You are on page 1of 1

Quick links

A computational model of a small periodic box


Water Models containing 216 water molecules

Water molecular models are computational techniques that have


been developed to help discover the structure of water.

Parameterization of the water models


Water model descriptions
Water model properties
Aqueous solutions
The Lennard-Jones relationship

'with four parameters I can fit an elephant,


and with five I can make him wiggle his trunk'
Enrico Fermi as stated in [2438]

'although the model can be very accurate in many respects,


it should not be confused with the real liquid'
Nilsson, Schlesinger and Pettersson [2569]

For reviews of the development of water models, see [275,2537],


for a review of the development of analytical potential energy functions, see [2641], for a review of their use in
supercritical water see [433], for an appraisal of their accomplishments see [400], for a comparison of some important
nonpolarizable models see [1478], for a review of progress in ab initio methods see [1611], and for a discussion of their
use in biomolecular simulations see [3395]. It should be noted that most simulations using models can only investigate
structural changes and interactions over very short timescales, generally less than a few microseconds. Although
computers have become more powerful, simulations often involve more complicated calculations with greater numbers of
constituent molecules and exponentially growing interactions. Reliable experimental data are always to be preferred over
results obtained by simulation methods

Water models parameterization

Water models are helpful given the basis that if the (known but hypothetical) model (that is, computer water) can
successfully predict the physical properties of liquid water, then the (unknown) structure of liquid water is determined.
There is a trade-off between the computational complexity of the model and the system's size and complexity that can be
computed in a realistic period. Even as computing power increases considerably year on year, the limits imposed by the
system size, model complexity, and time restrictions are tested. Simple models can be used in large systems (>10,000
molecules) or for long simulation periods (>10 ns). In contrast, complex but more accurate models (particularly ab initio
methods) may only be used for relatively small systems (100-1000) and simulation times (≈ ps). The small scale and
short periods of such computations may lead to the, often ignored, possibility of artifacts. Generally, they cannot
appreciate any real similar or larger (longer) scale events and should only be used to predict such events with great
caution. At present, no water model can be used in simulations involving 100 nm cubes of water or for simulation periods
of milliseconds or more. Despite the tremendous investment in time and effort, extrapolated predictions from molecular
modeling should not be treated as equal to or superior to experimental evidence.

Models involve orienting electrostatic effects and Lennard-Jones sites that may or may not coincide with one or more of
the charged sites.f The Lennard-Jones interaction accounts for the size of the molecules. It is repulsive at short
distances,a ensuring that the structure does not completely collapse due to the electrostatic interactions. At intermediate
distances, it is significantly attractive but non-directional and competes with the directional attractive electrostatic
interactions. This competition ensures a tension between an expanded tetrahedral network and a collapsed non-
directional one (for example, similar to that found in liquid noble gases). Generally, each model is developed to fit well
with one particular physical structure, or parameter (for example, the density anomaly, radial distribution function or the
critical parameters). It comes as no surprise when a model developed to fit specific parameters gives good compliance
with these same parameters (for example, see [984]). It is also the case that, despite the heavy computational investment
in the calculations, the final agreement (or otherwise) with experimental data is often 'by eye' and not statistically tested
or checked for parametric sensitivity. Also, tests for 'fit' often seem to be completed with publication in mind rather than
rigor. Thus, many papers use the radial distribution fit with diffraction data as their 'gold standard', despite the major fitted
peaks (where the agreement looks so impressive 'by eye') being derived from the tetrahedral nature of water that is built
into every model, and overpowering any disagreement in the fine detail. In particular, the O···O radial distribution function
masks any underlying complexity and seems to be a poor discriminator between widely differently performing models
[1224, 3671]. Indeed, current X-ray and neutron diffraction data cannot distinguish between popular models [1579, 1624,
1667, 1757, 2090]. Also, unfortunately, the purity, isotopic mix, and perhaps even the ortho/para spin state present in real
water may cause difficulty over the choice of the value of the physical parameter [400], as models only use one isotopic
form, and ignore the spin state and the presence of other entities, such as ions. Also ignored, in all the models described
here, is the ease of proton hopping between water molecules (≈ ps) and the positive effect that this must have on cluster
stability and formation.

