ArticlePDF Available

Comparison of DNA hydration patterns obtained using two distinct computational methods, molecular dynamics simulation and three-dimensional reference interaction site model theory

Authors:

Abstract

Because proteins and DNA interact with each other and with various small molecules in the presence of water molecules, we cannot ignore their hydration when discussing their structural and energetic properties. Although high-resolution crystal structure analyses have given us a view of tightly bound water molecules on their surface, the structural data are still insufficient to capture the detailed configurations of water molecules around the surface of these biomolecules. Thanks to the invention of various computational algorithms, computer simulations can now provide an atomic view of hydration. Here, we describe the apparent patterns of DNA hydration calculated by using two different computational methods: Molecular dynamics (MD) simulation and three-dimensional reference interaction site model (3D-RISM) theory. Both methods are promising for obtaining hydration properties, but until now there have been no thorough comparisons of the calculated three-dimensional distributions of hydrating water. This rigorous comparison showed that MD and 3D-RISM provide essentially similar hydration patterns when there is sufficient sampling time for MD and a sufficient number of conformations to describe molecular flexibility for 3D-RISM. This suggests that these two computational methods can be used to complement one another when evaluating the reliability of the calculated hydration patterns.
Comparison of DNA hydration patterns obtained using two distinct
computational methods, molecular dynamics simulation
and three-dimensional reference interaction site model theory
Yoshiteru Yonetani,1Yutaka Maruyama,2Fumio Hirata,2,aand Hidetoshi Kono1,3,a
1Computational Biology Group, Quantum Beam Science Directorate, Japan Atomic Energy Agency,
8-1 Umemidai, Kizugawa, Kyoto 619-0215, Japan
2Department of Theoretical and Computational Molecular Science, Institute for Molecular Science,
Okazaki 444-8585, Japan
3PRESTO, Japan Science and Technology Agency, 4-1-8 Kawaguchi, Saitama 332-0012, Japan
Received 28 December 2007; accepted 12 March 2008; published online 9 May 2008
Because proteins and DNA interact with each other and with various small molecules in the
presence of water molecules, we cannot ignore their hydration when discussing their structural and
energetic properties. Although high-resolution crystal structure analyses have given us a view of
tightly bound water molecules on their surface, the structural data are still insufficient to capture the
detailed configurations of water molecules around the surface of these biomolecules. Thanks to the
invention of various computational algorithms, computer simulations can now provide an atomic
view of hydration. Here, we describe the apparent patterns of DNA hydration calculated by using
two different computational methods: Molecular dynamics MDsimulation and three-dimensional
reference interaction site model 3D-RISMtheory. Both methods are promising for obtaining
hydration properties, but until now there have been no thorough comparisons of the calculated
three-dimensional distributions of hydrating water. This rigorous comparison showed that MD and
3D-RISM provide essentially similar hydration patterns when there is sufficient sampling time for
MD and a sufficient number of conformations to describe molecular flexibility for 3D-RISM. This
suggests that these two computational methods can be used to complement one another when
evaluating the reliability of the calculated hydration patterns. © 2008 American Institute of Physics.
DOI: 10.1063/1.2904865
I. INTRODUCTION
When discussing the structural and energetic properties
of biological molecules such as proteins and nucleic acids,
one must consider not only the solute biomolecules but also
the solvent or surrounding water molecules.1Indeed, the
stable states of biological molecules are significantly affected
by the solvent conditions. A good example of this is the
structure of DNA, which exhibits structural transition among
different A, B, and Z forms, depending on the solvent and its
ionic strength. Although high-resolution x-ray crystal struc-
ture analyses have provided us with detailed pictures of hy-
dration states,2,3these pictures remain incomplete because i
the positions of hydrogen atoms cannot be determined by
x-ray diffraction; iithe hydrated structure is not viewed
under physiological conditions, i.e., the solution state is not
necessarily obtained through analysis of crystallized
samples, as the effect of crystal packing is not negligible;
and iiithe hydrated structure may be affected by the salt
species and concentration. The positions of hydrogen atoms
can be determined through neutron crystal structure
analysis,2but so far the structures of only 19 biomolecules
have been solved using neutron diffraction as of Nov. 20,
2007.
An alternative approach to the structural analysis of the
hydration of biomolecules is to use computational methods,
which can be complementary to experimental measurements.
Such computational approaches are classified into two cat-
egories. One is many-particles simulation,4such as the mo-
lecular dynamics MDand Monte Carlo methods, in which
molecular configurations are generated according to New-
ton’s equations of motion or stochastic walking, respectively.
Another is a more theoretical approach based on the integral
equations for molecular distribution functions.5A good ex-
ample is reference interaction site model RISMtheory.6
Both the simulation and theoretical approaches are capable
of providing structures of the hydration around a solute mol-
ecule but have distinct features: The simulation can be used
with any complex molecular system because all atomic po-
sitions are numerically derived. This approach has therefore
been widely used to analyze the hydration of proteins,7
nucleic acids,8,9and their complexes.10 One shortcoming of
the simulation approach is the large computational resource
needed to obtain a complete statistical ensemble. If atomic
configurations are not sufficiently sampled, a proper descrip-
tion of the statistical properties will not be achieved. For
example, conventional MD techniques require very pro-
longed typically approximately millisecondssimulations to
estimate the probability of finding water molecules within a
protein cavity completely isolated from the bulk solvent.11
aAuthors to whom correspondence should be addressed. Electronic
addresses: hirata@ims.ac.jp and kono.hidetoshi@jaea.go.jp.
THE JOURNAL OF CHEMICAL PHYSICS 128, 185102 2008
0021-9606/2008/12818/185102/9/$23.00 © 2008 American Institute of Physics128, 185102-1
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
The theoretical approach does not suffer from this sampling
problem because the configuration integral is made over the
entire configuration space of the solvent at equilibrium with
a fixed solute conformation. Consequently, it naturally
samples solvent configurations that correspond to the con-
fined space within a biomolecule cavity. On the other hand,
the theoretical calculation of a fixed solute conformation is
not necessarily sufficient to clarify the physics of DNA hy-
dration because structural flexibility is an essential feature of
most biomolecules and will have non-negligible effects on
the behavior of the surrounding water. In fact, the effect of
structural conformation on the hydration characteristics of
proteins has been investigated in our recent work.12
The purpose of the present work is to compare two rep-
resentative computational approaches, MD simulation and
RISM theory, and to evaluate their applicability for descrip-
tion of a biomolecule’s hydration. Both MD and RISM ap-
pear to have the potential to describe the structural features
of hydration, and the calculation conditions i.e., trajectory
length and solute treatmentrequired to describe the hy-
drated structures of biomolecules by using the two methods
will be clarified through the present work. We chose the
double helix B-form of DNA as a test case. MD simulation
has always been capable of providing three-dimensional
3Ddistributions of molecules, but until about 10 years ago
RISM theory was limited to the calculation of a one-
dimensional 1Ddistribution function e.g., radial distribu-
tion function. Since then, however, extension to the
3D-RISM method enabled RISM to be also applicable for
calculation of 3D distributions.6A DNA-water system is suit-
able for the evaluation of calculated hydrations because
DNA molecules composed of particular sequences are
known to have a well-defined hydration pattern called a
“water spine,”13 which is composed of highly ordered water
molecules aligned along the narrow minor groove. Moreover,
it appears that the surface localized water plays important
roles in various biological and chemical processes. For in-
stance, the affinities with which proteins and drugs bind to
DNA are more or less influenced by entropic contributions
from water molecules released from the DNA surface upon
their binding.14 Surface water also affects the electronic
properties of DNA. The rate of charge transfer along DNA is
known to depend on the DNA sequence, and in this case,
surface water is also thought to be a major determinant of the
electronic properties.
