You are on page 1of 7

Article

pubs.acs.org/JPCB

Water−Methanol Mixtures: Simulations of Mixing Properties over the


Entire Range of Mole Fractions
Jean-Christophe Soetens* and Philippe A. Bopp
Institut des Sciences Moléculaires, UMR CNRS no. 5255, and Department of Chemistry, Université de Bordeaux, 351 Cours de la
Libération, 33405 Talence, France

ABSTRACT: Numerous experimental and theoretical investigations


have been devoted to the hydrogen bond in pure liquids and mixtures.
Among the different theoretical approaches, molecular dynamics (MD)
simulations are predominant in obtaining detailed information, on the
molecular level, simultaneously on the structure and the dynamics.
Water and methanol are the two most prominent hydrogen-bonded
liquids, and they and their mixtures have consequently been the subject
of many studies; we revisit here the problem of the mixtures. An
important first step is to check whether a classical potential model, the
components of which are deemed to be satisfactory for the pure liquids,
is able to reproduce the known thermodynamic excess properties of the
mixtures sufficiently well. We have used the available BJH (water) and
PHH (methanol) flexible models because they are by construction
mutually compatible and also well suited to study, in a second step, some
dynamic property characteristic of hydrogen-bonded liquids. In this article we show that these models, after a slight
reparametrization for use in NpT simulations, reproduce the essential features of the excess mixing and molar properties of
water−methanol mixtures. Furthermore, in the pure liquids, the agreement of the radial distribution functions with experiment
remains as satisfactory as before. Similarly, the translation self-diffusion coefficients D are modified by less than 10%. In the
mixtures, they evolve nonmonotonously as a function of mole fraction.

■ INTRODUCTION
Water (H2O) and methanol (CH3OH, MeOH) are completely
neat liquids, for which detailed analyses of the hydrogen bond
network have been carried out by combining MD simulations
miscible liquids at room temperature and ambient pressure, and vibrational spectroscopy experiments23−25 Applied to
as is to be expected from the polar character of these two mixtures such as water−methanol, we expect that this strategy
molecules. The molecular dipole moments (in the gas phase will lead to an improved qualitative and quantitative microscopic
μH2O = 1.86 D, μMeOH = 1.69 D) are similar; however, the static understanding of the peculiar properties of these systems. To
dielectric constant (relative permittivities) of the two neat achieve this goal, an important first step is to check whether a
liquids (εH2O ≈ 80, εMeOH ≈ 33) are markedly different because classical (flexible) potential model is able to reproduce the well-
of the different structures. known excess properties of these mixtures sufficiently well.
The presence of a hydrophobic group CH3 nevertheless Interaction models for mixtures are usually constructed by
contributes to peculiar properties such as the appearance of combining existing models developed and tested for the pure
a minimum in the partial molar volume of methanol. Thus, liquids, in most cases by keeping the partial charges of the initial
alcohols are sometimes seen as soluble hydrocarbons, and models and applying empirical combination rules (Lorentz−
models to explain the properties of their mixtures with water Berthelot,26 Kong,27 and others) for the Lennard-Jones 12−6
are still discussed.1−4 Molecular simulations are certainly the terms. In the present study we have started from the available
methods of choice to elucidate the properties of these systems BJH28 and PHH29 flexible models for water and methanol,
on a microscopic level. Most theoretical efforts have used respectively; they have already been used extensively. (See ref 30
molecular dynamics (MD) simulations to study the structure, and the references therein.) Having been devised to be mutually
the dynamic properties, and the properties of mixing, separately compatible, they are also suited to study spectroscopic pro-
or together.5−12 A number of experimental investigations have perties because they describe with good accuracy, in the
also been performed over the years, and most properties such as approximation of classical mechanics, the vibrational modes of
the density,13−15 the diffusion,16 the properties of mixing,17−20 water and methanol and in particular the shifts in the OH
and the structure1,21,22 are well known. stretching region.5
Even though there is a wealth of studies on mixtures, not
many combined efforts to obtain thermodynamic, structural, Received: April 7, 2015
and dynamic properties in a consistent fashion have been Revised: May 29, 2015
reported so far. This has, however, been accomplished for some