There is still disagreement over which value of some physical parameters to use such as for the dipole moment. Whether
model results agree with other physical properties of water then acts as proof (or otherwise) of their utility. A 15-
parameter model based on water monomer and dimer properties shows good agreement with the liquid-vapor
coexistence line (clearly influenced by monomer-dimer properties), the critical point, and the surface tension-temperature
curve [2334], but important dynamic properties remain to be determined [2333]. However, it does show surprisingly good
agreement with the unrelated temperature of maximum density. By and large, the more fitting parameters required by the
model (and some require over 50), the better the fit. Some models lack robustness due to their sensitivity to the precise
model parameters [206], the system size, or the calculation method [619, 649]. A study of the sensitivity of water's
behavior to changes in the parameters using the TIP4P model potential showed that σ, followed by the O-H bond length,
had significant effects on the density, enthalpy of vaporization and radial distribution function fits [494]. A separate
sensitivity analysis showed that the thermodynamic properties of water models were most sensitive to the van der Waals
repulsive, the short-range coulomb, and the polarization components of the potential [1042]. Classical force field
simulations have been shown to underestimate the anomalous fluctuations in water extending from very low to ambient
temperatures [3742]

It can be noted that a number of these models use water molecules with a wider (more tetrahedral) H-O-H angle and
longer H-O bond length than those expected of gaseous or liquid water, and indicative of the importance of including
parameters giving strong hydrogen-bonding [2339]. Water molecules in liquid water are all non-equivalent (differing in
their molecular orbitals, their precise geometry, and molecular vibrations; for an extreme case, see the water dimer) due
to their hydrogen-bonding status, which is influenced by the arrangement of the surrounding water molecules. Some
models are polarizable [867] to make some allowance for this,c,d whereas other simpler models try to reproduce
'average' structures. The Franzese-Stanley coarse grained model for hydration water predicts a phase transition
between two liquids with different densities and energies in the supercooled water region, ending in a liquid-liquid critical
point [3979].
[Back to Top ]

Water model descriptions Water models

A review listed 46 distinct models [400], so indirectly


indicating their lack of success in quantitatively reproducing
the properties of real water. They may, however, offer
helpful insight into water's behavior.

Some of the more successful simple models are opposite


with their parameters given below. Model types a, b, and c
are all planar, whereas type d is almost tetrahedral. The
mid-point site (M) in c, and the lone pair sites (L) in d are
labeled q2

Some models are most helpful in understanding the water


structure-property relationships. Good examples are the
simplified two- and three-dimensional Mercedes-Benz (MB)
water models, proposed initially in 1971. These interact with
other particles through Lennard-Jones interactions and
orientation-dependent hydrogen bonding interactions
through three radial arms arranged as in the MB logo
[4400].

Parameters for some molecular water models

Model Type σÅ6 ε kJ ˣ mol−1 6 l1 Å l2 Å q1 (e) q2 (e) θ° φ°


SSD [511] -8 3.016 15.319 - - - - 109.47 109.47
SPC [94] a 3.166 0.650 1.0000 - +0.410 -0.8200 109.47 -
SPC/E [3] a 3.166 0.650 1.0000 - +0.4238 -0.8476 109.47 -
SPC/HW (D2O) [220] a 3.166 0.650 1.0000 - +0.4350 -0.8700 109.47 -
SPC/Fw 2 [994] a 3.166 0.650 1.0120 - +0.410 -0.8200 113.24 -
TIP3P [180] a 3.15061 0.6364 0.9572 - +0.4170 -0.8340 104.52 -
TIP3P/Fw2 [994] a 3.1506 0.6368 0.9600 - +0.4170 -0.8340 104.5 -
iAMOEBA2 [2031] a 3.6453 0.8235 0.9584 - +0.29701 -0.59402 106.48 -
uAMOEBA2 [2401] a19 3.7553 0.5945 0.9499 - - 19 105.95 -
QCT[1251] a15 3.140 0.753 0.9614 - +0.6064 -1.2128 104.067 -
PPC 1, 2 [3] b 3.23400 0.6000 0.9430 0.06 +0.5170 -1.0340 106.00 127.00
TIP4P [180] c 3.15365 0.6480 0.9572 0.15 +0.5200 -1.0400 104.52 52.26
TIP4P-Ew [649] c 3.16435 0.680946 0.9572 0.125 +0.52422 -1.04844 104.52 52.26
TIP4P-FQ [197] c 3.15365 0.6480 0.9572 0.15 +0.631 -1.261 104.52 52.26
TIP4P/Ice [838] c 3.1668 0.8822 0.9572 0.1577 +0.5897 -1.1794 104.52 52.26
TIP4P/2005 [984] c 3.1589 0.7749 0.9572 0.1546 +0.5564 -1.1128 104.52 52.26
TIP4P/2005f [1765] c 3.1644 0.7749 0.9572 0.1546 +0.5564 -1.1128 104.52 52.26
TIP4P/ε [2444] c 3.165 0.7732 0.9572 0.105 +0.5270 -1.054 104.52 52.26
OPC [2168] c 3.1666 0.8903 0.8724 0.1594 +0.6791 -1.3582 103.6 51.8
OPC3 [2722] a 3.17427 0.68369 0.9789 - +0.4476 -0.8952 109.47 -
SWFLEX-AI 2 [201] c four terms used 0.9681 0.141,3 +0.6213 -1.2459 102.71 51.351
COS/G3 [704] 9 c 3.17459 0.9445 1.0000 0.15 +0.450672 -0.901344 109.47 -
COS/D2 [1617] 9 16 c 18 18 0.9572 0.2472 +0.57 -1.14 104.52 -
GCPM2 [859] 10 c 3.69 4,11 0.9146 4 0.9572 0.27 +0.6113 -1.2226 104.52 52.26
SWM4-NDP2 13 [933] c 3.18395 0.88257 0.9572 0.24034 0.55733 -1.11466 104.52 52.26
+0.584 -1.168
BK32[2080] c 17 17 0.975 0.2661
esu esu
104.52 -
0.247M -1.1334M 101.098
SWM62 13 [1999] c/d7 3.19833 0.67781 0.9572
0.315L
+0.53070 -0.1080L
104.52

