Culka 2017
Culka 2017
Computational Biochemistry—
Enzyme Mechanisms Explored
Martin Culka2, Florian J. Gisdon2, G. Matthias Ullmann1
Computational Biochemistry, University of Bayreuth, Bayreuth, Germany
1
Corresponding author: e-mail address: [email protected]
Contents
1. Introduction 78
2. Structural Models 81
2.1 Continuum-Electrostatics Models 82
2.2 Quantum-Mechanical Models 83
2.3 Empirical Molecular-Mechanical Models 84
2.4 Hybrid QM/MM Models 86
2.5 Pseudo-Atomistic Models 89
3. Calculating Enzymatic Mechanisms 91
3.1 Thermodynamic Properties of Biological Systems 91
3.2 Kinetic Properties of Biochemical Systems 93
3.3 Mechanic and Energetic Properties of Molecular Reaction Paths 97
3.4 Path Search Strategy 104
4. Going Beyond the Exploration of the Reaction Paths 104
References 105
Abstract
Understanding enzyme mechanisms is a major task to achieve in order to comprehend
how living cells work. Recent advances in biomolecular research provide huge amount
of data on enzyme kinetics and structure. The analysis of diverse experimental results
and their combination into an overall picture is, however, often challenging. Micro-
scopic details of the enzymatic processes are often anticipated based on several hints
from macroscopic experimental data. Computational biochemistry aims at creation of a
computational model of an enzyme in order to explain microscopic details of the cat-
alytic process and reproduce or predict macroscopic experimental findings. Results of
such computations are in part complementary to experimental data and provide an
explanation of a biochemical process at the microscopic level. In order to evaluate
the mechanism of an enzyme, a structural model is constructed which can be analyzed
by several theoretical approaches. Several simulation methods can and should be com-
bined to get a reliable picture of the process of interest. Furthermore, abstract models of
2
These authors contributed equally.
Advances in Protein Chemistry and Structural Biology, Volume 109 # 2017 Elsevier Inc. 77
ISSN 1876-1623 All rights reserved.
http://dx.doi.org/10.1016/bs.apcsb.2017.04.004
78 Martin Culka et al.
1. INTRODUCTION
“Certainly no subject or field is making more progress on so many
fronts at the present moment than biology, and if we were to name the most
powerful assumption of all, which leads one on and on in an attempt to
understand life, it is that all things are made of atoms, and that everything
that living things do can be understood in terms of the jigglings and wig-
glings of atoms” (Feynman, 1964). These words were said by the well-
known physicist Richard Feynman in his famous Lectures on Physics now
more than 50 years ago. Until today, this sentence has not lost its validity
and is the basis for much of the biomolecular research, maybe today even
more than 50 years ago. Today, we are beginning to understand how the
interplay of atoms and molecules lead to the complex processes that we find
in living systems. Recent advances in genomics and systems biology help to
gain more and more insights into the molecular organizations of the cell,
which are nevertheless still too complex to allow a full overview of all
the biochemical processes at atomic detail. New techniques in structural
biology such as the free-electron laser allow to analyze structurally the kinet-
ics of molecular processes at an atomic level (Nango et al., 2016; Pande et al.,
2016). On the other hand, modern electron microscopy allows us to gain
insights into large molecular assemblies that were not accessible to structural
investigations a few years ago (Bartesaghi et al., 2015; K€ uhlbrandt, 2014).
These new techniques complement traditional techniques of structural biol-
ogy such as X-ray crystallography and NMR. In the past, a structural char-
acterization of a protein was considered the final goal of an investigation.
Nowadays to gain a deep understanding of an enzymatic mechanism, a
structure is just the beginning. The structural information needs to be com-
pleted by other experimental information for instance from spectroscopy,
kinetics, or electrochemistry. Often, all these different aspects are difficult
to merge. Thus, it is important to approach the enzymatic mechanism also
from a theoretical side. With the help of modern methods from computa-
tional chemistry, it is possible to gain insights into enzymatic mechanisms
and to complete the picture.
Computational Biochemistry—Enzyme Mechanisms Explored 79
Experimental
data Numerical
mathematics
Analytical
Simulation solution
data
Fig. 1 Building theoretical models of real systems. This scheme depicts the stages of
building theoretical models that can help to analyze real systems. These theoretical
models are abstractions of the real systems that help to understand the behavior of
the system. Mathematical modeling allows a quantitative comparison between the real
system and the model.
model using methods from numerical mathematics. This discrete model can
then be implemented in computer programs. The computer model can then
be used to perform computer simulations and calculations using sets of
parameters. The resulting simulation data can be compared with experimen-
tal data and so the quality of the model can be judged. The quantities that
have been calculated from the model can lead to a better understanding of
the real system. Moreover, the analysis of the effect of well-defined changes
of the initial parameter can provide deeper insights into the behavior of the
real system.
One of the most basic model that is widely used in the description of
chemical systems is the Born–Oppenheimer approximation according to
which one can separate the motion of nuclei from the motion of the elec-
trons. This approximation allows us to describe molecular structures by
defining the nuclear coordinates and to speak about different electronic
states of a molecule. One fundamental concept that is used in molecular bio-
physics is the molecular energy landscape which describes the energy of a
molecular system with N atoms in dependence of its 3N coordinates. This
energy landscape contains minima which represent stable states of the sys-
tems, (first-order) saddle points which represent transition states of chemical
reactions, and minimum energy paths which represent the trajectory along
which chemical reactions occur. A large part of theoretical biochemistry is
concerned with exploring this energy landscape and to extract different kind
Computational Biochemistry—Enzyme Mechanisms Explored 81
of data from it. In this article, we review some of the most important
methods to explore this energy landscapes and to explain how they can
be used to explore enzymatic mechanisms.
2. STRUCTURAL MODELS
The most computational enzymology techniques construct a structural
model of the enzyme in question, usually using its experiment-derived 3D
structure as a starting point. A structural model provides a way to calculate
energy of the given structure, by using various approximations of the real
picture to stay computationally feasible. Based on energy differences
between different states within structural computer models of enzymes, ther-
modynamics, and kinetics of enzyme catalysis can be addressed. With the use
of structural models (Fig. 2), researchers aim to track the events within the
Implicit
Continuum-electrostatics
Pseudo-atomistic mechanics
Structural models
Molecular mechanics
Quantum mechanics
Explicit
ns
es
ls
es
om
el
tro
ul
ul
ex
C
ec
ec
At
ec
pl
ol
ol
om
El
C
ro
ac
M
Fig. 2 Structural models for the simulation of catalytic events within enzymes. The ver-
tical axis represents an ascending gradient from explicit to implicit models. The system
description lose the detailed (explicit) character in favor of averaged (implicit) descrip-
tion of physical properties. The horizontal axis shows the information that can be
obtained from the models, and together illustrates the size of the biological system,
which is computationally feasible to treat. The model bars indicate possible applications.