© XXXX American Chemical Society A DOI: 10.1021/acs.jpcb.5b03344


J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

Most of the previous MD studies of water−methanol mixtures Table 1. Refitted Parameters for the BJH28 and PHH29
have, however, been performed at constant volume (NVE or Models (Equations 1−4)
NVT ensembles), mostly at the experimental density at room
old new
temperature and normal pressure. Because both models were not parameters28,29 parameters
devised to reproduce the experimental densities, their use in
p1 1.045 1.463 kJ/mol
NpT simulations is not satisfactory. In some cases (depending,
p2 418 10.45 kJ/mol
for example, on mixture compositions) this may prove to be
p3 3.47 3.61 Å, water−methanol interactions only
impossible as they then predict unrealistic pressures. For these
p4 3.86 3.52 Å
reasons we did not investigate the energy properties in any
depth. The mixing effects are small, and comparing NVE and
NpT simulations, at different average pressures and temperatures oxygen−oxygen (VOO), oxygen−hydrogen (VOH), and hydrogen−
and with different box sizes and truncation schemes, would hydrogen (VHH) contributions are identical for the water−
require extensive work. Suffice it to say that we obtain the correct water, water−methanol, and water−methanol interactions, and
qualitative behavior of the potential energies as a function of the any change in the parametrization should retain this homogeneity.
mole fractions. 111 889
VOO = − p1 [exp(−4(r − 3.5)2 ) + exp(−1.5(r − 4.5)2 )]
In this article we first introduce a new parametrization for the r 8.86
BJH and PHH models to allow their use in NpT simulations (1)
of both the pure liquids and the mixtures over the entire 26.07 41.79 16.74
miscibility range, i.e. χ = 0 to 1. We have attempted to keep the VOH = 9.2 −
1 + exp(40(r − 1.05))

1 + exp(5.439(r − 2.2))
r
modifications as small as possible. We show that we can in this
(2)
way reproduce the essential features of the density, excess
volume, and partial molar volumes of water and methanol over p2
VHH =
the whole range of molar fraction reasonably well. These 1 + exp(29.9(r − 1.968)) (3)
properties of the binary systems are found to be quite sensitive
⎡⎛ p ⎞12 ⎛ p ⎞6⎤
to details of the intermolecular interactions, so the results VMeO = 4 × 0.3713⎢⎜ 3 ⎟ − ⎜ 3 ⎟ ⎥
presented below can be said to be a quite stringent test of the ⎢⎣⎝ r ⎠ ⎝ r ⎠ ⎥⎦
models. Of course, we also want to preserve the features that ⎡⎛ p ⎞12 ⎛ p ⎞6⎤
made the original models successful. We therefore proceeded to VMeMe = 4 × 0.1883⎢⎜ 4 ⎟ − ⎜ 4 ⎟ ⎥
⎢⎣⎝ r ⎠ ⎝ r ⎠ ⎥⎦ (4)
reevaluate some structural and dynamics properties of the pure
liquids and mixtures in a consistent fashion. In the next sections We note that the main modification concerns parameter p2,
we present first the modified model used in this study and the which essentially controls the H−H repulsion (cf. the original
details of the MD simulations. Then we look at the properties CF models32,33). The original large value of this parameter,
that we considered to be important and that were consequently which had essentially been carried over in later developments,
used to parametrize the model. After comparing the radial dis- was one of the main causes of the unsatisfactory pressures
tributions functions (RDFs) for the modified and unmodified obtained originally. Earlier work34 aiming essentially at bringing
models in the pure liquids, we proceed to discuss the radial pair the H−H radial distribution into better agreement with
distribution functions and translational self-diffusion constants experimental data35 involved adjusting the distance (1.968 Å)
in the mixtures.