ST2 [872] 12 d 3.10000 0.31694 1.0000 0.80 +0.24357 -0.24357 109.47 109.47
TIP5P [180] d 3.12000 0.6694 0.9572 0.70 +0.2410 -0.2410 104.52 109.47
TIP5P-Ew [619] d 3.097 0.7448 0.9572 0.70 +0.2410 -0.2410 104.52 109.47
-0.07358
TIP5P-2018 [3480] d 3.145 0.79 0.9572 0.70 +0.394137 qO 104.52 109.47
-0.641114
TIP7P [3644] d 20 3.156 0.6284 0.9572 0.41 +0.58014 -0.17724 104.52 109.47
TTM2-F [1027] 14 c five parameters used 0.9572 0.70 +0.574 -1.148 104.52 52.26
POL5/TZ 2 [256] d 2.9837 4 4 0.9572 0.5 varies 5 -0.42188 104.52 109.47
3.115OO 0.715OO 0.8892L -0.044L
Six-site [491] c/d7 0.673HH 0.115HH
0.980
0.230M
+0.477
-0.866M
108.00 111.00

1 Average value. 2 Polarizable models. 3 charge = -2.48856; 4 Buckingham potential a. This exponential potential presents a
more flexible (that is, softer) surface than the Lennard-Jones r−12 interaction; 5 with the charge on the oxygen atom; 6σ and ε
are Lennard-Jones parameters. The separation and depth of the potential energy minimum between two similar molecules
(equivalent to diameter). 7 has charges on the lone pair sites (L) as in model type d and the mid-point site (M) as in model type
c.8 has only a single, center of mass, interaction site with a tetrahedrally coordinated sticky potential that regulates the
tetrahedral coordination of neighboring molecules. 9 a polarization charge qpol (-8) is connected by a spring to site q2, the total
charge (qpol+q2) being given in the table as q2.10 The charges are smeared (that is, not point charges) using Gaussian
distributions with widths of 0.455 Å and 0.610 Å for q1 and q2 respectively.11 Zero potential position at 3.25 Å.. 12 This model
over-structures the water. 13 A 'Drude' particle carrying a negative charge -1.71636e is attached by a harmonic spring (4184 kJ
ˣ mol−1 Å−2 while the oxygen carries a charge +1.71636e. 14 Induced dipoles are placed on the atoms. The van der Waals
terms give a deeper, steeper, and more distant energy minimum (-1.187 kJ ˣ mol−1 at 3.726 Å) than the typical Lennard-Jones
potential. 15 Calculated electrostatics contains dipole, quadrupole, octupole, and hexadecapole terms. 16 The induced dipole
has a sublinear dependence on the electric field. 17 The Lennard-Jones function has an exponential form. 18 The Lennard-
Jones function has terms for O and H. 19 Calculated electrostatics contains dipole and quadrupole, terms for the oxygen atom.
20 plus two further sites halfway towards the H atoms with -0.45837 charge, The O atom has +0.11094