II. METHODS
A. Molecular models
To compare the MD and 3D-RISM methods, we had to
construct a system in which the conditions used for the two
calculations were consistent. So as much as possible, we
used the same system and the same conditions for both cal-
culations. A 12 bare pair bpfragment of B-form DNA
d5CGCGATATCGCG3was prepared as the solute. We
used the force field parameters of AMBER PARM 99 Ref. 15
to simulate the DNA molecule. This parameter set is known
to be suitable for reproducing the structure of B-DNA in MD
simulations.16 For water molecules, we used the TIP3P
model.17 One distinction of the water treatment between the
3D-RISM and MD methods was that we used an additional
Lennard–Jones term with the parameters =192.5 J/mol and
=0.4 Å for the water hydrogen site in order to converge the
RISM calculation.18 K+and Clions were included in the
MD simulation system, whereas they were ignored in the
RISM calculations.
Molecular flexibility is usually considered in MD simu-
lations of biomolecules, but the solute molecule is usually
fixed in the RISM calculations. Consequently, MD and 3D-
RISM calculations give different results, as the former con-
siders solute flexibility and the latter does not. In order to
evaluate the effects of solute treatment and to make an accu-
rate comparison, we carried out four types of calculations
shown in Table I, which included normal MD MDflex,MD
with a position constraint to mimic the solute treatment in
3D-RISM MDrigid, conventional 3D-RISM RISMrigid, and
3D-RISM with consideration of the solute flexibility
RISMflex. We will explain how we performed these calcu-
lations in the following sections.
B. MD simulation
In MD simulations, time-dependent development of the
atoms in a system is derived from the numerical solution of
Newton’s equations of motion, where forces acting on the
atoms come from intra- and intermolecular interactions.4Af-
ter this numerical procedure, statistical properties, such as
the spatial distribution of solvent molecules, are calculated
over the trajectory. To perform our MD simulations i.e.,
MDflex and MDrigid, the DNA-water system was prepared as
follows. We initially generated the B-form structure of the 12
bp DNA d5CGCGATATCGCG3fragment using the
Nucgen module of AMBER7.19 Then, 5114 water molecules
were placed around the DNA, and a periodic boundary con-
dition with a truncated octahedral box about 6060
60 Å3in size was imposed. Counterions consisting of 39
K+and 17 Clalso were added so as to realize both a physi-
ological salt concentration of 0.15Mand electrical neutrality.
Using these systems, we computed the time-dependent
development of the atoms by numerically solving the equa-
tions of motion. Integration of the equations was accom-
plished using the Leapfrog algorithm with a time increment
of 1 fs. At this point, to stabilize the numerical integration,
stretching motions of the covalent bonds involving hydrogen
atoms were removed using the SHAKE constraint.20 When
evaluating the interaction forces, van der Waals components
were truncated at 9 Å. Coulomb components were evaluated
using the particle mesh Ewald21 method to account for their
long-ranged nature, which does not require any truncation.
For MDrigid, we further imposed a positional restraint on the
TABLE I. Methods and solute treatments adopted in the present work.
Method
Solute treatment
Flexible Rigid
MD MDflex MDrigid
3D-RISM RISMflex RISMrigid
185102-2 Yonetani et al. J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
solute DNA atoms with a harmonic function of
500 kcal mol−1 Å−2 to suppress flexibility and mimic the
RISM calculation. The MD simulations were performed us-
ing the AMBER SANDER module.19
A 10 ns trajectory was generated in which the pressure
and temperature were set to 1 atm and 300 K, respectively,
using the weak coupling method.22 We calculated the spatial
distribution of the water molecules using the last 8 ns of
data. In this process, the system space was divided into cubic
boxes with an edge size of 0.5 Å, after which the probability
that water molecules are present within each box was calcu-
lated. Note that before this calculation, we eliminated trans-
lational and rotational displacement of the DNA and ob-
tained a water distribution relative to the solute DNA. This
elimination was carried out by applying the translational and
rotational transformation to every instantaneous configura-
tion, so that the instantaneous DNA structure could be fitted
to a reference DNA structure. We chose the instantaneous
structure that was most similar to the time-averaged structure
as the reference structure.
C. 3D-RISM theory
The 3D-RISM theory6is an integral equation theory
based on statistical mechanics used to obtain the molecular
distribution functions from the intermolecular potential func-
tions and thermodynamic conditions i.e., temperature and
density. This theoretical procedure has two steps. The first
step is to calculate the density pair correlation function in the
aqueous solution based on the 1D-RISM theory, which rep-
resents the microscopic structure of the distribution of sol-
vent molecules. In the second step, we immerse a DNA
molecule into the solvent and calculate the 3D distribution of
solvent atoms around the solute molecule based on the
3D-RISM theory. The following is a brief outline of the 3D-
RISM method.
The 3D-RISM integral equation for the 3D solute-
solvent site total and direct correlation functions, h
uvrand
c
uvr, is written as2325
h
uvr=
c
uvr*
vv +
vc
vv r兲兲,1
where
vv r=
rl
vv is the intramolecular matrix of sol-
vent molecules with site separations l
vv ,
and
are inter-
action sites for solvent molecules denoted with superscripts u
and v, respectively,
vis the solvent number density, and
means convolution in direct space. The radial site-site corre-
lation functions for pure solvent h
vv rwere independently
obtained from the conventional 1D-RISM theory modified
with the dielectrically consistent bridge corrections of
Perkyns and Pettitt.26 In the context of 3D-RISM, this en-
sures a proper macroscopic dielectric constant for the solvent
around the solute.25,27
In addition to 3D-RISM equation Eq. 1, the following
Kovalenko-Hirata KHclosure equations,24,28 which include
corrections for the supercell periodicity artifact in both the
direct and total 3D site correlation functions, c
uvrand
h
uvr, are employed to obtain the correlation functions
g
uvr=
exp
+Q
uvfor
0
1+
+Q
uvfor
0,
2
=−u
uvr
kBT+h
uvrc
uvrQ
uv,
where g
uvr=h
uvr+1 is the 3D solute-solvent site distri-
bution function, the interaction potential u
uvrbetween sol-
vent site
and the whole solute is calculated on the supercell
grid using the Ewald summation method, and
Q
uv=4
VcellkBTq0lim
k0
q
k2
vv k+
vh
vv k兲兲 3
is the shift in the distribution functions due to the supercell
background for the solute with net charge q=
q
ucom-
prised of the partial site charges q
u. The 3D solute-solvent
site direct correlation functions c
uvrcalculated from 3D-
RISM equation Eq. 1with the closure equation Eq. 2
were corrected by subtracting the long-ranged electrostatic
asymptotic of the periodic potential u
uvrsynthesized using
the Ewald summation, and adding back that of the single,
nonperiodic solute simply tabulated on the 3D grid within
the supercell.25,27
In conventional RISM calculations, the positions of the
solute atoms do not change. We refer to such a calculation as
RISMrigid, and the solute DNA structure used was the same
as that used for MDrigid. We calculated RISMflex, which took
into account the flexibility of DNA, by repeating the 3D-
RISM calculation for every instantaneous configuration of
the solute, which were obtained from the MDflex calculation,
and then averaging the resultant water distributions. These
calculations were performed under ambient thermodynamic
conditions of 300 K and 0.997 g/cm3, in accordance with
the MD condition.