where the repulsion sets in. It succeeded in improving the
structure but failed otherwise, among other things, because the
MODELS AND SIMULATION DETAILS self-diffusion coefficient increased by almost 40% in the case
Interaction Model. All simulations presented in this article where the best agreement of the gOO values was achieved.
make use of the BJH28 and PHH29 models for water and Simulation Runs. After a large number of trial runs to
methanol, respectively. In both models, the molecules are determine which parameters in the original model to keep and
represented by three interacting sites located on the oxygen and which ones to readjust (eqs 1−4) and determine the new values
hydrogen atoms of the hydroxyl group and on the methyl group (reported in Table 1), we have first carried out MD simulations
(which is thus a united atom, a drastic simplification). The total of the pure liquids in the NEV ensemble at ⟨T⟩ = 298 K and
potential energy is the sum of intra- and intermolecular con- experimental density.17 The electrostatic interactions were
tributions. The intramolecular parts, the anharmonicity of treated here, as everywhere else, with the reaction (RF)
which has been adjusted31 in the water case to yield, roughly, method36,37 (here for charge−charge interactions); the dielectric
the observed vibrational frequency shifts between the gas and constant of the surrounding continuum was set to εRF = ∞.
liquid phases,31 were not modified. The modifications of the A molecular cutoff equal to half of the edge of the box was
intermolecular potentials were kept, as already stated, minimal applied. The other empirical potential terms were truncated
and are not expected to affect the vibrational frequencies,31 so that numerical problems (e.g., under/overflows in the
which just should not require any modification of the intra- exponential functions) were avoided. These simulations were
molecular part. More details can be found in ref 5. carried out with N = 1000 molecules. The equations of motions
The model for the intermolecular potential energy is derived were integrated with the velocity Verlet algorithm38 with a time
from the central force model.32,33 The electrostatic part consists step equal to 0.2 fs. The simulations were run with the MDpol39
of Coulomb terms between partial charges (qO = −0.66e and program package, and each simulation consisted of two
qH = 0.33e for water; qO = −0.60e, qH = 0.35e, and qMe = 0.25e independent calculations of 500 ps (2.5 M steps) each after
for methanol). We recall below the empirical parts of the appropriate equilibration periods. During the simulations, a
intermolecular pair potential (in kJ mol−1 and Å); p1−p4 are small velocity rescaling has nevertheless been applied every
the parameters that we selected to be refitted in this work, and 0.1 ps in order to reach the desired average temperature with a
they are summarized in Table 1. It is important to note that the deviation of less than 1 K. The NpT simulations (at 298.15 K
B DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

and 1 atm) were carried out in a similar manner; i.e., all of the
technical details were the same excepted for the thermodynamic
ensemble controlled using the Nosé−Andersen algorithm40−42
(with coupling constants of 5 and 1 ps−1 with a heat bath and
piston, respectively).
We studied nine mixtures with mole fractions varying from
0.1 to 0.9 as well as the two pure liquids in NpT-MD simula-
tions (details in Table 3). In all cases, as above, two runs starting
from different and uncorrelated initial conditions were first
equilibrated before starting the 500 ps productions runs. The
results reported below stem from these simulations.
Thermodynamics. Mixtures of liquids such as the water− Figure 1. Density of water−methanol mixtures at T = 298 K,
methanol binary system can be characterized by their excess p = 1 atm as a function of the mole fraction of methanol: results from
mixing functions. For the volume, for instance, one can express NpT-MD simulations (dashed line and circles) and experimental
the corresponding excess property VE as results17 (solid line). The ideal densities, shown for comparison
(dashed−dotted line), are obtained from a linear combination of the
mole volumes of the pure compounds.
V E = Vmix − ∑ nivi(0)
i (5)

where Vmix is the volume of the mixture, v(0)i is the molar


volumes of the pure compounds, and ni is the number of moles
of each component. In the present binary mixtures, the partial
molar volume of each component, vi, can be obtained from
the following relations

Table 2. Comparison of Some Properties of Liquid Water


and Liquid Methanol under Ambient Conditions Obtained
from NEV-MD Simulations Using the Original Models
(BJH28 and PHH,29Respectively) and NpT-MD Simulations
with the New Parametrization for These Models, as Figure 2. Excess molar volumes obtained from NpT-MD simulations
Described in the Text at T = 298 K, p = 1 atm (dashed line and circles), experimental
results17 (solid line), and results from previous simulation work9
NEV NpT (dashed−dotted line).
⟨T⟩ =298 K T = 298.15 K, p = 1 atm
Water BJH Model Modified Model Table 4. Parameters Obtained from a Fit Using Equation 7
ρ (g cm−3) 0.99717 0.997 on Experimental Data17 and Previous Work from Diego
⟨Epot⟩ (kJ mol−1) −41.9 −45.8 González-Salgado et al.9 and the Present Worka
D (10−9 m2 s−1) 1.5 1.5
⟨p⟩ (atm) 4015 1 Ak exp previous work this work
Methanol PHH Model Modified Model A0 −4.033 −3.209 −3.997
ρ (g cm−3) 0.78717 0.788 A1 −0.294 −0.291 1.043
⟨Epot⟩ (kJ mol−1) −26.6 −28.2 A2 0.478 0.224
D (10−9 m2 s−1) 4.0 4.4 A3 0.851 1.033
a
⟨p⟩ (atm) 1000 1 The corresponding curves are presented in Figure 2.