Some of the above values are varied slightly by different workers. Other workers use diffuse electron density [203] or
polarizable versions of the nonpolarizable models, using flexible bonding (for example, SWFLEX-AI), induced dipoles
(for example, [181]), energy optimization (for example, the TIP4P-FQ version of TIP4P) or movable charge (for example,
SWFLEX-AI), all of which generally give better fit but at a significantly increased computational cost [198]. A comparison
of eleven nonpolarizable classical water models (TIP3P, SPC, SPC/E, flexible SPC, TIP4P, TIP4P/Ice, TIP4P/2005,
flexible TIP4P/2005, TIP4P-Ew, TIP5P, TIP5P-Ew) has been made with their hydrogen bond networks [4193].
Polarization mutually strengthens the hydrogen bonding and partially compensates for the absence (except statistically)
of the known long-range interactions and the dependence of these models on short-ranged forces. Diffuse electron
density [203] varies the effective charges with distance. Such models generally perform better away from the ambient
conditions under which they are parameterized than the simpler models.

In some circumstances, coarse-grained models are helpful. These are computationally cheap for the number of water
molecules used and are particularly useful for biomolecular hydration. They suffer in being less useful as predictors.
However, some such models have shown promise, such as one that uses cluster units of four water molecules (an
incomplete water tetrahedron) [2656]. Another, using a so-called 'Quantum Drude Oscillator' consisting of a negative
charge bound to a positive center by a quantum harmonic oscillator, generates many-body polarization and dispersion
interactions [2720]. This model predicts density, surface tension, enthalpy of vaporization, and dielectric constant across
phases and at surfaces, demonstrating good transferability.

An eleven-parameter analytical model has been proposed that allows the calculation of properties of water without
running simulations. It involves cage structures, two-molecule hydrogen bonding, and Lennard-Jones interactions a
[3518]. This model appears similar to my cluster model.

Another modeling system is density functional theory (DFT), which describes the system in terms of its three-dimensional
electron density and involves an approximate solution to the Shrödinger equation of a many-body system. Although much
valuable and interesting work has been achieved and improvements are ongoing [3064, 4454], this methodology is still
incapable of accurately predicting the properties of liquid water over a range of conditions [2546].

A variety of methods include nuclear quantum effects in dynamics simulations by combining quantum Boltzmann
statistics with classical dynamics. These are computationally expensive methods that can be used to compare with
infrared spectra. A comparison of models has given the quasi-centroid molecular dynamics (QCMD) model as best for
reproducing fundamental transitions in the spectrum [3756]. [Back to Top ]

Water model properties

The calculated physical properties of some of the water models are given below.