III. RESULTS AND DISCUSSION
A. Checking the required conditions
for the comparison of MD and 3D-RISM
Before comparing the MD and 3D-RISM results, we de-
termined whether or not the quality of each result was satis-
factory. A major determinant of the quality of the MD results
was the length of the trajectory. If the possible atomic con-
figurations were not sufficiently sampled in the simulation,
we would not obtain a reliable molecular distribution. To
examine the trajectory length dependence of the spatial dis-
tribution of water oxygen, sampling times of 1, 2, 4, and 8 ns
were compared. In Fig. 1, regions colored in blue have a
water oxygen distribution 2.25 times denser than that of the
bulk water. At trajectory lengths below 4 ns, the oxygen dis-
tribution varied with the trajectory length. By contrast, the
distribution showed no remarkable changes between 4 and
8 ns, which means a trajectory of 8 ns is long enough to
get a reliable water distribution.
The 3D-RISM method is not subject to sampling prob-
lems because the 3D-RISM-derived distribution function is a
statistical average obtained from the integral over all the sol-
vation configurations. A problem does occur, however, when
185102-3 DNA hydration patterns J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
considering the solute flexibility in the RISM calculation. In
conventional RISM calculations, solute atoms are fixed, so
the solute flexibility is not considered. To consider the effect
of solute flexibility, in the present work, we performed nu-
merous 3D-RISM calculations for different DNA conforma-
tions obtained from MDflex and averaged the results. To see
the dependence of the results on the number of conforma-
tions used, the spatial distributions of water oxygen were
calculated for four sets of DNA conformations with different
sizes 10, 50, 1200, and 2400兲共see Fig. 2. Each set was
prepared by randomly choosing the indicated number of
snapshots from the MDflex trajectory. As shown in Fig. 2, the
results obtained using 1200 and 2400 conformations were in
good agreement, while that obtained with the set of 10 or 50
conformations largely differed from the others, suggesting
that 1200 solute conformations are sufficient for realization
of a 3D-RISM calculation that considers solute flexibility.
On the basis of the above examination, we compared
3D-RISM and MD by using a trajectory length of 8 ns for
MDflex and 2400 conformations for RISMflex.
B. Comparison of the MD and 3D-RISM results
In our comparison of MD and 3D-RISM, we first com-
pared two cases of MDflex and RISMrigid because these two
calculations treated the solute DNA in the conventional man-
ner for each method. Figure 3shows the distributions of
water oxygen obtained using different threshold density val-
ues, which are multiples of the density of the bulk water. We
found that RISMrigid produces a much larger solvent distri-
bution than MDflex Figs. 3aand 3d. At a threshold of
2.25, for example, the water is distributed along the minor
groove when calculated using MDflex, but is widely distrib-
uted over the entire DNA molecule when calculated using
RISMrigid. Upon increasing the threshold to 4.5, the hydra-
tion pattern seen with RISMrigid becomes comparable to that
of MDflex with a threshold of 2.25. The same tendency was
confirmed by viewing the solvent distribution in a slice plane
parallel to the DNA base pairs Fig. 4. With MDflex
Fig. 4a, regions of high density are concentrated at the
groove surface, whereas with RISMrigid Fig. 4d, they are
widely distributed and are present around the phosphates of
the DNA backbone as well as around the groove surface.
The difference in the apparent hydration patterns de-
scribed above mainly reflects the difference in the treatment
of the solute flexibility, not a difference in the methodologi-
cal details. We confirmed this statement by comparing the
results obtained when the calculations were made with con-
sistent solute treatment. MDflex Figs. 3aand Fig. 4aand
RISMflex Figs. 3cand 4c, which both consider the solute
flexibility, yielded similar patterns of water density. Like-
wise, MDrigid Figs. 3band 4band RISMrigid Figs. 3d
and 4d, which treat the solute as fixed, also gave water
densities that were similar to one another. Thus, MD and
RISM produce nearly the same hydration profile when the
solute is treated in the same manner. Irrespective of the
FIG. 1. ColorMD trajectory length dependence of the distribution of water oxygen around DNA. Shown in blue are regions in which the density of the water
oxygen is 2.25 times greater than that of the bulk water. All the figures except Fig. 4were created using gOpenMol Ref. 37.
FIG. 2. ColorVariation in the distribution of the water oxygen around DNA, depending on the number of DNA conformations used in the RISMflex
calculation. Shown in yellow are regions in which the density of the water oxygen is 1.7 times greater than that of the bulk water.
185102-4 Yonetani et al. J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
method used, the flexible models MDflex and RISMflexgave
less dense and broader distributions of water oxygen than the
rigid models MDrigid and RISMrigid. This is because the
structural variability of a flexible solute increases possible
hydration patterns.
To assess the patterns of DNA hydration in more detail,
we compared the MDflex and RISMflex results while focusing
on three representative regions Fig. 5. The water density
thresholds were set at 2.25 for MD and 1.7 for RISM. The
value of 1.7 was chosen for RISM so that the resultant spa-
tial distribution of the water resembled the MD result ob-
tained with a density threshold of 2.25. This slightly lower
threshold is likely attributable to an approximation employed
in the current 3D-RISM calculation. It is known that 3D-
RISM theory has a tendency to broaden the resultant distri-
bution functions.11 Another noteworthy difference between
the MD and RISM results is that, with RISM, high-density
regions are continuous with neighboring high-density re-
gions. This is also attributable to the characteristic spread of
the distribution functions seen without a treatment of inho-
mogeneity near the solute molecule.29
Except for the differences mentioned above, the MD and
3D-RISM results illustrated in Fig. 5are very similar. In the
central region of the ATAT sequence Fig. 5a, the hydra-
tion sites are aligned along the minor groove, which is the
characteristic hydration pattern for DNA so called “spine
hydration”. At one end of the DNA molecule, the CGCG
region Fig. 5b, the minor groove is slightly wider than in
the central ATAT region and shows a hydration pattern dif-
ferent from that seen in the ATAT region. In the CGCG re-
gion, water hydration sites are split into two branches, which
is consistent with an observation made in an earlier simula-
FIG. 3. ColorDistributions of water oxygen obtained using the MDflex a,MD
rigid b, RISMflex c, and RISMrigid dcalculations. Shown are the
distributions of water oxygen obtained with the indicated density thresholds, which are multiples of the water oxygen density in the bulk water.
FIG. 4. ColorWater oxygen distribution on a slice
plane parallel to the DNA base pairs: aMDflex,
bMDrigid,cRISMflex, and dRISMrigid. Water oxy-
gen densities expressed as multiples of that in the bulk
water are given in the color scale. The figure was
created by Chimera Ref. 38.
185102-5 DNA hydration patterns J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
tion study.8Compared to the minor groove, the major groove
exhibits a less dense distribution with both MD and 3D-
RISM Fig. 5c. In the major groove, various network pat-
terns of water molecules become possible as a result of the
width of the groove. Consequently, water localization at the
groove surface becomes weaker. In this respect, the
3D-RISM and MD calculations produced consistent
hydration results.