Table 3. Details of the NpT-MD Simulations of Water−Methanol Mixtures, Mole Fractions, Number of Molecules in the
Simulation Boxes, Experimental17 and Simulated Densities, and Excess Volumes of the Systemsa
density (g cm−3) VE (cm3 mol−1)

χMeOH NMeOH NH2O exp simulation exp simulation


0.0 0 1000 0.99710 0.9970 ± 0.0001 0.0000 0.000 ± 0.000
0.1 100 900 0.97015 0.9656 ± 0.0014 −0.3174 −0.221 ± 0.026
0.2 200 800 0.94719 0.9428 ± 0.0005 −0.6166 −0.510 ± 0.017
0.3 300 700 0.92523 0.9204 ± 0.0000 −0.8441 −0.708 ± 0.007
0.4 400 600 0.90339 0.9014 ± 0.0007 −0.9758 −0.904 ± 0.014
0.5 500 500 0.88175 0.8820 ± 0.0003 −1.0083 −1.001 ± 0.020
0.6 600 400 0.86073 0.8638 ± 0.0002 −0.9509 −1.040 ± 0.016
0.7 700 300 0.84075 0.8437 ± 0.0004 −0.8176 −0.910 ± 0.005
0.8 800 200 0.82195 0.8256 ± 0.0007 −0.6189 −0.749 ± 0.019
0.9 900 100 0.80417 0.8077 ± 0.0010 −0.3535 −0.490 ± 0.060
1.0 1000 0 0.78690 0.7876 ± 0.0003 0.0000 0.000 ± 0.000

a
The results and uncertainties were obtained from two sets of uncorrelated calculations, as discussed in the text.

C DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

Figure 3. Partial molar volumes of water and methanol in their


mixtures obtained from NpT-MD simulations at T = 298 K, p = 1 atm
(dashed line and circles), experimental results17 (solid line), and
results from previous simulation work9 (dashed−dotted line). The
lines are given by eq 7, and the parameters are listed in Table 4.

Figure 5. Radial distribution functions of liquid methanol. Comparison


of the results obtained from the original PHH water model29 under
NEV conditions (dashed lines) with the modified model under NpT
(T = 298 K) conditions (solid lines).

■ RESULTS AND DISCUSSION


Excess Properties. The simulated properties of the pure
liquids in the NVE and NpT ensembles are given in Table 2.
The goal of having the correct experimental pressure is achieved
in both cases while the average potential energies of both systems
are lowered. The self-diffusion constant of water remains the
same, i.e., a bit too low, while that of methanol, already too high
compared to experiment,16 is increased by about 10%.
We present in Table 3 the densities and the excess volumes
Figure 4. Radial distribution functions of liquid water. Comparison of of the water−methanol mixtures as a function of the methanol
the results obtained from the original BJH water model28 under mole fraction χMeOH. The variation of these properties is also
NEV conditions (dashed lines) with the modified model under NpT shown in Figures 1 and 2 and compared with experimental
(T = 298 K) conditions (solid lines). results. The densities and their associated error bars have been
⎛ ⎞ ⎛ ⎞ obtained from the two independent 500 ps productions runs
∂V ∂Δv ⎟
vi = ⎜⎜ mix ⎟⎟ = vi(0) + Δv − χj ⎜⎜ ⎟
(as described in a previous section). The statistical errors
⎝ ∂ni ⎠T , P , n ≠ n ⎝ ∂χj ⎠T , P appear to be negligible, and the agreement with experiment
j i (6)
is fairly good as the maximum deviation is about 0.5%. The
where χi = ni/(n1 + n2) is the mole fraction of component i and calculated excess volume VE also agrees well with experiment,
Δv = ((VE)/(n1 + n2)). The following equation17 can be used and the maximum deviation with respect to ideal mixtures is
to fit the simulated volumes: about −1 cm3 mol−1. The extremum is slightly shifted toward
3 larger mole fractions. Figure 2 also shows the curve resulting
V E = χ1 χ2 ∑ Ak (χ2 − χ1 )k from a fit of simulated data using eq 7, and the corresponding
k=0 (7) optimized parameters are given in Table 4.
D DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