The fading areas show the potential of future developments enabled by improvements
of algorithmic and computational power. All models can be combined into the so-called
hybrid or multiscale models, which treat different parts of the biological system with the
appropriate method, to enhance the effectiveness of the simulation and the quality of
results.
82 Martin Culka et al.
boundary between the protein and the solvent, as well as from the distribu-
tion of ions in the solution. Electrostatic energy can be obtained by integrat-
ing the potential distribution over the space.
Continuum-electrostatics models such as the above described Poisson–
Boltzmann model have various applications. The most straightforward is cal-
culation of solvation energies and visualization of electrostatic potentials
(Baker, Sept, Joseph, Holst, & McCammon, 2001). More advanced appli-
cations allow to calculate energies of different protonation and oxidation
microstates of the protein, as well as ligand binding energies and many other
applications (Ullmann et al., 2008). Furthermore, the Poisson–Boltzmann
model or other popular continuum models such as COSMO (Klamt &
Sch€ €rmann, 1993) are often combined with more detailed models to sim-
uu
ulate solvent effects in a computationally affordable way (Chen, Noodleman,
Case, & Bashford, 1994; Li, Nelson, Peng, Bashford, & Noodleman, 1998;
Liu et al., 2004).
! X X X
V ðR Þ ¼ Kb ðb b0 Þ2 + Kθ ðθ θ0 Þ2 + Kφ ð1 + cos ðnφ δÞÞ
bonds angles torsions
8 9
>
> >
>
>
> >
>
>
> >
>
>
> 2 >
>
>
> !12 !6 3 >
>
X X < Rijmin Rijmin =
εijmin 4 5 qi qj
+ Kω ðω ω0 Þ2 + 2 +
>
> >
0 εrij >
r r
nonbonded> >
ij ij 4πε
improper >
> |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl {zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
ffl } |fflfflfflffl{zfflfflfflffl} >
>
>
> >
>
>
>
pairs > >
>
torsions
: LennardJones Coulomb >
;
potential law
(2)
where b is the bond length, θ is the bond angle, ϕ the torsion angle, ω the
improper torsion angle. The 0 indices represent the equilibrium values of
given parameters and K the respective spring constants. Van der Waals
interactions are represented by the Lennard–Jones potential where Rijmin is
the distance energy minimum position, rij is the distance between two
atoms and εminij is the interaction constant. Electrostatic interactions are
described by the Coulomb law where qi and qj are the partial charges of
the interacting atoms, rij is their distance and ε0ε is the environment
permittivity.
Several empirical MM force fields for biomolecules have been devel-
oped over the years, e.g., CHARMM (MacKerell et al., 1998), AMBER
(Cornell et al., 1996), GROMOS (Oostenbrink, Villa, Mark, & Van
Gunsteren, 2004), or OPLS (Jorgensen, Maxwell, & Tirado-Rives,
1996). MM models proved to be very useful especially for studying protein
dynamics by molecular dynamics (MD) simulations. In context of enzyme
catalysis, pure MM models can be used to study conformational changes or
substrate binding dynamics (Costa, Batista, Bisch, & Perahia, 2015;
86 Martin Culka et al.
Gilson & Zhou, 2007). When one wants to model details of a catalytic
mechanism in the enzyme active site, however, the conventional MM
models are insufficient. They are, however, often used to model the non-
reactive parts of the enzyme in hybrid QM/MM approaches, as described
below.
Nevertheless, attempts have been made to allow for chemical reactivity
within the empirical force fields without employing QM principles. These
approaches are referred to as reactive force fields. One of the most widely
used reactive force field, that has been also utilized for biological systems,
is called ReaxFF (Senftle et al., 2016; van Duin, Dasgupta, Lorant, &
Goddard III, 2001). In ReaxFF, the empirical energy equation (Eq. 2) is
modified to follow a more complex bond order formalism rather than the
balls-on-springs formalism. Although reactive force fields have been origi-
nally developed for material chemistry applications (Liang et al., 2013), first
attempts have been made to model peptides, small proteins (Golkaram,
Shin, & van Duin, 2014; Monti et al., 2013), and DNA (Verlackt et al.,
2015) using ReaxFF.
Empirical force fields are popular because of their speed and their often
realistic description of molecules, which is a consequence of extensive
parameterization to reproduce experimentally derived structures and behav-
iors. It has been shown that MM methods are able to reproduce experimen-
tally derived molecular structures, except for nonstandard regions (Kulik,
Luehr, Ufimtsev, & Martı́nez, 2012). In such regions, ab initio QM methods
consistently perform better. Despite this benefit, as already mentioned
above, full ab initio studies of proteins are restricted to rather small systems
(Kulik et al., 2012). Instead, hybrid approaches (referred to as QM/MM)
have been developed to combine accurate ab initio models and fast empirical
models.
within the MM model, the MM region can react to it. A major drawback is,
however, that the QM calculation is performed in the absence of the elec-
trostatic MM environment, thus the atoms in the QM part cannot react to
their full environment. But especially this electrostatic environment is cru-
cial for enzymatic catalysis (Zhang, 2013). The electrostatic embedding
approach (Bakowies & Thiel, 1991) treats this important interaction by pro-
viding the electrostatic environment for the QM calculation. The electro-
static environment appears as a one-electron term in the QM Hamiltonian.
Thus the charge distribution in the QM region is polarized according to the
MM charges. Despite the higher accuracy in the calculation, charge leakage
effects can occur at the boundary of the QM region, where the QM charge
density is polarized in immediate proximity by MM charges. In both embed-
ding schemes the MM charges are rigid and do not react to the QM charge
density. In a polarized embedding scheme the polarization happens in both,
the QM region direction and the MM region direction. For the polarization
of the MM region, a polarizable force field has to be applied on the MM level
(Thompson & Schenter, 1995). This approach of polarized embedding is in
general the most accurate, though computational demanding one.
