes of the two outer
isoalloxazine rings) to estimate the energetic contribution
that the protein makes to catalysis by imposing a planar
form to the reduced isoalloxazine (Fig. 8). The most stable
conformations of the bound reduced cofactor have the
following bending angles: reduced form before the charge
relay, 171°, and after the charge relay, 175° (Fig. 8). In the
free isoalloxazine, these conformations have energies of
1.9 and 2.5 kcal/mol, respectively, above the energy of the
most stable conformation (151°). Thus, interactions with
the protein provide close to 2 kcal/mol for the stabilization
of the planar reduced form. Equally important, the stable
conformation of the free isoalloxazine (bent angle ⫽ 151°)
is destabilized in the protein environment by 14.8 kcal/mol
(Fig. 8).
The interaction of the isoalloxazine ring with rQR1 also
has an effect on the energy of reduction. Thus, the analysis
can be complemented by the estimation of the difference in
energy between the reduction of the isolated/solvated and
431
AB INITIO STUDY OF NAD(P)H:QUINONE REDUCTASE
TABLE II. Energies in the Thermodynamic Cycle:
Stabilization Energy
Reduced FAD
configuration
FADH2
FADH2 (SCRF)
Fig. 9. Stabilization energy and reduction energy of the flavin: the
thermodynamic cycle considered for FAD reduction isolated and in the
QR environment. This thermodynamic cycle permits us to calculate an
experimentally inaccessible quantity, ⌬E2 ⫺ ⌬E4, by the results of the
quantum calculations, because ⌬E2 ⫺ ⌬E4 is equal to ⌬E1 ⫺ ⌬E3, and the
latter can be found from the calculations. Species QR is the enzyme
environment alone, with the oxidized flavin isoalloxazine ring deleted.
Species QR⬘ is the same enzyme environment but with the reduced flavin
isoalloxazine ring deleted. It is different from QR because Tyr-155 in the
protein environment has transferred a proton to the reduced flavin as part
of the charge-transfer relay.
TABLE I. Species Considered in the Thermodynamic Cycle
Reduced (R)
Oxidized (OX)
Isolated
FAD
FAD in QR
environment
QR
environment
alone
FAD (R)
FAD (OX)
FAD (R) 䡠 QR⬘
FAD (OX) 䡠 QR
QR⬘
QR
protein-bound isoalloxazine rings (⌬Ebound ⫺ ⌬Efree). This
energy difference can be estimated by the study of the
thermodynamic cycle shown in Figure 9. Calculations
based on this cycle with the species shown in Table I
overcome the problem of calculating energy differences
between structures that have a different number of atoms.
The structures correspond to the cases shown in Figures
3(A,C) and 5(A,C). The energies of the thermodynamic
cycle were calculated with the various model chemistries
explained in the Materials and Methods section.
It is clear from Figure 9 that
⌬E1 ⫺ ⌬E3 ⫽ ⌬E2 ⫺ ⌬E4
where
⌬E1 ⫽ EFAD(R)䡠QR⬘ ⫺ (EQR⬘ ⫹ EFAD(R))
⌬E3 ⫽ EFAD(OX)䡠QR ⫺ (EQR ⫹ EFAD(OX))
QR⬘ refers to the protein environment QR without the
proton, which has been transferred in the first step of the
charge relay to O2F. The results of the calculations in
vacuo and with an SCRF are shown in Table II. The
contribution of the enzyme to the FAD two-electron obligatory reduction with respect to the reduction of the isolated
FAD (i.e., ⌬E2 ⫺ ⌬E4, which is equal to ⌬E1 ⫺ ⌬E3) is to
make the reduction less favorable by 2.39 – 4.74 kcal/mol
(Table II). That means that the reduction potential of the
rQR1 flavin is 50 –100 mV more negative than that of the
⌬E1
(kcal/mol)
⌬E3
(kcal/mol)
⌬E1 ⫺ ⌬E3 ⫽
⌬E2 ⫺ ⌬E4
(kcal/mol)
-
⫺2.48
8.18
4.74
2.39
free flavin. This change may be very important for the
function of the enzyme. Even with this change of potential,
o⬘
the bound flavin can be reduced by NAD(P)H (⌬Efree
䡠 flavin
o⬘
⫽ ⫺180 mV; ⌬Efree 䡠 NAD(P) H ⫽ ⫺320 mV;). However,
having a more negative reduction potential, the rQR1bound flavin has a greater driving force for the reduction of
quinones, allowing the enzyme to reduce a large variety of
quinones, even those with a large negative reduction
potential. This conclusion agrees with the observed broad
range of quinones that can be reduced by QR1. In addition,
a greater driving force may result in a more efficient
hydride transfer or higher catalytic rates for that step.
This study, in addition to providing valuable insights
into the mechanism of QR, presents an excellent example
of the contribution that quantum mechanical calculations
can make to understanding enzymatic mechanisms.