Calculated physical properties of the water models

Average Density Expansion


Dielectric self-diffusion, 10−5
Model Dipole moment e constant 2
configurational maximum, coefficient,
cm /s energy, kJ ˣ mol−1 °C 10−4 °C−1
SSD 2.35 [511] 72 [511] 2.13 [511] -40.2 [511] -13 [511] -
SPC 2.27 [181] 65 [185] 3.85 [182] -41.0 [185] -45 [983] 7.3 [704] **
SPC/E 2.35 [3] 71 [3] 2.49 [182] -41.5 [3] -38 [183] 5.14 [994]
SPC/Fw 2.39 [994] 79.63 [994] 2.32 [994] - - 4.98 [994]
PPC 2.52 [3] 77 [3] 2.6 [3] -43.2 [3] +4 [184] -
TIP3P 2.35 [180] 82 [3] 5.19 [182] -41.1 [180] -91 [983] 9.2 [180]
TIP3P/Fw 2.57 [994] 193 [994] 3.53 [994] - - 7.81 [994]
iAMOEBA 2.78 [2031] 80.7 [2031] 2.54 [2031] - 4 [2031] 2.5 [2031]
uAMOEBA 2.80 [2401] 76.3 [2401] 2.41 [2401] - - 3.38 [2401]
QCT ** 1.85 [1251] - 1.5 [1251] -42.7 [1251] +10 [1251] 3.5 [1251]
TIP4P 2.18 [3,180] 53 a [3] 3.29 [182] -41.8 [180] -25 [180] 4.4 [180]
TIP4P-Ew 2.32 [649] 62.9 [649] 2.4 [649] -46.5 [649] +1[649] 3.1[649]
TIP4P-FQ 2.64[197] 79 [197] 1.93 [197] -41.4 [201] +7 [197] -
TIP4P/2005 2.305 [984] 60 [984] 2.08 [984] - +5 [984] 2.8 [984]
TIP4P/2005f 2.319 [1765] 55.3 [1765] 1.93 [1765 ] - +7 [1765 ] -
TIP4P/ε 2.4345 [2444] 78.3 [2444] 2.10 [2444] - +4 [2444] -
OPC 2.48 [2168] 78.4 [2168] 2.3 [2168] - -1 [2168] 2.7 [2168]
OPC3 2.43 [2722] 78.4 [2722] 2.3 [2722] - -13 [2722] 4.3 [2722]
SWFLEX-AI 2.69 [201] 116 [201] 3.66 [201] -41.7 [201] - -
COS/G3 ** 2.57 [704] 88 [704] 2.6 [704] -41.1 [704] -78[1939] 7.0 [704]
COS/D2 2.55 [1617] 78.9 [1617] 2.2 [1617] -41.8 [1617] - 4.9 [1617]
GCPM 2.723 [859] 84.3 [859] 2.26 [859] -44.8 [859] -13 [859] -
SWM4-NDP 2.461 [933] 79 [933] 2.33 [933] -41.5 [933] <-53 [1999] -
BK3 2.644 [2080] 79 [2080] 2.28 [2080] -43.32 [2080] +4 [2080] 3.01 [2080]
SWM6 2.431 [1999] 78.1 [1999] 2.14 [1999] -41.5 [1999] -48 [1999] -
TIP5P 2.29 [180] 81.5 [180] 2.62 [182] -41.3 [180] +4 [180] 6.3 [180]
TIP5P-Ew 2.29 [619] 92 [619] 2.8 [619] - +8 [619] 4.9[619]
TIP5P-2018 2.504 [3480] 127 [3480] 2.34 [3480] - +4 [3480] 4.2 [3480]
TTM2-F 2.67 [1027] 67.2 [1027] 1.4 [1027] -45.1 [1027] - -
TIP7P [3644] 2.378 76.098 2.3 - +4 2.319
POL5/TZ 2.712 [256] 98 [256] 1.81 [256] -41.5 [256] +25 [256] -
Six-site * 1.89 [491] 33 [491] - - +14 [491] 2.4 [491]
MB-pol [3133] - 68.4 2.3 - -10 3.5
Experimental 2.95 78.4 2.30 -41.5 [180] +3.984 2.53
All the data is at 25 °C and 1 atm, except * at 20 °C and ** at 27 °C.

Many of the data values given in the table vary significantly between different researchers (see, [185]). A comparison of
some of the properties of the gas phase dimers for various models is given [704]. b Generally, rigid models give
excessive stabilization of the dimer compared with polarizable models [1241]. As can be deduced from the data given
(and other data), although such simple models are of great utility, no universally applicable model can be identified at this
time. It should also be noted that many simulations are performed with just a few hundred water molecules within
rectangular periodic boxes no more than 2.5 nm along each edge for times equivalent to a few picoseconds; conditions
that reduce discovery of long-range effects and introduce artifacts. The use of cut-off lengths (even long ones) in the
intermolecular interactions may also introduce artifacts [761]. It should be noted that there is a strong correlation between
the length scale of any water structuring and the timescale which is required to see it. The predictive value of water
models has been questioned [202] and, even with current developments, their general application should be approached
with caution [203]. Errors found with rigid water models, 25 °C [1858]

The water molecule is a flexible molecule with % of the experimental value


electronic polarization and models that do not include Model Specific Shear Thermal
both of these characteristics together with their three heat, CP viscosity conductivity
body interactions are unlikely to be good predictors.
SPC 102 31 144
This can be seen from the table opposite where the
physical parameters determined from the best SPC/E 108 37 153
available rigid models are seen to be unpredictable TIP3P 107 36 146
and unreliable [1858]. Also, such models do somewhat TIP4P 118 47 135
worse as the temperature is lowered. TIP4P-Ew 115 64 147
TIP4P/2005 120 65 149
The agreement of the icosahedral cluster model of TIP5P 120 88 111
water with the O···O radial distribution function, and
TIP5P-Ew 141 91 102
the long-range structure apparent from X-ray
diffraction [1476], are in marked contrast to the use of many polarizable and nonpolarizable models for water, which do
not show any fine structure. The popular TIP4P model underestimates the tetrahedral nature of the water molecules'
environments, which explains its poor estimate of the relative permittivity (dielectric constant). It is, however, remarkably
good at qualitatively describing water's phase diagram [669], and this has been developed further in TIP4P/Ice [838] and
TIP4PQ/2005 [1895] where 16 CPU.years of computation was required. The SPC/E, PPC, and TIP4P, [3] and BSV, CC,
DC, SPC/E, and TIP4P, [93] models are reported as failing to describe the experimental O···O radial distribution function
properly. The TIP3P and SPC show particularly poor agreement. The TIP4P, SPC/E, and PPC show improved
agreement but the models TIP4P-FQ, and the increasingly used TIP5P, give further improvement [199] at an increased
computational cost. However different the model's O···O radial distribution functions are, they give similar numbers of
neighbors and hydrogen bonds. So the O···O radial distribution function is a poor discriminator for these factors [199] and
certainly should not be used as the sole comparator for a model. The first oxygen peak of many models' radial
distribution functions rise too fast, go too high, and then fall too fast to the first minimum [3524b].