Figure 6. Selected radial distribution functions (left ordinate) and corresponding integrals (right ordinate) for methanol−water mixtures as a
function of the mole fraction of methanol. (Left) Top: gOwOw(r). Middle: gOmOm(r). Bottom: gOwOm(r). (Right) Top: gOwHw(r). Middle: gOmHm(r).
Bottom: gOwHm(r). Subscripts w and m denote water and methanol, respectively. For clarity, only the results for the systems with the following mole
fractions of methanol have been plotted: 0.1, 0.3, 0.5, 0.7, and 0.9. Because the sequence of the curves is monotonous, only the lowest mole fraction
of each set is indicated.

Using these parameters one can calculate the partial molar selected self and cross g functions in mixtures of various
volumes of water and methanol; the results are presented in compositions. We define as usual
Figure 3. As the densities of the pure liquids are very well re-
Nβ r
produced the calculated partial molar volumes of water (vH(0)2O = nαβ (r ) = 4π ∫0 r′2 gαβ (r′) dr′
18.07 cm3 mol−1) and methanol (v(0) 3
MeOH = 40.66 cm mol )
−1 V (8)
virtually coincide with experimental results. For both species
where Nβ is the number of β particles in the system and ⟨V⟩ is
the overall agreement is also good for intermediate mole
the average volume (in NpT simulations, otherwise the fixed
fractions. To the best of our knowledge the minimum in the
volume) of the simulation box. nαβ(r) is thus the total number of
partial volume of methanol at low concentration has been β-type neighbors around a particle of type α.
reproduced here for the first time. Because this feature depends The results presented in Figures 4 and 5 allow a comparison
on the slope of the excess volume curve, the present model with the original model, BJH and PHH, respectively, under
seems to be able to describe the subtle changes occurring in the NEV conditions, with the present modified versions under NpT
structure of the mixture when only a few alcohol molecules are conditions. In both systems gOO(r) and gOH(r) remain constant
mixed with water. and only very slight differences are observed in the gHwHw(r) and
Structure. The structures of the pure liquids and of the gMeMe(r) functions. The new parametrization obviously does
mixtures have been analyzed by computing radial distributions not alter the interactions responsible for the hydrogen bonding
functions (RDFs) together with the corresponding running yet modifies other parts of the potential enough to maintain a
coordination numbers. In Figure 4 we present the three water− reasonable pressure in the systems.
water partial RDFs in pure water, namely, gOwOw(r), gOwHw(r), The cross g functions in mixtures, presented in Figure 6,
and gHwHw(r), where subscript w denotes water (and later on show an increase in the height of the peaks as the mole fraction
subscript m denotes methanol). Figure 5 shows the correspond- of methanol increases. The positions of the first and second
ing three functions for pure methanol. Finally, Figure 6 shows peaks seem not to be affected by the change in composition in
E DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

Figure 7. Running coordination numbers (text and eq 8), nOwOw (solid


line and triangles), and nOwOm (dashed line and circles) as a function of
the mole fraction of methanol.