ACKNOWLEDGEMENTS
The authors thank Mario Bianchet for his always useful
and generous collaboration and B. Brooks and his group
for helpful discussions and suggestions. Thanks are also
due to the Advanced Biomedical Computing Center of the
National Cancer Institute (Frederick, MD) for the Cray
computer time and some of the software used and a very
efficient and friendly system administration. G. Cavelier’s
work was supported in part by an NIGMS/NIH-National
Research Service Award.
REFERENCES
1. Ross D, Siegel D, Beall H, Prakash AS, Mulcahy RT, Gibson NW.
DT-diaphorase in activation and detoxification of quinones. Bioreductive activation of mitomycin C. Cancer Metastasis Rev 1993;12:
83–101.
2. Knox RJ, Boland MP, Friedlos F, Coles B, Southan C, Roberts JJ.
The nitroreductase enzyme in Walker cells that activates 5-(aziridin-1-yl)-2,4-dinitrobenzamide (CB 1954) to 5-(aziridin-1-yl)-4hydroxylamino-2-nitrobenzamide is a form of NAD(P)H dehydrogenase (quinone) (EC 1.6.99.2). Biochem Pharmacol 1988;37:-. Knox RJ, Friedlos F, Jarman M, Roberts JJ. A new cytotoxic, DNA
interstrand crosslinking agent, 5-(aziridin-1-yl)-4-hydroxylamino2-nitrobenzamide, is formed from 5-(aziridin-1-yl)-2,4-dinitrobenzamide (CB 1954) by a nitroreductase enzyme in Walker carcinoma cells. Biochem Pharmacol 1988;37:-. Tedeschi G, Chen S, Massey V. DT-diaphorase. Redox potential,
steady-state, and rapid reaction studies. J Biol Chem 1995;270:
1198 –1204.
5. Li R, Bianchet MA, Talalay P, Amzel LM. The three-dimensional
structure of NAD(P)H:quinone reductase, a flavoprotein involved
in cancer chemoprotection and chemotherapy: mechanism of the
two-electron reduction. Proc Natl Acad Sci U S A 1995;92:8846 –
8850.
6. Talalay P, De Long MJ, Prochaska HJ. Identification of a common
chemical signal regulating the induction of enzymes that protect
against chemical carcinogenesis. Proc Natl Acad Sci U S A
1988;85:-.
432
G. CAVELIER AND L.M. AMZEL
7. Skelly JV, Sanderson MR, Suter DA, Baumann U, Read M,
Gregory DS, Bennett M, Hobbs SM, Neidle S. Crystal structure of
human DT-diaphorase: a model for interaction with the cytotoxic
prodrug 5-(aziridin-1-yl)-2,4-dinitrobenzamide (CB1954). J Med
Chem 1999;42:-. Faig M, Bianchet MA, Talalay P, Chen S, Winski S, Ross D, Amzel
LM. Structures of recombinant human and mouse NAD(P)H:
quinone oxidoreductases: species comparison and structural
changes with substrate binding and release. Proc Natl Acad Sci U
S A 2000;97:-. Foster CE, Bianchet MA, Talalay P, Zhao Q, Amzel LM. Crystal
structure of human quinone reductase type 2, a metalloflavoprotein. Biochemistry 1999;38:-. Walsh CT. Enzymatic reaction mechanisms. San Francisco: Freeman;- p.
11. Bugg T. An introduction to enzyme and coenzyme chemistry.
Oxford: Blackwell Science;- p.
12. Silverman RB. The organic chemistry of enzyme-catalyzed reactions. San Diego: Academic;- p.
13. Lee T-S, York DM, Yang W. Linear-scaling semiempirical quantum calculations for macromolecules. J Chem Phys 1996;105:2744 –
2750.
14. Sánchez-Portal D, Ordejón P, Artacho E, Soler JM. Densityfunctional method for very large systems with LCAO basis sets.
Int J Quantum Chem 1997;65:-. Ordejón P, Artacho E, Soler JM. Self-consistent order-N densityfunctional calculations for very large systems. Phys Rev B 1996;53:
R10441–R-. Ordejón P, Drabold DA, Martin RM, Grumbach MP. Linear
system-size scaling methods for electronic-structure calculations.
Phys Rev B 1995;51:1456 –1476.
17. Kohen A, Klinman JP. Enzyme catalysis: beyond classical paradigms. Acc Chem Res 1998;31:-. Kohen A, Cannio R, Bartolucci S, Klinman JP. Enzyme dynamics
and hydrogen tunneling in a thermophilic alcohol dehydrogenase.
Nature 1999;399:496 – 499.
19. Ringe D, Petsko GA. Tunnel vision. Nature 1999;399:-. Yagi K. Hydrogen bonding in flavoproteins. Chem Scr A 1987;27:
27–30.
21. Mestres J, Duran M, Bertran J. Characterization of the transition
state for the hydride transfer in a model of the flavoprotein
reductase class of enzymes. Bioorg Chem 1996;24:69 – 80.