Next, we compared the calculated distribution of water
molecules with the positions of water molecules determined
by x-ray crystallographic analysis.30 In Fig. 6, the experi-
mental water oxygen atoms are represented by five different
colors, depending on their B-factor. Within the crystal struc-
ture, highly localized water molecules blue, B-factor
10 Å2are aligned along the minor groove. Both MD- and
3D-RISM-derived water distributions agreed well with the
water positions in the crystal structure, although there were
slight positional displacements. This small difference prob-
ably reflects the different sample conditions used i.e., solu-
tion for the calculations and crystal for the experimentor
imperfection of the interaction potential employed in the cal-
culations. Then, we analyzed patterns of hydration around
the phosphate moieties of the DNA backbone. At a density
threshold of 2.25, MD-derived hydration was not observed
Fig. 6a. As already shown in Fig. 3a, upon reducing the
density threshold to 1.7, the MD-derived hydration pat-
terns became visible and were similar to the RISM-derived
patterns Fig. 6b, which indicates that the water distribu-
tion is less dense and broader in the backbone region than the
minor groove region. This tendency is consistent with the
crystal structure, where only a few oxygen positions with
large B-factors were observed in the backbone region.
The distribution of water hydrogen atoms together with
that of oxygen atoms is shown in Fig. 7. In the MD result
Fig. 7a, hydrogen atoms are distributed on two sides of
the oxygen distribution, indicating that a hydrogen-bond net-
work is formed with the surface water molecules. A particu-
larly clear picture of spine hydration is seen in the central
ATAT region Fig. 7c: Water molecules in the first layer
form bridges between two acceptor atoms Nblueor O
red兲兴 in diagonally opposing bases belonging to different
DNA strands, and the second layer of water molecules com-
prises a bridge that connects two neighboring water mol-
ecules in the first layer. Unfortunately, a similarly detailed
hydration picture could not be specified from the RISM cal-
culation Fig. 7b. This problem may be related to the
broadening of the distribution functions caused by neglecting
the effect of inhomogeneity and employing a relatively large
grid size for the feasible calculations at present. Recently,
Ishizuka et al. have proposed a new integral equation theory
for inhomogeneous molecular fluids.29 We will report further
investigations on the hydration/solvation structures around
biomolecules including inhomogeneity with a finer grid size.
Finally, we mention the behavior of counterions around
DNA. Although ions were not treated as a subject of the
current MD-RISM comparison, they are expected to have a
profound effect on the DNA stability. As previously
recognized,31,32 it is not easy for MD simulation to explore
FIG. 5. ColorCalculated hydration patterns in the mi-
nor groove of the central ATAT region a, the minor
groove of the CGCG end region b, the major groove
of the central ATAT region c.MDblueand RISM
yellowresults are given in the left and central views,
respectively. MD and RISM results are superimposed in
the right view.
185102-6 Yonetani et al. J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
the properties of ions because of the increased difficulty with
their configuration sampling. Since the number of ions in-
volved in the calculation is very small compared to the water
number e.g., the present MD system has only 39 K+and 17
Clin 5000 H2O, it becomes difficult to obtain reliable
statistics of ions. Furthermore, ions will not stay at specific
sites. This also makes sampling of the ion configurations
more difficult. Using the present 8 ns trajectories, we calcu-
lated 3D distributions of K+and Clions in the same manner
to that of water molecules. The result was roughly similar to
the recent results31 which were derived from much longer
simulations of 50 ns: Cations preferentially distributed
along the minor groove between the two DNA strands. How-
ever, it is apparent that the time length of 8 ns was too short
to obtain a reliable ion distribution. Because the present
DNA of CGCGATATCGCG has a palindromic sequence, the
FIG. 6. ColorStereo graphics in which the calculated
results are compared to experimental data. In the upper
MD; blueand bottom RISM; yellowpanels, the re-
spective density thresholds were 2.25 and 1.7. Oxygen
positions determined by x-ray crystallography Ref. 30
PDB ID 287dare indicated by spheres colored ac-
cording to the experimental B-factors. Only water mol-
ecules located within 5 Å of the DNA surface atoms are
shown; those situated at a short distance from
other crystal packed DNA molecules are excluded.
FIG. 7. ColorDistribution of water
hydrogen and oxygen. The oxygen
distribution is the same as in Fig. 5
and is colored blue for MD aand
yellow for RISM b. The hydrogen
distribution is shown in white with
both MD and RISM. Note that differ-
ent values are used for the density
thresholds in the MD and RISM calcu-
lations see the text for the details.c
Closer inspection of the MD distribu-
tion in a. Red and blue spheres are
oxygen and nitrogen atoms in the
DNA bases, respectively, which act as
acceptors for hydrogen atoms in the
first layer water molecules.
185102-7 DNA hydration patterns J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
resultant ion distribution should have a corresponding sym-
metry if sufficient sampling is achieved. Contrary to this ex-
pectation, the present result from the 8 ns trajectories exhib-
ited an explicit deviation from such symmetry. A similar
deviation has been confirmed from a previous short-time
simulation.32 On the basis of the recent thorough
evaluations,31 time length of 50 ns is required to remove
such an undesirable deviation. Accordingly, the time length
required for constructing a satisfactory 3D distribution of
ions will become several times as long as the current study.
In such a case, it would be necessary to carefully consider a
recently mentioned force-field problem on ions.33 A necessity
of readjusting ion parameters has been inferred from a recent
study using ionic solutions. Solving the spatial distribution of
DNA surrounding ions will be a good subject to illustrate the
advantage of the RISM approach, because RISM is free from
the sampling problem unlike MD simulation.
IV. CONCLUDING REMARKS
Using two distinct computational methods i.e., MD
simulation and 3D-RISM theory, the spatial distributions of
water molecules around a DNA fragment were calculated to
assess the respective capacities of the two methods to simu-
late patterns of biomolecule hydration. This study confirmed
that the results agree well when the solute treatment is con-
sistent. If that is not the case, however, MD and 3D-RISM
will yield differing results. We demonstrated that when as-
suming a solute DNA molecule to be rigid, as in conven-
tional RISM treatments, the solvent distribution near the sol-
ute is about twice as dense as that obtained with the more
realistic MDflex calculation. This suggests that we can never
obtain a true picture of the hydration of solute biomolecules
without considering their flexibility. To cope with this prob-
lem, some improvements in RISM theory based on combi-
nation with the simulation techniques have been proposed
and tested.34 We anticipate that the agreement between the
simulation and theoretical approaches will form the basis for
further methodological development.
The most advantageous aspect of computational ap-
proaches is that they can tackle subjects that are experimen-
tally difficult to observe. A good example is determination of
the positions of the hydrogen atoms of hydrating water,
which is quite difficult to achieve with x-ray crystallography
because of the small atomic scattering factor. By contrast,
with a computational approach, hydrogen atoms are readily
observable. We anticipate that in the near future much infor-
mation on the distribution of water hydrogen will be ob-
tained through neutron diffraction experiments carried out at
the Japan Proton Accelerator Research Complex J-PARC
facility. We look forward to comparing the present computa-
tional results with those experimental data. Another interest-
ing problem that can be computationally approached is the
sequence-dependent properties of DNA.35 To systematically
investigate the effect of sequence on hydration, measure-
ments from many samples with differing sequences are
needed. Computational approaches are particularly well
suited for such an investigation because they are not subject
to any of the experimental difficulties associated with crys-
tallization. We are currently performing an analysis of
sequence-dependent hydration patterns vis-à-vis the
conformational properties of DNA,36 which will be reported
elsewhere.