the system. The values of the integrals of the gOwOw(r) and


gOwOm(r) RDFs at the first minimum are shown in Figure 7.
The first peak in the gOwOw(r) functions becomes higher as
χMeOH increases, but the coordination number associated with
its integration decreases, which indicates the presence of
methanol molecules around water molecules even at very low
mole fractions. The “inversion” in the number of neighbors
occurs around χMeOH = 0.65, i.e., the number of methanol
molecules in the first solvation shell of water becomes larger
than the number of water molecules. These results show that Figure 8. Self-diffusion coefficient of water (top) and methanol
the correct evolution of the density of the system seen here (bottom) in water−methanol mixtures obtained from NpT simulations
results from a progressive mixing of the two species and not (dashed lines and circles) as a function of the mole fraction of
from a segregation process. methanol. Experimental data are from ref 16 (solid lines and triangles).
Dynamics. We have characterized the translational mobility
of the molecules by their self-diffusion constant D. This pro- Because these models have essentially been constructed to
perty was obtained from the mean square displacements using study spectroscopic properties, they are particularly well suited to
the Einstein relation.26 For the pure liquids under ambient con- study in detail the hydrogen bond networks and its evolution: the
ditions, the experimental self-diffusion constants of water and shifts of the vibrational OH-streching frequencies induced by
methanol are similar, namely, 2.3 × 10−9 and 2.4 × 10−9 m2 s−1, the presence of hydrogen bonds (or other perturbations) can be
respectively.16 In addition to other properties it is well known easily detected and correlated with structural features. We expect
that these values are difficult to reproduce by most potentials. that the combination of such studies with vibrational spectroscopy
(For water, see ref 43.) This remark applies in particular to the experiments will lead to an improved qualitative and quantitative
understanding of the peculiar properties of these systems.


PHH model because it makes use of only three interacting sites.
It has already been observed that the reduced steric hindrance
resulting from this choice, compared to an all-atom six-site AUTHOR INFORMATION
model, leads to an increase in the diffusion processes.44 Corresponding Author
The results are presented in Figure 8. In the mixtures, D goes *Phone: +33 (0)5 4000 2242. E-mail: jean-christophe.
through a minimum as a function of composition of both the soetens@u-bordeaux.fr.
water and methanol molecules. This feature is moderately well Notes
reproduced in our simulations for water; i.e., we see a minimum The authors declare no competing financial interest.
but not at the mole fraction where it is found experimentally. For
methanol, the simulated self-diffusion increases monotonously
with increasing mole fraction of MeOH and rises to a value that
■ ACKNOWLEDGMENTS
Computer time for this study was provided by MCIA
is too large for the pure system. The very slight minimum seen (Mésocentre de Calcul Intensif Aquitain), the public research
experimentally around χMeOH ≈ 0.3 does not appear.


HPC center in Aquitany, France.
CONCLUDING REMARKS
The modified versions of the BJH and PHH models presented
■ REFERENCES
(1) Dixit, S.; Crain, J.; Poon, W. C. K.; Finney, J. L.; Soper, A. K.
in this article demonstrate that it is possible within the frame- Molecular Segregation Observed in a Concentrated Alcohol-Water
work of pair interactions to describe the pure liquids and to Solution. Nature 2002, 416, 829−832.
reproduce the excess mixing and molar properties of mixtures. (2) Soper, A. K.; Dougan, L.; Crain, J.; Finney, J. L. Excess Entropy in
To the best of our knowledge, one of these properties, the Alcohol-Water Solutions: A Simple Clustering Explanation. J. Phys.
Chem. B 2006, 110, 3472−3476.
minimum in the partial volume of methanol at low concentra- (3) Pascal, T. A.; Goddard, W. A. Hydrophobic Segregation, Phase
tion in water, has been reproduced here for the first time. Only Transitions and the Anomalous Thermodynamics of Water/Methanol
minor changes in the model parameters were required to use Mixtures. J. Phys. Chem. B 2012, 116, 13905−13912.
these models in NpT simulations, thus retaining their intrinsic (4) Davis, J. G.; Gierszal, K. P.; Wang, P.; Ben-Amotz, D. Water
qualities such as the interactions responsible for the hydrogen Structural Transformation at Molecular Hydrophobic Interfaces.
bonding and the adjusted intramolecular parts. Nature 2012, 491, 582−585.

F DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry B Article

(5) Pálinkás, G.; Bakó, I.; Heinzinger, K.; Bopp, P. Molecular (26) Allen, M. P.; Tildesley, D. J. Computer Simulations of Liquids;
Dynamics Investigation of Inter- and Intramolecular Motion in Liquid Oxford Science Publications, 1987.
Methanol and Methanol-Water Mixtures. Mol. Phys. 1991, 73, 897− (27) Kong, C. Combining Rules for Intermolecular Potential
915. Parameters. II. Rules for the Lennard-Jones (12−6) Potential and
(6) Noskov, S. Y.; Kiselev, M. G.; Kolker, A. M.; Rode, B. M. the Morse Potential. J. Chem. Phys. 1973, 59, 2464.
Structure of Methanol-Methanol Associates in Dilute Methanol-Water (28) Bopp, P. A.; Jancsó, G.; Heinzinger, K. An Improved Potential
Mixtures from Molecular Dynamics Simulation. J. Mol. Liq. 2001, 91, for Non-Rigid Water Molecules in the Liquid Phase. Chem. Phys. Lett.
157−165. 1983, 98, 129−133.
(7) Kiselev, M.; Noskov, S.; Puhovski, Y.; Kerdcharoen, T.; (29) Pálinkás, G.; Hawlicka, E.; Heinzinger, K. A Molecular
Hannongbua, S. The Study of Hydrophobic Hydration in Supercritical Dynamics Study of Liquid Methanol with a Flexible Three-Site
Water-Methanol Mixtures. J. Mol. Graphics Modell. 2001, 19, 412−416. Model. J. Phys. Chem. 1987, 91, 4334−4341.
(8) Wensink, E. J. W.; Hoffmann, A. C.; van Maaren, P. J.; van der (30) Rybicki, M.; Hawlicka, E. Influence of Ions on Molecular
Spoel, D. Dynamic Properties of Water/Alcohol Mixtures Studied by Vibrations and Hydrogen Bonds in Methanol-Water Mixtures, MD
Computer Simulation. J. Chem. Phys. 2003, 119, 7308−7317. Simulation Study. J. Mol. Liq. 2014, 196, 300−307.
(9) González-Salgado, D.; Nezbeda, I. Excess Properties of Aqueous (31) Jancsó, G.; Bopp, P. A. The Dependence of the Internal
Mixtures of Methanol: Simulation versus Experiment. Fluid Phase Vibrational Frequencies of Liquid Water on Central Force Potentials.
Equilib. 2006, 240, 161−166. Z. Naturforsch. 1982, 38a, 206−213.
(10) Georgiev, G. M.; Vasilev, K.; Gyamchev, K. Hydrogen Bonds in (32) Lemberg, H. L.; Stillinger, F. H. Central Force Model for Liquid
Water-Methanol Mixture. Bulg. J. Phys. 2007, 34, 103−107. Water. J. Chem. Phys. 1975, 62, 1677.
(11) Vlček, L.; Nezbeda, I. Excess Properties of Aqueous Mixtures of (33) Rahman, A.; Stillinger, F. H.; Lemberg, H. L. Study of a Central
Methanol: Simple Models versus Experiment. J. Mol. Liq. 2007, 131− Force Model for Liquid Water by Molecular Dynamics. J. Chem. Phys.
132, 158−162. 1975, 63, 5223.
(12) Bakó, I.; Megyes, T.; Bálint, S.; Grósz, T.; Chihaia, V. Water- (34) Preisler, Z.; Bopp, P. A. J. Unpublished work.
Methanol Mixtures: Topology of Hydrogen Bonded Network. Phys. (35) Pusztai, L.; Pizio, O.; Sokolowski, S. Comparison of Interaction
Chem. Chem. Phys 2008, 10, 5004−11. Potentials of Liquid Water with Respect to their Consistency with
(13) Sentenac, P.; Bur, Y.; Rauzy, E.; Berro, C. Density of Methanol Neutron Diffraction Data of Pure Heavy Water. J. Chem. Phys. 2008,
+ Water between 250 and 440 K and up to 40 MPa and Vapor - Liquid 129, 184108.
Equilibria from 363 to 440 K. J. Chem. Eng. Data 1998, 43, 592−600. (36) Barker, J. A.; Watts, R. O. Monte Carlo Studies of the Dielectric
(14) Mikhail, S. Z.; Kimel, W. R. Densities and Viscosities of Properties of Water-Like Models. Mol. Phys. 1973, 26, 789.