In the majority of QM/MM simulations of enzymes, one has to deal with
covalent bonds on QM/MM boundary (e.g., bond between an amino acid
side chain and the protein backbone) in addition to the nonbonded inter-
actions. The boundary QM atom valency needs to be saturated to allow
for proper electronic structure calculation. The simplest approach to treat
the QM/MM boundary is the link atom approach (Field, Bash, &
Karplus, 1990; Singh & Kollman, 1986), where the QM atom is capped
by an auxiliary atom (usually hydrogen) that is constraint in the direction
of the QM/MM bond. This generates a problem of QM density over-
polarization near the boundary that has to be treated. Alternative approaches
are based on frozen hybrid orbitals (Amara, Field, Alhambra, & Gao, 2000;
Thery, Rinaldi, Rivail, Maigret, & Ferenczy, 1994). A frontier atom is cho-
sen in the QM/MM boundary and a set of suitably oriented localized orbitals
is placed on it. This treatment allows to converge on more proper electronic
structure in the boundary region. On the other hand frozen orbital appro-
aches are more technically demanding and require calibration for the specific
bond and QM method.
A conceptually different QM/MM approach that has been used to study
enzyme catalysis is the empirical valence bond (EVB) (Kamerlin & Warshel,
2010, 2011; Warshel, 2003). EVB uses the valence-bond (VB) approach for
quantum description of the enzyme active site. The QM methods discussed
Computational Biochemistry—Enzyme Mechanisms Explored 89
so far are based on molecular orbital theory that combines the atomic orbitals
into a molecular orbital wave-function. The VB methods instead describe
the system as a linear combination of all possible states where electrons
occupy localized orbitals. In EVB, the stationary points along the reaction
path are described by an MM force field, while the transitions between them
are treated by an SE valence-bond QM approach (Shurki, Derat, Barrozo, &
Kamerlin, 2015). Thus for every reaction step simulated by EVB, a set of
empirical parameters has to be derived. This is usually done on a model reac-
tion in solution or in gas phase, where the parameters are fitted to reproduce
the experimental data or ab initio QM results (Åqvist & Warshel, 1993).
Once well calibrated, extensive conformational sampling along the reaction
pathway can be achieved in reasonable time with EVB in order to get a
proper free energy landscape (see Section 3.3.2). On the other hand, the
major disadvantage of EVB in comparison to common QM and QM/
MM methods is the need of specific calibration for every reaction step. In
fact, prior knowledge of the reaction mechanism is required to perform
an EVB simulation, while unknown mechanism alternatives can be discov-
ered within conventional ab initio QM and QM/MM models.
ΔG ¼ ΔH T ΔS ¼ ΔU + PV T ΔS (6)
ΔA ¼ ΔU T ΔS (7)
Fi ¼ ri E ¼ mi ai (8)
The force Fi acting on atom i with mass mi determines its acceleration ai,
while the force is defined by the negative gradient of the potential energy
function riE with respect to the coordinates of atom i. By integrating
the equations of motion for all atoms over small time steps, it is possible
to obtain their time dependent positions, just providing the initial positions
and the initial velocities. The initial positions are directly given by the coor-
dinates of the atoms, and the initial velocities are usually distributed ran-
domly with a certain probability distribution. Only the acceleration is
needed, which can be obtained by the gradient of the potential energy func-
tion and the atomic mass (Eq. 8). The velocities of the atoms define the
temperature, which is an important thermodynamic property in MD simu-
lations. An equilibration simulation is performed until the system reaches an
equilibrium state, which is a global minimum on the energy surface. During
this process it is important that the system has enough time and energy
(i.e., temperature) to escape local minima. The system can in some cases any-
way remain kinetically trapped in a local minimum. Annealing simulations
start at higher temperature to overcome such energy barriers and gradually
Computational Biochemistry—Enzyme Mechanisms Explored 93
A B
Fig. 3 (A) Reaction coordinate for interconversion between two stable states is thermo-
dynamically characterized by reaction free energy ΔG° and kinetically by the free acti-
vation energy ΔG6¼ . (B) Marcus model of two parabolas representing the two stable
electronic configurations of the system. Reaction free energy ΔG° and the reorganiza-
tion energy λ used for calculation of the free activation energy ΔG6¼ of the electron
transfer are highlighted.
barrier between two stable states of the system (Fig. 3A). The reaction barrier
can be related to the reaction rate by the Arrhenius law:
kðT Þ ¼ A exp βΔG6¼ (9)
where ΔG6¼ is the free activation energy, β is 1 =kB T (kB is the Boltzmann
constant, T is the absolute temperature), and A is a preexponential factor.
Eq. (9) is further adopted to theoretically describe rates of different processes
involved in the enzyme function.
The first category are long-range electron transfer reactions, which can
be regarded as nonadiabatic processes. A Marcus model of two harmonic
potentials representing an initial and a final electronic configuration
(Fig. 3B) can be used to describe the reaction coordinate. The free activation
energy can be calculated from reaction free energy ΔG° and reorganization
energy λ, which is the energy needed to perform the structural changes
within the system:
ðΔG° + λÞ2
ΔG6¼ ¼ (10)
4λ
The reaction rate constant (9) can then be adopted for electron transfers
using Marcus theory for the activation energy and Fermi’s golden rule
(Marcus & Sutin, 1985) for the preexponential factor:
!
2π 2 1 ðΔG° + λÞ2
kET ðT Þ ¼ HDA pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi exp β (11)
ℏ 4πβ1 λ 4λ
Computational Biochemistry—Enzyme Mechanisms Explored 95
High
Pro
Energy
Energy
In In
Sub
Pro
Sub
Low
Reaction coordinate
Fig. 4 Reaction path can be imagined as a trail through an energy landscape where
basins represent the stable states and the passes represent the transition states. The
path from substrate (Sub) to product (Pro) usually involves several intermediate (In)
states. The energy change plotted against the position on the path is referred to as reac-
tion profile.
from a minimum along the shallowest ascent toward the transition state and
from there along the deepest descent toward the product state. Most of the
chemical reactions are composed of several elementary steps forming
together the minimum energy path (MEP) from reactant to product (Fig. 4).
Many methods have been developed to identify reaction paths and to
distinguish among mechanism variants based on corresponding energy
profiles. In general, one has to first identify the mechanistic options and then
get corresponding accurate energy profiles. At the beginning of a computa-
tional kinetic study, optimization techniques are usually used to find minima
and saddle points on the PES, which together define MEPs. The obtained
energy profile corresponds to the enthalpic part of the free energy (Eq. 6). If
the enthalpic differences among mechanism variants are big, one can already
make qualitative conclusions about the preferred mechanism. The entropic
part of Eq. (6) has to be considered to obtain proper reaction free energy
profile at finite temperature in general, although the effect on the reaction
barrier height is not always that profound (Kazemi, Himo, & Åqvist, 2016).