22. Zheng Y-J, Ornstein RL. A theoretical study of the structures of
flavin in different oxidation and protonation states. J Am Chem
Soc 1996;118:-. Niemz A, Imbriglio J, Rotello VM. Model systems for flavoenzyme
activity: one- and two-electron reduction of flavins in aprotic
hydrophobic environments. J Am Chem Soc 1997;119:-. Breinlinger EC, Keenan CJ, Rotello VM. Modulation of flavin
recognition and redox properties through donor atom interactions.
J Am Chem Soc 1998;120:8606 – 8609.
25. Vishveshwara S, Jacob R, Nadig G, Maizel JV Jr. The role of
lysine-41 in RNase A catalysis. A quantum chemical study on the
active site–ligand complex. J Mol Struct 1998;471:1–11.
26. Antony J, Medvedev DM, Stuchebrukhov A. Theoretical study of
electron transfer between the photolyase catalytic cofactor FADH⫺
and DNA thymine dimer. J Am Chem Soc 2000;122:-. Várnai P, Richards WG, Lyne PD. Modelling the catalytic reaction
in human aldose reductase. Proteins 1999;37:218 –227.
28. Chatfield DC, Eurenius KP, Brooks BR. HIV-1 protease cleavage
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
mechanism: a theoretical investigation based on classical MD
simulation and reaction path calculations using a hybrid QM/MM
potential. J Mol Struct 1998;423:79 –92.
Liu H, Müller-Plathe F, Gunsteren WFv. A combined quantum/
classical molecular dynamics study of the catalytic mechanism of
HIV protease. J Mol Biol 1996;261:454 – 469.
Tapia O, Andrés J, Safont VS. Transition structures in vacuo and
the theory of enzyme catalysis. Rubisco’s catalytic mechanism: a
paradigmatic case? J Mol Struct 1995;342:131–140.
Li J, Nelson MR, Peng CY, Bashford D, Noodleman L. Incorporating protein environments in density functional theory: a selfconsistent reaction field calculation of redox potentials of [2Fe2S]
clusters in ferredoxin and phthalate dioxygenase reductase. J
Phys Chem A 1998;102:-.
Rick SW, Abashkin YG, Hilderbrandt RL, Burt SK. Computational studies of the domain movement and the catalytic mechanism of thymidine phosphorylase. Proteins 1999;37:242–252.
Walsh CT. Flavin coenzymes: at the crossroads of biological redox
chemistry. Acc Chem Res 1980;13:148 –155.
Ghisla S, Massey V. Mechanisms of flavoprotein-catalyzed reactions. Eur J Biochem 1989;181:1–17.
Murthy YV, Massey V. Synthesis and properties of 8-CN-flavin
nucleotide analogs and studies with flavoproteins. J Biol Chem
1998;273:-.
Meyer M. Ab initio study of flavonoids. Int J Quantum Chem
2000;76:724 –732.
Meyer M. Density functional study of isoalloxazine and C4ahydroperoxidihydroisoalloxazine. J Mol Struct 1997;417:163–168.
Wouters J, Durant F, Champagne B, André J-M. Electronic
properties of flavins: implications on the reactivity and absorption
properties of flavoproteins. Int J Quantum Chem 1997;64:721–
733.
Nakai S, Yoneda F, Yamabe T. Theoretical study on the lowestfrequency mode of the flavin ring. Theor Chem Acc 1999;103:109 –
116.
Frisch MJ, Trucks GW, Schlegel HB, et al. Gaussian 98. Pittsburgh, PA: Gaussian;- p.
Becke AD. Density-functional exchange-energy approximation
with correct asymptotic behavior. Phys Rev A 1988;38:3098 –3100.
Lee C, Yang W, Parr RG. Development of the Colle–Salvetti
correlation-energy formula into a functional of the electron density. Phys Rev B 1988;37:785–789.
Bruice TC. Mechanisms of flavin catalysis. Acc Chem Res 1980;13:
256 –262.
Nishimoto K. Ab-initio MO study of the redox function of flavin.
In: Nagata C, editor. Biomolecules. Tokyo: Japan Scientific Societies Press; 1985. p 9 –19.
Wong MW, Frisch MJ, Wiberg KB. Solvent effects 1. The mediation of electrostatic effects by solvents. J Am Chem Soc 1991;113:
4776 – 4782.
Foresman JB, Frisch A. Exploring chemistry with electronic
structure methods. Pittsburgh, PA: Gaussian;- p.
Werner P-E, Ronnquist O. The crystal structure of 5-acetyl-9bromo-1,3,7,8,10-pentamethyl-1,5-dihydroisoalloxazine. Acta
Chem Scand 1970;24:-.
Lennon BW, Williams CH, Ludwig ML. Crystal structure of
reduced thioredoxin reductase from Escherichia coli: structural
flexibility in the isoalloxazine ring of the flavin adenine dinucleotide cofactor. Protein Sci 1999;8:2366 –2379.