Methanol-Water Mixtures. J. Chem. Eng. Data 1961, 6, 533−537. (37) Neumann, M. The Dielectric Constant of Water. Computer
(15) Kubota, H.; Tsuda, S.; Murata, M.; Yamamoto, T.; Tanaka, Y.; Simulation with the MCY Potential. J. Chem. Phys. 1985, 82, 5663.
Makita, T. Specific Volume and Viscosity of Methanol-Water Mixtures (38) Swop, W. C.; Andersen, H. C.; Berens, P. H.; Wilson, K. R. A
under High Pressure. Rev. Phys. Chem. Jpn. 1979, 49, 59−69. Computer Simulation Method for the Calculation of Equilibrium
(16) Derlacki, Z. J.; Easteal, A. J.; Edge, A. V. J.; Woolf, L. A.; Constants for the Formation of Physical Clusters of Molecules:
Roksandic, Z. Diffusion Coefficients of Methanol and Water and the Application to Small Water Clusters. J. Chem. Phys. 1982, 76, 637.
Mutual Diffusion Coefficient in Methanol-Water Solutions at 278 and (39) Soetens, J. C. Ph.D. Thesis, Université Henri Poincaré − Nancy
298 K. J. Phys. Chem. 1985, 89, 5318−5322. I, 1996.
(17) McGlashan, M. L.; Williamson, A. Isothermal Liquid-Vapor (40) Nosé, S. A Molecular Dynamics Method for Simulation in the
Equilibria for System Methanol-Water. J. Chem. Eng. Data 1976, 21, Canonical Ensemble. Mol. Phys. 1984, 52, 255.
196−199. (41) Nosé, S. A Unified Formulation of the Constant Temperature
(18) Lee, H.; Hong, W. H.; Kim, H. Excess Volumes of Binary and Molecular Dynamics Method. J. Chem. Phys. 1984, 81, 511.
Ternary Mixtures of Water, Methanol, and Ethylene Glycol. J. Chem. (42) Andersen, H. C. Molecular Dynamics Simulations at Constant
Eng. Data 1990, 35, 371−374. Pressure and/or Temperature. J. Comput. Phys. 1983, 52, 24.
(19) Xiao, C.; Bianchi, H.; Tremaine, P. R. Excess Molar Volumes (43) Guillot, B. A Reappraisal of What we Have Learnt During Three
and Densities of (Methanol+Water) at Temperatures between 323 Decades of Computer Simulations on Water. J. Mol. Liq. 2002, 101,
and 573 K and Pressures of 7.0 and 13.5 MPa. J. Chem. Thermodyn. 219.
1997, 29, 261−286. (44) Hawlicka, E.; Pálinkás, G.; Heinzinger, K. A Molecular
(20) Yilmaz, H.; Guler, S. Excess Properties of Methanol-Water Dynamics Study of Liquid Methanol with a Flexible Six-Site Model.
Binary System at Various Temperatures. Nuovo Cimento 1998, 20, Chem. Phys. Lett. 1989, 154, 255−259.
1853−1861.
(21) Takamuku, T.; Asato, M.; Matsumoto, M.; N, N. Structure of
Clusters in Methanol-Water Binary Solutions Studied by Mass
Spectrometry and X-Ray Diffraction. Z. Naturforsch. 2000, 55a,
513−525.
(22) Takamuku, T.; Saishoa, K.; Nozawa, S.; Yamaguchi, T. X-Ray
Diffraction Studies on Methanol-Water, Ethanol-Water, and 2-
Propanol-Pater Mixtures at Low Temperatures. J. Mol. Liq. 2005,
119, 133−146.
(23) Lalanne, P.; Andanson, J. M.; Soetens, J. C.; Tassaing, T.;
Danten, Y.; Besnard, M. Hydrogen Bonding in Supercritical Ethanol
Assessed by Infrared and Raman Spectroscopies. J. Phys. Chem. A
2004, 108, 3902.
(24) Andanson, J. M.; Soetens, J. C.; Tassaing, T.; Besnard, M.
Hydrogen-Bonding in Supercritical Tert-Butanol Assessed by Vibra-
tional Spectroscopies and Molecular Dynamics Simulations. J. Chem.
Phys. 2005, 122, 174512−1−9.
(25) Andanson, J. M.; Bopp, P.; Soetens, J. C. Relation Between
Hydrogen Bonding and Intramolecular Motions in Liquid and
Supercritical Methanol. J. Mol. Liq. 2006, 129, 101.

G DOI: 10.1021/acs.jpcb.5b03344
J. Phys. Chem. B XXXX, XXX, XXX−XXX

You might also like