The entropy-corrected PES is referred to as free energy surface at certain
temperature. To get a quantitative free energy surface, one needs to perform
extensive sampling along the reaction path, which can be extremely costly or
even prohibitive in complex structural models. Therefore, the sampling is
usually performed only when the most probable mechanism is identified
on PES.
In the next part, we first discuss methods to identify a reaction path on
PES. We then turn to the sampling methods that aim to get full free energy
reaction profiles. In the end, we return to the preexponential transmission
coefficient of Eq. (12) and discuss the role of nuclear quantum effects.
Computational Biochemistry—Enzyme Mechanisms Explored 97
preferred. The entropic aspects can also be accessed a posteriori once the
PES path is established as we shall see below.
E, Ren, & Vanden-Eijnden, 2005). Note that the single-ended string method
described above is a variant of the growing string method with even less
initial bias.
The above described COS methods do not a priori search for the tran-
sition state structures. Although CI-NEB makes a step in this direction, it is
not actually guaranteed that it will reach a first-order saddle point. One can,
of course, employ the above described eigenvector following approaches to
refine the maxima of a COS path to saddle points. However, these methods
are both computationally costly and also unreliable once the input structure
is not close to the actual saddle point. An elegant solution is the conju-
gate peak refinement (CPR) (Fischer & Karplus, 1992; Gisdon, Culka, &
Ullmann, 2016) method that gradually constructs chain of states between
the initial and final states while it aims at locating the first-order saddle
points. In contrast to NEB or SM, the number of states along the path is
not fixed and thus the sampling in the saddle point region can be increased
to facilitate its proper location. The CPR method is based on the fact that in
the vicinity of a saddle point, there is one direction, which points to an ener-
getic maximum, while all others lead to a minimum. The CPR algorithm
picks the highest energy structure along the discretized path and performs a
line maximization along the corresponding tangential path vector. This
corrected maximum is then minimized in conjugate space similar to the con-
jugated gradient minimization method. By staying conjugate to the original
path, falling to the neighboring minimum is prevented. The optimized
structure is added into the chain of states. Gradually the saddle point is
approached by providing more sampling in the transition region and occa-
sionally the conjugate optimization procedure can converge to locate the
first-order saddle point. A successful run ends, when all maxima along the
reaction path between the initial and final state are identified and optimized
to first-order saddle points.
entropic part, for which extensive sampling at finite temperature along the
reaction path is needed. In case of US and metadynamics the sampling is
already included in the path search procedure, so entropic influence is
explicitly included and thus free energy profile can be directly estimated.
For US, weighted histogram analysis method (WHAM) (Kumar,
Rosenberg, Bouzida, Swendsen, & Kollman, 1992) can be used to remove
the umbrella bias and to integrate the simulation windows into a free energy
profile. In case of metadynamics, the sum of added Gaussian hills plotted
against the collective variable directly represents the free energy profile.
TPS collected pathways are dynamic trajectories, thus kinetic information,
such as rate constants, can be extracted (Dellago, 2007). But since only tra-
jectories are considered that connect certain regions in configuration space,
the configurations are not distributed according to the equilibrium distribu-
tion of the system. To determine free energy profiles, one has to obtain
equilibrium-distributed configurations. One possibility is to apply biasing
procedures, such as US variants, to rarely visited states, and divide the
CV into overlapping windows (Dellago, 2007).
In case of static methods such as adiabatic PES mapping or chain-of-states
methods, the entropic aspect has to be added by sampling method (MD or
Monte Carlo) to account for entropic part of the free energy profile. One
option is to use these static paths as an input to above described path-based
approaches. A possible obstacle can be that achieving extensive sampling
when high-level QM methods are used to treat the active site may be com-
putationally prohibitive. In the same time, usage of, e.g., SE QM models can
introduce additional bias into the results. A way out of this dilemma offers
the free energy perturbation (FEP) methods. For instance, one can find the
reaction path by a chain-of-states method within QM/MM model (K€astner,
Senn, Thiel, Otte, & Thiel, 2006). Subsequently, an MD simulation is per-
formed for every state of the PES reaction path with QM region kept frozen.
Perturbation energy is calculated as energy for moving one step forward in
the PES QM reaction path while staying in the same MM conformational
ensemble:
where i and i + 1 are the indices of adjacent path steps and r are the coor-
dinate vectors. Free energy for the i ! i + 1 step is then calculated by
averaging over the MM ensemble at step i using Zwanzig equation
(Zwanzig, 1954):
Computational Biochemistry—Enzyme Mechanisms Explored 103
1
ΔGi!i + 1 ¼ ln h exp ðβΔEpert Þii (14)
β
specific circumstances and cannot account for all possible effects. Consider-
ing the constantly increasing number of complete genomes and partially
reconstructed metabolisms, it comes more and more important to get a more
realistic view of the metabolic reaction in its context. The challenge for
computational biochemistry today and in the future is to derive enzymatic
parameters from structure models by using methods that we reviewed in this
article. However, it will be required to go beyond such information. The
kinetic parameters can be combined in master equation approaches
(Becker et al., 2007; Bombarda & Ullmann, 2011) or kinetic Monte Carlo
simulations (Till, Becker, Essigke, & Ullmann, 2008) in order to simulate
complete catalytic cycles that are influenced by environmental parameters
such as pH or membrane potential. To carry the approach further, it might
be possible to model the whole cellular context in reaction–diffusion equa-
tions, which may allow in the future to model complex biochemical reac-
tions. Combining structural biology and systems biology may thus be a
promising direction, especially considering the pace in which both fields
make progress in recent years. With the help of computer models that rely
on a solid experimental basis, we may more and more understand how the
jigglings and wigglings of atoms leads to the complex phenomenon we
call life.
REFERENCES
Amara, P., Field, M. J., Alhambra, C., & Gao, J. (2000). The generalized hybrid orbital
method for combined quantum mechanical/molecular mechanical calculations: Formu-
lation and tests of the analytical derivatives. Theoretical Chemistry Accounts, 104, 336–343.
Åqvist, J., & Warshel, A. (1993). Simulation of enzyme reactions using valence bond force
fields and other hybrid quantum/classical approaches. Chemical Reviews, 93, 2523–2544.
Baker, N. A., Sept, D., Joseph, S., Holst, M. J., & McCammon, J. A. (2001). Electrostatics of
nanosystems: Application to microtubules and the ribosome. Proceedings of the National
Academy of Sciences, 98, 10037–10041.
Bakowies, D., & Thiel, W. (1991). MNDO study of large carbon clusters. Journal of the
American Chemical Society, 113, 3704–3714.