ACKNOWLEDGMENTS
This work was supported by a grant from Scientific
Research on Priority Area of “Water and Biomolecules,”
and in part by Grants-in-Aid for Scientific Research No.
18031042 H.K.from Ministry of Education, Culture,
Sports, Science and Technology in Japan. Y.M. and F.H.
were supported by the grant from the Next Generation
Supercomputing Project, Nanoscience Program of the minis-
try. Y.Y. was supported by JSPS Research Fellowships for
Young Scientists.
1T. M. Raschke, Curr. Opin. Struct. Biol. 16,1522006; P. Auffinger and
Y. Hashem, ibid. 17, 325 2007.
2S. Arai, T. Chatake, T. Ohhara, K. Kurihara, I. Tanaka, N. Suzuki, Z.
Fujimoto, H. Mizuno, and N. Niimura, Nucleic Acids Res. 33, 3017
2005.
3M. Nakasako, J. Mol. Biol. 289, 547 1999; N. Niimura, S. Arai, K.
Kurihara, T. Chatake, I. Tanaka, and R. Bau, Cell. Mol. Life Sci. 63,285
2006.
4M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids
Clarendon, Oxford, 1987.
5J.-P. Hansen and I. R. McDonald, Theory of Simple Liquids Academic,
New York, 1986.
6F. Hirata, Molecular Theory of Solvation Kluwer, Dordrecht, 2003.
7V. Lounnas, B. M. Pettitt, and G. N. Phillips, Jr., Biophys. J. 66,601
1994; J. Higo, H. Kono, N. Nakajima, H. Shirai, H. Nakamura, and A.
Sarai, Chem. Phys. Lett. 306, 395 1999; J. Higo, H. Kono, H.
Nakamura, and A. Sarai, Proteins 40, 193 2000.
8V. P. Chuprina, U. Heinemann, A. A. Nurislamov, P. Zielenkiewicz, R. E.
Dickerson, and W. Saenger, Proc. Natl. Acad. Sci. U.S.A. 88,593
1991.
9P. Auffinger and E. Westhof, J. Mol. Biol. 300, 1113 2000;305, 1057
2001; M. Feig and B. M. Pettitt, ibid. 286, 1075 1999.
10 C. Chen, B. W. Beck, K. Krause, and B. M. Pettitt, Proteins 62, 982
2006.
11 T. Imai, R. Hiraoka, A. Kovalenko, and F. Hirata, Proteins 66,804
2007.
12 T. Imai, Y. Harano, M. Kinoshita, A. Kovalenko, and F. Hirata, J. Chem.
Phys. 126, 225102 2007; T. Yamazaki, T. Imai, F. Hirata, and A.
Kovalenko, J. Phys. Chem. B 111, 1206 2007.
13 H. R. Drew and R. E. Dickerson, J. Mol. Biol. 151,5351981.
14 P. L. Privalov, A. I. Dragan, C. Crane-Robinson, K. J. Breslauer, D. P.
Remeta, and C. A. S. A. Minetti, J. Mol. Biol. 365,12007.
15 J. Wang, P. Cieplak, and P. A. Kollman, J. Comput. Chem. 21,1049
2000.
16 T. E. Cheatham III, P. Cieplak, and P. A. Kollman, J. Biomol. Struct.
Dyn. 16, 845 1999.
17 W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L.
Klein, J. Chem. Phys. 79, 926 1983.
18 B. M. Pettitt and P. J. Rossky, J. Chem. Phys. 77,14511982.
19 D. A. Case, D. A. Pearlman, J. W. Caldwell, T. E. Cheatham III, J. Wang,
W. S. Ross, C. L. Simmerling, T. A. Darden, K. M. Merz, R. V. Stanton,
A. L. Cheng, J. J. Vincent, M. Crowley, V. Tsui, H. Gohlke, R. J. Radmer,
Y. Duan, J. Pitera, I. Massova, G. L. Seibel, U. C. Singh, P. K. Weiner,
and P. A. Kollman, AMBER7, University of California, San Francisco,
2002.
20 J.-P. Ryckaert, G. Ciccotti, and H. J. C. Berendsen, J. Comput. Phys. 23,
372 1977.
21 T. Darden, D. York, and L. Pedersen, J. Chem. Phys. 98, 10089 1993.
22 H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. DiNola, and
J. R. Haak, J. Chem. Phys. 81, 3684 1984.
23 D. Beglov and B. Roux, J. Phys. Chem. B 101, 7821 1997;A.
Kovalenko and F. Hirata, Chem. Phys. Lett. 290,2371998.
24 A. Kovalenko and F. Hirata, J. Chem. Phys. 11 0, 10095 1999.
25 A. Kovalenko and F. Hirata, J. Chem. Phys. 11 2, 10391 2000.
185102-8 Yonetani et al. J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
26 J. S. Perkyns and B. M. Pettitt, J. Chem. Phys. 97, 7656 1992;Chem.
Phys. Lett. 190,6261992.
27 A. Kovalenko and F. Hirata, J. Chem. Phys. 11 2, 10403 2000.
28 A. Kovalenko and F. Hirata, J. Phys. Chem. B 103,79421999.
29 R. Ishizuka, S.-H. Chong, and F. Hirata, J. Chem. Phys. 128, 034504
2008.
30 M. Shatzky-Schwartz, N. D. Arbuckle, M. Eisenstein, D. Rabinovich, A.
Bareket-Samish, T. E. Haran, B. F. Luisi, and Z. Shakked, J. Mol. Biol.
267,5951997.
31 S. Y. Ponomarev, K. M. Thayer, and D. L. Beveridge, Proc. Natl. Acad.
Sci. U.S.A. 101, 14771 2004; P. Varnai and K. Zakrzewska, Nucleic
Acids Res. 32,42692004.
32 M. A. Young, G. Ravishanker, and D. L. Beveridge, Biophys. J. 73, 2313
1997.
33 P. Auffinger, T. E. Cheatham III, and A. C. Vaiana, J. Chem. Theory
Comput. 3, 1851 2007.
34 M. Kinoshita, Y. Okamoto, and F. Hirata, J. Am. Chem. Soc. 120, 1855
1998; A. Mitsutake, M. Kinoshita, Y. Okamoto, and F. Hirata, J. Phys.
Chem. B 108, 19002 2004; T. Miyata and F. Hirata, J. Comput. Chem.
29, 871 2008.
35 M. J. Arauzo-Bravo, S. Fujii, H. Kono, S. Ahmad, and A. Sarai, J. Am.
Chem. Soc. 127, 16074 2005; D. L. Beveridge, G. Barreiro, K. Suzie
Byun, D. A. Case, T. E. Cheatham III, S. B. Dixit, E. Giudice, F. Lankas,
R. Lavery, J. H. Maddocks, R. Osman, E. Seibert, H. Sklenar, G. Stoll, K.
M. Thayer, P. Varnai, and M. A. Young, Biophys. J. 87, 3799 2004.
36 Y. Yonetani, H. Kono, S. Fujii, A. Sarai, and N. Go, Mol. Simul. 33,103
2007.
37 D. L. Bergman, L. Laaksonen, and A. Laaksonen, J. Mol. Graphics
Modell. 15, 301 1997.
38 E. F. Pettersen, T. D. Goddard, C. C. Huang, G. S. Couch, D. M.
Greenblatt, E. C. Meng, and T. E. Ferrin, J. Comput. Chem. 25,1605
2004.