Bartesaghi, A., Merk, A., Banerjee, S., Matthies, D., Wu, X., Milne, J. L. S., &
Subramaniam, S. (2015). 2.2 Å resolution cryo-EM structure of Î2-galactosidase in com-
plex with a cell-permeant inhibitor. Science, 348, 1147–1151.
Becke, A. (1993). Density functional thermochemistry. III. The role of exact exchange. The
Journal of Chemical Physics, 98, 5648–5652.
Becker, T., Ullmann, R. T., & Ullmann, G. M. (2007). Simulation of the electron transfer
between the tetraheme subunit and the special pair of the photosynthetic reaction center
using a microstate description. The Journal of Physical Chemistry. B, 111, 2957–2968.
Bernardi, R. C., Melo, M. C. R., & Schulten, K. (2015). Enhanced sampling techniques in
molecular dynamics simulations of biological systems. Biochimica et Biophysica Acta, 1850,
872–877.
106 Martin Culka et al.
Bolhuis, P. G., Chandler, D., Dellago, C., & Geissler, P. L. (2002). Transition path sampling:
Throwing ropes over rough mountain passes, in the dark. Annual Review of Physical
Chemistry, 53, 291–318.
Bolhuis, P. G., Dellago, C., & Chandler, D. (1998). Sampling ensembles of deterministic
transition pathways. Faraday Discussions, 110, 421–436.
Bombarda, E., Becker, T., & Ullmann, G. M. (2006). The influence of the membrane poten-
tial on the protonation of bacteriorhodopsin: Insights from electrostatic calculations into
the regulation of proton pumping. Journal of the American Chemical Society, 128,
12129–12139.
Bombarda, E., & Ullmann, G. M. (2011). Continuum electrostatic investigations of charge
transfer processes in biological molecules using a microstate description. Faraday
Discussions, 148, 173–193.
Braams, B. J., & Manolopoulos, D. E. (2006). On the short-time limit of ring polymer molec-
ular dynamics. The Journal of Chemical Physics, 125, 124105.
Branduardi, D., Gervasio, F. L., & Parrinello, M. (2007). From A to B in free energy space.
The Journal of Chemical Physics, 126, 054103. http://dx.doi.org/10.1063/1.2432340.
Bryantsev, V. S., Diallo, M. S., Van Duin, A. C. T., & Goddard, W. A. (2009). Evaluation of
B3LYP, X3LYP, and M06-class density functionals for predicting the binding energies of
neutral, protonated, and deprotonated water clusters. Journal of Chemical Theory and
Computation, 5, 1016–1026.
Calimet, N., & Ullmann, G. M. (2004). The influence of a transmembrane pH gradient on
protonation probabilities of bacteriorhodopsin: The structural basis of the back-pressure
effect. Journal of Molecular Biology, 339, 571–589.
Cao, J., & Voth, G. A. (1994a). The formulation of quantum statistical mechanics based on
the Feynman path centroid density. I. Equilibrium properties. The Journal of Chemical
Physics, 100, 5093–5105.
Cao, J., & Voth, G. A. (1994b). The formulation of quantum statistical mechanics based on
the Feynman path centroid density. II. Dynamical properties. The Journal of Chemical
Physics, 100, 5106–5117.
Cerjan, C. J. (1981). On finding transition states. The Journal of Chemical Physics, 75, 2800.
Chen, J. L., Noodleman, L., Case, D., & Bashford, D. (1994). Incorporating solvation effects
into density functional electronic structure calculations. The Journal of Physical Chemistry,
98, 11059–11068.
Cı́žek, J. (1966). On the correlation problem in atomic and molecular systems. Calculation of
wavefunction components in Ursell-type expansion using quantum-field theoretical
methods. The Journal of Chemical Physics, 45, 4256–4266.
Cole, D. J., & Hine, N. D. M. (2016). Applications of large-scale density functional theory in
biology. Journal of Physics. Condensed Matter, 28, 393001.
Cornell, W. D., Cieplak, P., Bayly, C. I., Gould, I. R., Merz, K. M., Ferguson, D. M., …
Kollman, P. A. (1996). A second generation force field for the simulation of proteins,
nucleic acids, and organic molecules. Journal of the American Chemical Society, 118,
2309–2309.
Costa, M. G. S., Batista, P. R., Bisch, P. M., & Perahia, D. (2015). Exploring free energy
landscapes of large conformational changes: Molecular dynamics with excited normal
modes. Journal of Chemical Theory and Computation, 11, 2755–2767.
Craig, I. R., & Manolopoulos, D. E. (2004). Quantum statistics and classical mechanics: Real
time correlation functions from ring polymer molecular dynamics. The Journal of Chemical
Physics, 121, 3368–3373.
Cramer, C. J. (2004). Essentials of computational chemistry: Theories and models (2nd ed.).
Chichester: Wiley.
Dellago, C. (2007). Transition path sampling and the calculation of free energies. In C. Chipot
& A. Pohorille (Eds.), Free energy calculations (pp. 249–276). Berlin, Heidelberg: Springer.
Computational Biochemistry—Enzyme Mechanisms Explored 107
Dellago, C., & Bolhuis, P. G. (2007). Transition path sampling simulations of biological sys-
tems. In M. Reiher (Ed.), Atomistic approaches in modern biology (Vol. 268, pp. 291–317).
Berlin, Heidelberg: Springer.
Dellago, C., Bolhuis, P. G., Csajka, F. S., & Chandler, D. (1998). Transition path sampling
and the calculation of rate constants. The Journal of Chemical Physics, 108, 1964.
E, W., Ren, W., & Vanden-Eijnden, E. (2002). String method for the study of rare events.
Physical Review B, 66, 052301.
E, W., Ren, W., & Vanden-Eijnden, E. (2005). Finite temperature string method for the
study of rare events. The Journal of Physical Chemistry B, 109, 6688–6693.
Elstner, M. (2006). The SCC-DFTB method and its application to biological systems. The-
oretical Chemistry Accounts, 116, 316–325.
Feliks, M., & Ullmann, G. M. (2012). Glycerol dehydratation by the B12-independent
enzyme may not involve the migration of a hydroxyl group: A computational study.
The Journal of Physical Chemistry. B, 116, 7076–7087.
Feynman, R. (1964). The Feynman lectures on physics. New York, USA: Basic Books.
Retrieved from http://www.feynmanlectures.caltech.edu/.
Field, M. J. (2007). A practical introduction to the simulation of molecular systems (2nd ed.).
Cambridge, UK: Cambridge University Press.
Field, M. J., Bash, P. A., & Karplus, M. (1990). A combined quantum mechanical and molec-
ular mechanical potential for molecular dynamics simulations. Journal of Computational
Chemistry, 11, 700–733.
Fischer, S., & Karplus, M. (1992). Conjugate peak refinement: An algorithm for finding reac-
tion paths and accurate transition states in systems with many degrees of freedom. Chem-
ical Physics Letters, 194, 252–261.
Gamow, G. (1928). Zur quantentheorie des atomkernes. Zeitschrift fr Physiotherapie, 51,
204–212.
Gao, J., Ma, S., Major, D. T., Nam, K., Pu, J., & Truhlar, D. G. (2006). Mechanisms and free
energies of enzymatic reactions. Chemical Reviews, 106, 3188–3209.
Garcia-Viloca, M., Alhambra, C., Truhlar, D. G., & Gao, J. (2001). Inclusion of quantum-
mechanical vibrational energy in reactive potentials of mean force. The Journal of Chemical
Physics, 114, 9953–9958.
Georgieva, P., & Himo, F. (2010). Quantum chemical modeling of enzymatic reactions: The
case of histone lysine methyltransferase. Journal of Computational Chemistry, 31,
1707–1714.
Gilson, M. K., & Zhou, H.-X. (2007). Calculation of protein-ligand binding affinities.
Annual Review of Biophysics and Biomolecular Structure, 36, 21–42.
Gisdon, F. J., Culka, M., & Ullmann, G. M. (2016). PyCPR—A python-based implemen-
tation of the Conjugate Peak Refinement (CPR) algorithm for finding transition state
structures. Journal of Molecular Modeling, 22, 242.
Golkaram, M., Shin, Y. K., & van Duin, A. C. T. (2014). Reactive molecular dynamics study
of the pH-dependent dynamic structure of α-helix. The Journal of Physical Chemistry. B,
118, 13498–13504.
Gray, H. B., & Winkler, J. R. (2005). Long-range electron transfer. Proceedings of the National
Academy of Sciences, 102, 3534–3539.
Head-Gordon, M., Pople, J. A., & Frisch, M. J. (1988). MP2 energy evaluation by direct
methods. Chemical Physics Letters, 153, 503–506.
Henkelman, G., Uberuaga, B. P., & Jónsson, H. (2000). A climbing image nudged elastic
band method for finding saddle points and minimum energy paths. The Journal of Chemical
Physics, 113, 9901.
Heyden, A., & Truhlar, D. G. (2008). Conservative algorithm for an adaptive change of res-
olution in mixed atomistic/coarse-grained multiscale simulations. Journal of Chemical
Theory and Computation, 4, 217–221.
108 Martin Culka et al.
Honig, B., & Nicholls, A. (1995). Classical electrostatics in biology and chemistry. Science,
268, 1144–1149.
Hwang, J. K., & Warshel, A. (1993). A quantized classical path approach for calculations of
quantum mechanical rate constants. The Journal of Physical Chemistry, 97, 10053–10058.
Ingólfsson, H. I., Lopez, C. A., Uusitalo, J. J., de Jong, D. H., Gopal, S. M., Periole, X., &
Marrink, S. J. (2014). The power of coarse graining in biomolecular simulations. Wiley
Interdisciplinary Reviews: Computational Molecular Science, 4, 225–248.
Jonsson, H., Mills, G., & Jacobsen, K. W. (1998). Nudged elastic band method for finding
minimum energy paths of transitions. In B. J. Berne, G. Ciccotti, & D. F. Coker (Eds.),
Classical and quantum dynamics in condensed phase simulations (pp. 385–404). Singapore:
World Scientific.
Jorgensen, W. L., Maxwell, D. S., & Tirado-Rives, J. (1996). Development and testing of the
OPLS all-atom force field on conformational energetics and properties of organic liquids.
Journal of the American Chemical Society, 118, 11225–11236.
Jorgensen, W. L., & Tirado-Rives, J. (1996). Monte Carlo vs molecular dynamics for con-
formational sampling. The Journal of Physical Chemistry, 100, 14508–14513.
Kamerlin, S. C. L., & Warshel, A. (2010). The EVB as a quantitative tool for formulating
simulations and analyzing biological and chemical reactions. Faraday Discussions, 145,
71–106.
Kamerlin, S. C. L., & Warshel, A. (2011). The empirical valence bond model: Theory and
applications. Wiley Interdisciplinary Reviews: Computational Molecular Science, 1, 30–45.
Kar, P., & Feig, M. (2014). Recent advances in transferable coarse-grained modeling of pro-
teins. Advances in Protein Chemistry and Structural Biology, 96, 143–180.
K€astner, J., Senn, H. M., Thiel, S., Otte, N., & Thiel, W. (2006). QM/MM free-energy
perturbation compared to thermodynamic integration and umbrella sampling: Applica-
tion to an enzymatic reaction. Journal of Chemical Theory and Computation, 2, 452–461.
Kazemi, M., Himo, F., & Åqvist, J. (2016). Enzyme catalysis by entropy without Circe effect.
Proceedings of the National Academy of Sciences of the United States of America, 113,
2406–2411.
Klamt, A., & Sch€ €rmann, G. (1993). COSMO: A new approach to dielectric screening in
uu
solvents with explicit expressions for the screening energy and its gradient. Journal of the
Chemical Society, Perkin Transactions 2 (799–805).
Kmiecik, S., Gront, D., Kolinski, M., Wieteska, L., Dawid, A. E., & Kolinski, A. (2016).
Coarse-grained protein models and their applications. Chemical Reviews, 116,
7898–7936.
K€uhlbrandt, W. (2014). Cryo-EM enters a new era. eLife, 3, e03678.
Kulik, H. J., Luehr, N., Ufimtsev, I. S., & Martı́nez, T. J. (2012). Ab initio quantum chem-
istry for protein structures. The Journal of Physical Chemistry. B, 116, 12501–12509.
Kumar, S., Rosenberg, J. M., Bouzida, D., Swendsen, R. H., & Kollman, P. A. (1992). The
weighted histogram analysis method for free-energy calculations on biomolecules. I. The
method. Journal of Computational Chemistry, 13, 1011–1021.
Laio, A., & Parrinello, M. (2002). Escaping free-energy minima. Proceedings of the National
Academy of Sciences of the United States of America, 99, 12562–12566.
Leach, A. R. (2001). Molecular modelling: Principles and applications (2nd ed.). New Jersey:
Prentice Hall.
Levitt, M., & Warshel, A. (1975). Computer simulation of protein folding. Nature, 253,
694–698.