185102-9 DNA hydration patterns J. Chem. Phys. 128, 185102 2008
Downloaded 30 May 2008 to 133.48.169.109. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
... The three dimensional interaction site model (3D-RISM) (38)(39)(40) is a molecular solvation theory method that treats ion size effects and ion-ion correlation but unlike MD simulation, can be practically applied for the prediction of solvation thermodynamics over a wide range of conditions (41,42). 3D-RISM calculations on nucleic acids are much less computationally intensive than MD simulations and calculations can be easily converged in the presence of micro-molar to molar salt concentrations (36). ...
... 3D-RISM calculations on nucleic acids are much less computationally intensive than MD simulations and calculations can be easily converged in the presence of micro-molar to molar salt concentrations (36). 3D-RISM and MD have been shown to yield similar layered solvent and ion distributions around nucleic acids and other biopolymers (36,(38)(39)(40). ...
... There is no a priori guarantee that force fields developed for MD simulation should perform comparably when used with 3D-RISM. However the similarity between particle distributions obtained from MD and 3D-RISM (36,(38)(39)(40) suggest that current force fields can be used as a starting point for testing 3D-RISM against experimental data. Bulk properties of alkali halides determined using dielectrically consistent DRISM have been recently evaluated against a wide variety experimental measurements with the general conclusion that DRISM provides semi-quantitative agreement with experiment (70). ...
Article
Full-text available
The composition of the ion atmosphere surrounding nucleic acids affects their folding, condensation and binding to other molecules. It is thus of fundamental importance to gain predictive insight into the formation of the ion atmosphere and thermodynamic consequences when varying ionic conditions. An early step toward this goal is to benchmark computational models against quantitative experimental measurements. Herein, we test the ability of the three dimensional reference interaction site model (3D-RISM) to reproduce preferential interaction parameters determined from ion counting (IC) experiments for mixed alkali chlorides and dsDNA. Calculations agree well with experiment with slight deviations for salt concentrations >200 mM and capture the observed trend where the extent of cation accumulation around the DNA varies inversely with its ionic size. Ion distributions indicate that the smaller, more competitive cations accumulate to a greater extent near the phosphoryl groups, penetrating deeper into the grooves. In accord with experiment, calculated IC profiles do not vary with sequence, although the predicted ion distributions in the grooves are sequence and ion size dependent. Calculations on other nucleic acid conformations predict that the variation in linear charge density has a minor effect on the extent of cation competition. © The Author(s) 2015. Published by Oxford University Press on behalf of Nucleic Acids Research.
... The 3D-RISM-KH theory has been validated on both simple and complex associating liquids and solutes with different chemical functionalities [14][15][16][17][18][19][20][21][22][23][24][25], including ionic liquids [19], and polyelectrolyte gels [20], in a range of fluid thermodynamic conditions [4,[21][22][23][24], in various environments such as interfaces with metal [10,12], metal oxide [26], zeolite [13,27], clay [28,29], and in confinement of carbon nanotubes [16], synthetic organic rosette nanotubes [14,[30][31][32], nanocellulose based bionanomaterials [33][34][35], and biomolecular systems [14,. The latter include case studies ranging from the structure of hydrated alanine dipeptide [36][37][38][39], miniprotein 1L2Y and protein G [39]; structural water, xenon, and ions bound to lysozyme protein [40,41]; selective binding and permeation of water, ions and protons in channels [14,20,[41][42][43][44][45]; salt-induced conformational transitions of DNA [45][46][47]; partial molar volume and pressure induced conformational transitions of biomolecules [48][49][50][51][52]; formation and conformational stability of β-sheet Amyloid-β (Aβ) oligomers [53], HET-s Prion and Aβ fibrillous aggregates [54,55]; ligand-binding affinities 32601-2 of seven biotin analogues to avidin in aqueous solution [56]; binding modes of inhibitors of pathologic conversion and aggregation of prion proteins [20,55], binding modes of thiamine against extracytoplasmic thiamine binding lipoprotein MG289 [57,58], binding modes and ligand efflux pathway in multidrug transporter AcrB [59], and maltotriose ligand binding to maltose-binding protein in which structural water triggers binding modes with protein domain motion resulting in the holo-closed structure that binds the ligand [60]; to aqueous electrolyte solution at physiological concentrations in biomolecular systems as large as the Gloebacter violaceus pentameric ligand-gated ion channel (GLIC) homologue in a lipid bilayer [14,20,57] and structural water promoting folding in the GroEL/ES chaperonin complex [61]. The 3D-RISM-KH theory provided an insight into such soft matter phenomena as the structure and stability of gels formed by oligomeric polyelectrolyte gelators in different solvents [20]. ...
Article
Full-text available
Integral equation theory of molecular liquids based on statistical mechanics is quite promising as an essential part of multiscale methodology for chemical and biomolecular nanosystems in solution. Beginning with a molecular interaction potential force field, it uses diagrammatic analysis of the solvation free energy to derive integral equations for correlation functions between molecules in solution in the statistical-mechanical ensemble. The infinite chain of coupled integral equations for many-body correlation functions is reduced to a tractable form for 2- or 3-body correlations by applying the so-called closure relations. Solving these equations produces the solvation structure with accuracy comparable to molecular simulations that have converged but has a critical advantage of readily treating the effects and processes spanning over a large space and slow time scales, by far not feasible for explicit solvent molecular simulations. One of the versions of this formalism, the three-dimensional reference interaction site model (3D-RISM) integral equation complemented with the Kovalenko-Hirata (KH) closure approximation, yields the solvation structure in terms of 3D maps of correlation functions, including density distributions, of solvent interaction sites around a solute (supra)molecule with full consistent account for the effects of chemical functionalities of all species in the solution. The solvation free energy and the subsequent thermodynamics are then obtained at once as a simple integral of the 3D correlation functions by performing thermodynamic integration analytically.
... 3D-RISM calculations are thus potentially very powerful as tools to study the ion atmosphere for nucleic acid systems that can be represented by a relatively small ensemble of rigid conformations. Both MD and 3D-RISM use molecular force fields and, unlike NLPB, yield similar layered solvent and ion distributions Maruyama, Yoshida, & Hirata, 2010;Yonetani, Maruyama, Hirata, & Kono, 2008;Howard, Lynch, & Pettitt, 2011) (Figure 3). A challenge for both 3D-RISM calculations and MD simulations that has been fully recognized only recently (A. A. Chen, Draper, & Pappu, 2009;Giambaşu et al., 2014;Yoo & Aksimentiev, 2012) is the need to consider a sufficiently large amount of solvent such that regions far from the solute exhibit bulk behavior. ...
Article
RNA catalysis is of fundamental importance to biology and yet remains ill-understood due to its complex nature. The multidimensional "problem space" of RNA catalysis includes both local and global conformational rearrangements, changes in the ion atmosphere around nucleic acids and metal ion binding, dependence on potentially correlated protonation states of key residues, and bond breaking/forming in the chemical steps of the reaction. The goal of this chapter is to summarize and apply multiscale modeling methods in an effort to target the different parts of the RNA catalysis problem space while also addressing the limitations and pitfalls of these methods. Classical molecular dynamics simulations, reference interaction site model calculations, constant pH molecular dynamics (CpHMD) simulations, Hamiltonian replica exchange molecular dynamics, and quantum mechanical/molecular mechanical simulations will be discussed in the context of the study of RNA backbone cleavage transesterification. This reaction is catalyzed by both RNA and protein enzymes, and here we examine the different mechanistic strategies taken by the hepatitis delta virus ribozyme and RNase A. © 2015 Elsevier Inc. All rights reserved.