Li, J., Nelson, M. R., Peng, C. Y., Bashford, D., & Noodleman, L. (1998). Incorporating
protein environments in density functional theory: A self-consistent reaction field calcu-
lation of redox potentials of [2Fe2S] clusters in ferredoxin and phthalate dioxygenase
reductase. The Journal of Physical Chemistry. B, 102, 6311–6324.
Computational Biochemistry—Enzyme Mechanisms Explored 109
Li, J., & Ryde, U. (2014). Comparison of the active-site design of molybdenum oxo-transfer
enzymes by quantum mechanical calculations. Inorganic Chemistry, 53, 11913–11924.
Liang, T., Shin, Y. K., Cheng, Y.-T., Yilmaz, D. E., Vishnu, K. G., Verners, O., … van
Duin, A. C. (2013). Reactive potentials for advanced atomistic simulations. Annual
Review of Materials Research, 43, 109–129.
Liu, T., Han, W.-G., Himo, F., Ullmann, G. M., Bashford, D., Toutchkine, A., …
Noodleman, L. (2004). Density functional vertical self-consistent reaction field theory
for solvatochromism studies of solvent-sensitive dyes. The Journal of Physical Chemistry.
B, 108, 11157–11169.
MacKerell, A. D., Bashford, D., Bellott, M., Dunbrack, R. L., Evanseck, J. D., Field, M. J.,
… Karplus, M. (1998). All-atom empirical potential for molecular modeling and dynam-
ics studies of proteins. The Journal of Physical Chemistry. B, 102, 3586–3616.
Manta, B., Raushel, F. M., & Himo, F. (2014). Reaction mechanism of zinc-dependent
cytosine deaminase from Escherichia coli: A quantum-chemical study. The Journal of Phys-
ical Chemistry. B, 118, 5644–5652.
Marcus, R., & Sutin, N. (1985). Electron transfers in chemistry and biology. Biochimica et
Biophysica Acta, Reviews on Bioenergetics, 811, 265–322.
Maseras, F., & Morokuma, K. (1995). IMOMM: A new integrated ab initio + molecular
mechanics geometry optimization scheme of equilibrium structures and transition states.
Journal of Computational Chemistry, 16, 1170–1179.
Meier, K., Choutko, A., Dolenc, J., Eichenberger, A. P., Riniker, S., & Van
Gunsteren, W. F. (2013). Multi-resolution simulation of biomolecular systems:
A review of methodological issues. Angewandte Chemie, International Edition, 52,
2820–2834.
Møller, C., & Plesset, M. S. (1934). Note on an approximation treatment for many-electron
systems. Physics Review, 46, 618–622.
Monti, S., Corozzi, A., Fristrup, P., Joshi, K. L., Shin, Y. K., Oelschlaeger, P., …
Bourne, P. E. (2013). Exploring the conformational and reactive dynamics of biomol-
ecules in solution using an extended version of the glycine reactive force field. Physical
Chemistry Chemical Physics, 15, 15062.
Moser, C. C., Keske, J. M., Warncke, K., Farid, R. S., & Dutton, P. L. (1992). Nature of
biological electron transfer. Nature, 355, 796–802.
Munro, L. J., & Wales, D. J. (1999). Defect migration in crystalline silicon. Physical Review B,
59, 3969–3980.
Nango, E., Royant, A., Kubo, M., Nakane, T., Wickstrand, C., Kimura, T., … Iwata, S.
(2016). A three-dimensional movie of structural changes in bacteriorhodopsin. Science,
354, 1552–1557.
Noid, W. G. (2013). Perspective: Coarse-grained models for biomolecular systems. The Jour-
nal of Chemical Physics, 139, 090901.
Oostenbrink, C., Villa, A., Mark, A. E., & Van Gunsteren, W. F. (2004). A biomolecular
force field based on the free enthalpy of hydration and solvation: The GROMOS
force-field parameter sets 53A5 and 53A6. Journal of Computational Chemistry, 25,
1656–1676.
Page, C. C., Moser, C. C., Chen, X., & Dutton, P. L. (1999). Natural engineering principles
of electron tunneling in biological oxidation-reduction. Nature, 402, 47–52.
Pande, K., Hutchison, C. D. M., Groenhof, G., Aquila, A., Robinson, J. S., Tenboer, J., …
Schmidt, M. (2016). Femtosecond structural dynamics drives the trans/cis isomerization
in photoactive yellow protein. Science, 352, 725–729.
Paquet, E., & Viktor, H. L. (2015). Molecular dynamics, Monte Carlo simulations, and
Langevin dynamics: A computational review. BioMed Research International, 2015,
183918.
110 Martin Culka et al.
Peters, B., Heyden, A., Bell, A. T., & Chakraborty, A. (2004). A growing string method for
determining transition states: Comparison to the nudged elastic band and string methods.
The Journal of Chemical Physics, 120, 7877–7886.
Riniker, S., Allison, J. R., & van Gunsteren, W. F. (2012). On developing coarse-grained
models for biomolecular simulation: A review. Physical Chemistry Chemical Physics, 14,
12423.
Rowley, C. N., & Woo, T. K. (2009). New shooting algorithms for transition path sampling:
Centering moves and varied-perturbation sizes for improved sampling. The Journal of
Chemical Physics, 131, 234102.
R€ucker, P., Wieninger, S. A., Ullmann, G. M., & Sticht, H. (2012). pH-dependent molec-
ular dynamics of vesicular stomatitis virus glycoprotein G. Proteins, 80, 2601–2613.
Saunders, M. G., & Voth, G. A. (2013). Coarse-graining methods for computational biology.
Annual Review of Biophysics, 42, 73–93.
Senftle, T. P., Hong, S., Islam, M. M., Kylasa, S. B., Zheng, Y., Shin, Y. K., … van
Duin, A. C. T. (2016). The ReaxFF reactive force-field: Development, applications
and future directions. Computational Materials, 2, 15011.
Senn, H. M., & Thiel, W. (2007). QM/MM methods for biological systems. In M. Reiher
(Ed.), Atomistic approaches in modern biology (Vol. 268, pp. 173–290). Berlin, Heidelberg:
Springer.
Senn, H. M., & Thiel, W. (2009). QM/MM methods for biomolecular systems. Angewandte
Chemie, International Edition, 48, 1198–1229.
Sharp, K. E. (1998). Calculation of electron transfer reorganization energies using the Finite
Difference Poisson-Boltzmann model. Biophysical Journal, 73, 1241–1250.