Article
We calculated solvent accessibility of DNA backbone hydrogen sites, H1’-H5’ by using molecular dynamics simulation of DNA. The result of accessibility is well correlated with the site-dependent reactivity with OH radicals experimentally reported, indicating that the different DNA-radical reactivity is mainly caused by the difference in the solvent accessibility of each hydrogen site. Compared with the previous calculation with solvent-accessible surface area, the present MD-based counting of molecular access provided a slightly improved result, which suggests importance of more realistic molecular components such as electrostatic interactions and DNA conformational fluctuation.
Article
Full-text available
The integration equation theory (IET) provides highly efficient tools for the calculation of structural and thermodynamic properties of molecular liquids. In recent years, the 3D reference interaction site model (3DRISM), the most developed IET for solvation, has been widely applied to study protein solvation, aggregation, and drug‐receptor binding. However, hydrophobic solutes with sufficient size (>nm) can induce water density depletion at the solute–solvent interface. This density depletion is not considered in the original 3DRISM theory. The authors here review the recent developments of 3DRISM at hydrophobic surfaces and related theories to address this challenge. At hydrophobic surfaces, an additional hydrophobicity‐induced density inhomogeneity equation is introduced to 3DRISM theory to consider this density depletion. Accordingly, several new closures equations including D2 closure and D2MSA closures are developed to enable stable numerical solutions of 3DRISM equations. These newly developed theories hold great promise for an accurate and rapid calculation of the solvation effect for complex molecular systems such as proteins. At the end of the report, the authors also provide a perspective on other challenges of the IETs as an efficient solvation model.
Book
The idea of theoretically predicting the useful properties of various materials using multiscale simulations has become popular in recent years. Of special interest are nanostructured, organic functional materials, which have a hierarchical structure and are considered materials of the future because of their flexibility and versatility. Their functional properties are inherited from the molecule that lies at the heart of the hierarchical structure. On the other hand, the properties of this functional molecule, in particular its absorption and emission spectra, strongly depend on its interactions with its molecular environment. Therefore, the multiscale simulations used to predict the properties of organic functional materials should be atomistic, that is, they should be based on classical and/or quantum methods that explicitly take into account the molecular structure and intermolecular interactions at the atomic level. This book, written by well-known specialists in theoretical chemistry, focuses on the basics of classical mechanics, quantum chemistry methods used for molecular disordered materials, classical methods of molecular simulations of disordered materials, vibronic interactions, and applications (presented as multiscale strategies for atomistic simulations of photonic materials). It has been edited by Professor Mikhail Alfimov, a renowned Russian scientist, a full member of the Russian Academy of Sciences, Russia, and the founder, first director, and now research supervisor of the Photochemistry Center of the Russian Academy of Science, Russia. Professor Alfimov’s main research interests are in the field of photochemistry and photophysics of molecular and supramolecular systems. The book is a great reference for advanced undergraduate- and graduate-level students of nanotechnology and molecular science and researchers in nano- and molecular science, nanotechnology, chemistry, and physical chemistry, especially those with an interest in functional materials. © 2018 by Pan Stanford Publishing Pte. Ltd. All rights reserved.
Chapter
In this chapter, we present a theoretical treatment on the molecular recognition, one of elementary processes of life phenomena, based on the 3D-RISM/RISM theory. The theory has been applied successfully to a variety of molecular recognition processes. It has proved its ability to discriminate molecules having different charges and structures, which is an essential requirement for the in silico drug design. The structural fluctuation plays a crucial role in the process where a protein expresses its function. A new theory to characterize the structural fluctuation of protein around its native state, which combines the 3D-RISM/RISM theory with the generalized Langevin theory, is also presented here.
Article
The so-called three-dimensional version (3D-RISM) can be used to describe the interactions of solvent components (here we treat water and ions) with a chemical or biomolecular solute of arbitrary size and shape. Here we give an overview of the current status of such models, describing some aspects of “pure” electrolytes (water plus simple ions) and of ionophores, proteins and nucleic acids in the presence of water and salts. Here we focus primarily on interactions with water and dissolved salts; as a practical matter, the discussion is mostly limited to monovalent ions, since studies of divalent ions present many difficult problems that have not yet been addressed. This is not a comprehensive review, but covers a few recent examples that illustrate current issues.
Article
We develop a method to extract site-site bridge functions from simulations of molecular liquids with the use of dielectric consistent reference interaction site model (DRISM). We reduce the problem to evaluations of short-range parts of direct correlation functions, providing consistent corrections to the DRISM integral equations. We reconstruct site-site bridge functions with the use of the corrected DRISM equations and Fourier transforms of simulated radial distribution functions. The performance of the algorithm has been tested for the SPC/E water model. The obtained bridge functions are quite close to those obtained with the use of the RISM approach. We evaluate structural properties of ambient water within the DRISM framework, using the bridge functions. All the calculated structural properties are in a good agreement with available experimental data.
Article
We calculated the binding free energy of short oligonucleotides (9–20-mers DNA) by using molecular dynamics, followed with post-processing based on molecular theory of solvation (MM/3D-RISM-KH approach). A comparison with the PBSA and GBSA continuum solvation models was performed to identify the approach in the best agreement with experimental results for the binding free energy. Compared to the PBSA or GBSA methods, the 3D-RISM-KH molecular theory of solvation provides a more accurate description of the nonpolar contribution to the solvation free energy from the first principles of statistical mechanics. The binding free energy was calculated by using separate trajectories for the DNA complex and its two strands, as well as based on a single trajectory for the complex, both with and without account for explicit counter ions in post-processing of molecular dynamics trajectories to calculate the binding free energy of oligonucleotides. Overall, both GBSA and 3D-RISM-KH predict the binding free energy obtained from “separate trajectory” calculations with implicit account for counter ions in a good agreement with experiment, the latter showing a better performance for larger oligonucleotides.
Article
Full-text available
The method to calculate the potential of mean force in ion-molecular solution directly as a difference between the chemical potential of solvation of a pair of solutes at a given relative arrangement and at infinite separation was elaborated. The solvation chemical potential was obtained by using the 3D generalization of the reference interaction site model (RISM)/hypernetted chain (HNC) theory. The method allows the evaluation of the solvation structure and thermodynamics for complex ionic and polar molecular solvent and substantially improve the understanding of solvation properties of aqueous electrolytes as well as other ion-molecular solutions.
Article
Full-text available
DNA tetramer sequences AATT and TTAA are known to be conformationally more rigid and flexible, respectively. In this study, we carry out molecular dynamics (MD) simulations of these two sequences and investigate the characteristic hydration pattern. The rigid AATT is found to be more likely to construct the hydration spine in the minor groove, than the flexible TTAA. The result suggests that the hydration water molecules play a critical role, for determining the sequence dependent deformability of DNA conformation.
Book
Molecular Theory of Solvation presents the recent progress in the statistical mechanics of molecular liquids applied to the most intriguing problems in chemistry today, including chemical reactions, conformational stability of biomolecules, ion hydration, and electrode-solution interface. The continuum model of "solvation" has played a dominant role in describing chemical processes in solution during the last century. This book discards and replaces it completely with molecular theory taking proper account of chemical specificity of solvent. The main machinery employed here is the reference-interaction-site-model (RISM) theory, which is combined with other tools in theoretical chemistry and physics: the ab initio and density functional theories in quantum chemistry, the generalized Langevin theory, and the molecular simulation techniques. This book will be of benefit to graduate students and industrial scientists who are struggling to find a better way of accounting and/or predicting "solvation" properties.
Article
In this study, we present conformational energies for a molecular mechanical model (Parm99) developed for organic and biological molecules using the restrained electrostatic potential (RESP) approach to derive the partial charges. This approach uses the simple “generic” force field model (Parm94), and attempts to add a minimal number of extra Fourier components to the torsional energies, but doing so only when there is a physical justification. The results are quite encouraging, not only for the 34-molecule set that has been studied by both the highest level ab initiomodel (GVB/LMP2) and experiment, but also for the 55-molecule set for which high-quality experimental data are available. Considering the 55 molecules studied by all the force field models for which there are experimental data, the average absolute errors (AAEs) are 0.28 (this model), 0.52 (MM3), 0.57 (CHARMm [MSI]), and 0.43 kcal/mol (MMFF). For the 34-molecule set, the AAEs of this model versus experiment and ab initio are 0.28 and 0.27 kcal/mol, respectively. This is a lower error than found with MM3 and CHARMm, and is comparable to that found with MMFF (0.31 and 0.22 kcal/mol). We also present two examples of how well the torsional parameters are transferred from the training set to the test set. The absolute errors of molecules in the test set are only slightly larger than in the training set (differences of <0.1 kcal/mol). Therefore, it can be concluded that a simple “generic” force field with a limited number of specific torsional parameters can describe intra- and intermolecular interactions, although all comparison molecules were selected from our 82-compound training set. We also show how this effective two-body model can be extended for use with a nonadditive force field (NAFF), both with and without lone pairs. Without changing the torsional parameters, the use of more accurate charges and polarization leads to an increase in average absolute error compared with experiment, but adjustment of the parameters restores the level of agreement found with the additive model. After reoptimizing the Ψ, Φ torsional parameters in peptides using alanine dipeptide (6 conformational pairs) and alanine tetrapeptide (11 conformational pairs), the new model gives better energies than the Cornell et al. ( J Am Chem Soc 1995, 117, 5179–5197) force field. The average absolute error of this model for high-level ab initio calculation is 0.82 kcal/mol for alanine dipeptide and tetrapeptide as compared with 1.80 kcal/mol for the Cornell et al. model. For nucleosides, the new model also gives improved energies compared with the Cornell et al. model. To optimize force field parameters, we developed a program called parmscan, which can iteratively scan the torsional parameters in a systematic manner and finally obtain the best torsional potentials. Besides the organic molecules in our test set, parmscan was also successful in optimizing the Ψ, Φ torsional parameters in peptides to significantly improve agreement between molecular mechanical and high-level ab initio energies. © 2000 John Wiley & Sons, Inc. J Comput Chem 21: 1049–1074, 2000
Article
A liquid state theory based on site–site integral equations is constructed to have the asymptotics given by angular expansion theory. This results in a theory which shows dielectric consistency, e.g., the dielectric constant as viewed from the solvent is the same as that viewed by the ions. Such consistency is lacking in other extended reference interaction site model (XRISM)-based theories and leads to unrealistic structural predictions. The Kirkwood–Buff route to thermodynamics is used and allows a physical partitioning of the terms responsible for the solvation process. Sample results for a 1–1 salt are given.
Article
A reformulation of reference interaction site model theory is proposed. The approach makes use of the formally correct asymptotic form of the correlations obtained from the one-center angular expansion technique. A modified closure, or equivalently, a modified propagation equation for site—site correlations is shown to incorporate the necessary information to allow dielectric consistency in finite-concentration salt solutions. Examples of the correlations and thermodynamics are given.
Article
An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolutions using fast Fourier transforms. Timings and accuracies are presented for three large crystalline ionic systems. The Journal of Chemical Physics is copyrighted by The American Institute of Physics.
Article
We applied the three-dimensional reference interaction site model (3D-RISM) integral equation theory with the 3D hypernetted chain (3D-HNC) closure or its partial linearization (3D-PLHNC) to obtain the potentials of mean force (PMFs) and the solvation structure of sodium chloride in ambient water. The bulk solvent correlations are treated by the dielectrically consistent site-site RISM/HNC theory (DRISM/HNC) to provide a proper description of the dielectric properties of solution and to include the case of a finite salt concentration. The PMF is calculated as a difference in the solvation free energy of an ion pair and of individual ions. We obtained and analyzed in detail the PMFs and solvation structure for ion pairs of NaCl at infinite dilution and a concentration of 1 M. The results are in reasonably good agreement with molecular dynamics simulations for the same model of the solution species. Positions and orientations of water molecules in the first solvation shell around the ion pair are deduced. The short-range hydration structure of the ion pairs at infinite dilution and at moderate concentration is very similar. Ionic ordering and clustering is found in 1 M solution.
Article
An application of our recently developed extended RISM equation formulation to several three‐site models of water is presented. The site–site correlation functions are obtained and compared to available computer simulation results. Further, the variation of liquid state structure with the model site charge is examined. The analysis of these results has demonstrated that the integral equation approach provides a correct qualitative description of the liquid structure, although the amplitudes of most structural features are somewhat less accurate that their positions. Comparison to our earlier results for simpler models suggests that the nature of the quantitative deficiencies of the approach is predictable. The charging study has shown that the development of waterlike structure with increasing site charge follows a qualitatively different pattern for oxygen–oxygen pairs, compared to those involving hydrogen. This is attributed to interference between the amplitudes characteristic of liquid water and of simple liquids. These results suggest that this is the origin of a relatively flat 0–0 correlation function for several models studied in the past, and, further, that such results should not be properly characterized as ’’unstructured.’’
Article
We have developed a self-consistent description of an interface between a metal and a molecular liquid by combination of the density functional theory in the Kohn–Sham formulation (KS DFT) for the electronic structure, and the three-dimensional generalization of the reference interaction site model (3D RISM) for the classical site distribution profiles of liquid. The electron and classical subsystems are coupled in the mean field approximation. The procedure takes account of many-body effects of dense fluid on the metal–liquid interactions by averaging the pseudopotentials of liquid molecules over the classical distributions of the liquid. The proposed approach is substantially less time-consuming as compared to a Car–Parrinello-type simulation since it replaces molecular dynamics with the integral equation theory of molecular liquids. The calculation has been performed for pure water at normal conditions in contact with the (100) face cubic centered (fcc) surface of a metal roughly modeled after copper. The results are in good agreement with the Car–Parrinello simulation for the same metal model. The shift of the Fermi level due to the presence of water conforms with experiment. The electron distribution near an adsorbed water molecule is affected by dense water, and so the metal–water attraction follows the shapes of the metal effective electrostatic potential. For the metal model employed, it is strongest at the hollow site adsorption positions, and water molecules are adsorbed mainly at the hollow and bridge site positions rather than over metal atoms. Layering of water molecules near the metal surface is found. In the first hydration layer, adsorbed water molecules are oriented in parallel to the surface or tilted with hydrogens mainly outwards the metal. This orientation at the potential of zero charge agrees with experiment. © 1999 American Institute of Physics.