Shen, L., & Hu, H. (2014). Resolution-adapted all-atomic and coarse-grained model for bio-
molecular simulations. Journal of Chemical Theory and Computation, 10, 2528–2536.
Shurki, A., Derat, E., Barrozo, A., & Kamerlin, S. C. L. (2015). How valence bond theory
can help you understand your (bio) chemical reaction. Chemical Society Reviews, 44,
1037–1052.
Singh, U. C., & Kollman, P. A. (1986). A combined ab initio quantum mechanical and
molecular mechanical method for carrying out simulations on complex molecular sys-
tems: Applications to the CH3Cl + Cl exchange reaction and gas phase protonation
of polyethers. Journal of Computational Chemistry, 7, 718–730.
Sokkar, P., Boulanger, E., Thiel, W., & Sanchez-Garcia, E. (2015). Hybrid quantum
mechanics/molecular mechanics/coarse grained modeling: A triple-resolution approach
for biomolecular systems. Journal of Chemical Theory and Computation, 11, 1809–1818.
Stewart, J. J. P. (2013). Optimization of parameters for semiempirical methods VI: More
modifications to the NDDO approximations and re-optimization of parameters. Journal
of Molecular Modeling, 19, 1–32.
Svensson, M., Humbel, S., Froese, R. D. J., Matsubara, T., Sieber, S., & Morokuma, K.
(1996). ONIOM: A multilayered integrated MO + MM method for geometry
optimizations and single point energy predictions. A test for Diels–Alder reactions
and Pt(P(t-Bu)3)2 + H2 oxidative addition. The Journal of Physical Chemistry, 100,
19357–19363.
Swenson, D. W. H., & Bolhuis, P. G. (2014). A replica exchange transition interface sam-
pling method with multiple interface sets for investigating networks of rare events. The
Journal of Chemical Physics, 141, 044101.
Thery, V., Rinaldi, D., Rivail, J.-L., Maigret, B., & Ferenczy, G. G. (1994). Quantum
mechanical computations on very large molecular systems: The local self-consistent field
method. Journal of Computational Chemistry, 15, 269–282.
Thompson, M. A., & Schenter, G. K. (1995). Excited ates of the bacteriochlorophyll b dimer
of Rhodopseudomonas viridis: A QM/MM study of the photosynthetic reaction center
that includes MM polarization. The Journal of Physical Chemistry, 99, 6374–6386.
Computational Biochemistry—Enzyme Mechanisms Explored 111
Till, M. S., Becker, T., Essigke, T., & Ullmann, G. M. (2008). Simulating the proton transfer
in Gramicidin A by a sequential dynamical Monte Carlo method. The Journal of Physical
Chemistry. B, 112, 13401–13410.
Todorovic, M., Bowler, D. R., Gillan, M. J., & Miyazaki, T. (2013). Density-functional the-
ory study of gramicidin A ion channel geometry and electronic properties. Journal of The
Royal Society Interface, 10, 20130547.
Torrie, G., & Valleau, J. (1977). Nonphysical sampling distributions in Monte Carlo free-
energy estimation: Umbrella sampling. Journal of Computational Chemistry, 23, 187–199.
Tozzini, V. (2005). Coarse-grained models for proteins. Current Opinion in Structural Biology,
15, 144–150.
Truhlar, D. G., Gao, J., Alhambra, C., Garcia-Viloca, M., Corchado, J., Sánchez, M. L., &
Villà, J. (2002). The incorporation of quantum effects in enzyme kinetics modeling.
Accounts of Chemical Research, 35, 341–349.
Ullmann, G. M. (2000). The coupling of protonation and reduction in proteins with multiple
redox centers: Theory, computational method, and application to cytochrome c3. The
Journal of Physical Chemistry. B, 104, 6293–6301.
Ullmann, G. M., & Bombarda, E. (2014). Continuum electrostatic analysis of proteins.
In G. Naray-Szabo (Ed.), Protein modelling (pp. 135–163). Cham, Switzerland:
Springer International Publishing.
Ullmann, G. M., Kloppmann, E., Essigke, T., Krammer, E.-M., Klingen, A. R., Becker, T.,
& Bombarda, E. (2008). Investigating the mechanisms of photosynthetic proteins using
continuum electrostatics. Photosynthesis Research, 97, 33–53.
Ullmann, G. M., Mueller, L., & Bombarda, E. (2016). Theoretical analysis of
electron transfer in proteins: From simple proteins to complex machineries. In
W. Cramer (Ed.), Cytochrome complexes: Evolution, structures, energy transduction, and
signaling (Vol. 37, pp. 99–127). Dordrecht, Netherlands: Springer International
Publishing.
Ullmann, R. T., & Ullmann, G. M. (2012). GMCT: A Monte Carlo simulation package for
macromolecular receptors. Journal of Computational Chemistry, 33, 887–900.
Vanden-Eijnden, E., & Venturoli, M. (2009). Revisiting the finite temperature string
method for the calculation of reaction tubes and free energies. The Journal of Chemical
Physics, 130, 194103.
van Duin, A. C. T., Dasgupta, S., Lorant, F., & Goddard, W. A., III. (2001). ReaxFF:
A reactive force field for hydrocarbons. The Journal of Physical Chemistry. B, 105,
9396–9409.
van Mourik, T., B€ uhl, M., & Gaigeot, M.-P. (2014). Density functional theory across chem-
istry, physics and biology. Philosophical Transactions. Series A, Mathematical, Physical, and
Engineering Sciences, 372, 20120488.
Verlackt, C. C. W., Neyts, E. C., Jacob, T., Fantauzzi, D., Golkaram, M., Shin, Y.-K., …
Bogaerts, A. (2015). Atomic-scale insight into the interactions between hydroxyl radicals
and DNA in solution using the ReaxFF reactive force field. New Journal of Physics, 17,
103005.
Řezáč, J., & Hobza, P. (2012). Advanced corrections of hydrogen bonding and dispersion for
semiempirical quantum mechanical methods. Journal of Chemical Theory and Computation,
8, 141–151.
Wales, D. J., & Walsh, T. R. (1998). Theoretical study of the water pentamer. American Insti-
tute of Physics, 105, 6957–6971.
Wang, Z., Antoniou, D., Schwartz, S. D., & Schramm, V. L. (2016). Hydride transfer in
DHFR by transition path sampling, kinetic isotope effects, and heavy enzyme studies.
Biochemistry, 55, 157–166.
Warshel, A. (1991). Computer modeling of chemical reactions in enzymes and solutions. New York:
Wiley & Sons.
112 Martin Culka et al.