The popular models SPC, SPC/E, TIP3P, and TIP4P produce a poor agreement with water's melting point (giving melting
points of 190 K, 215 K, 146 K, and 232 K respectively), and SPC, SPC/E, TIP3P, and TIP5P do not give ice1h as a stable
phase, replacing it with ice II [775] or improbable unrealistic crystal structures. Many models mistakenly predict
antiferroelectric character for the ordered phases of ice (for example, ice XI) [1051]. The popular SPC/E, TIP4P, and
TIP4P-Ew models also fail to predict correct critical data, vapor pressure, or second virial coefficients [1235]. The
commonly used SPC/E and TIP3P models have been proven unreliable even in the liquid phase [1706]. The SPC/E
model should not be used at high pressure [3956]. Different models also give dissimilar lowest energy structures for small
water clusters [857]. It is also true that models for liquid water, bearing little relationship to reality (for example, involving
only two dimensions or 8-molecule cubic arrangements), can be used to calculate similarly close results for a small
number of water's properties. Most models do not account for the predominantly pz2 character of the highest occupied
molecular orbital (HOMO; 1b1), or consequently water's large quadrupole [1731 ]. Nonpolarizable models have been
shown to be inherently unable to simultaneously predict certain physical properties, such as melting temperature and the
temperature of maximum density, whatever parameter values are chosen, due to the limited number of variable
parameters [1079, 2159, 3521]. It has been proposed that the model [2722] represents the accuracy limit of rigid 3-point
water models, but is still somewhat inaccurate. Nuclear quantum effects are rarely included explicitly but give rise to high
proton swapping between neighboring water molecules plus strengthening of hydrogen bonds at low temperatures and
their weakening at high temperatures [2286]. None of 40 rigid, flexible, polarizable and ab initio models could
simultaneously agree with both the experimental radial distribution model and the experimental internal energy [245].
Serious discrepancies concerning the first coordination shell hydrogen bonding, have been noted between the molecular
dynamics simulations using these models and X-ray absorption spectroscopy [613]. The computational costs of these
models have been compared [2159]. A model that shows some agreement over the range of physical states of ice and
water, including correctly reproducing the temperature dependence, if not the accurate values, of many structural,
thermodynamic, and dynamical properties, is the MB-pol many-body potential for water [3133].

Other modeling studies have failed to reproduce parts of water's vibrational spectrum even qualitatively [696]. Artifacts,
such as unnatural phase transitions, may be unexpectedly produced in water simulations [1056] or, worse, go unnoticed.
An improved agreement may require many-body parameters [221, 3894] as three-body effects have been shown to
contribute 14.5% (or more [728]) to the internal energy, and these cannot be adequately represented by potentials that
distort two-body effects [465]. Even good models are incapable of describing the situation in deeply supercooled water,
with none (up until 2018) able to reproduce the sharpness of the experimentally observed compressibility maximum
[3502]. The magnitude of experimental density fluctuations is severely underestimated in the simulations. However,a
model (GCPM ) using polarizable smeared charges, rather than the other models' point charges, has shown considerable
promise [859]. Other steps forward are the inclusion of quantization, which is shown to have significant consequences on
water's structuring [863], and the need for high order multipole components [3160], up to hexadecapole, to achieve the
correct ferroelectric structures for the ordered ice phases [1051]. A study attempting to combine diffraction, infrared and
x-ray absorption data, concluded that current water models show poor fit [1159]. It is also clear that care must be taken
when combining quantum chemical modeling (involving a small number of molecules) with a larger scale, but
computationally cheaper, empirical models as some, well-parameterized models (such as TIP5P) may be unsuitable for
mixed simulations [1643]. The best water models for liquid water have a charge distribution with a large dipole, a large
quadrupole, and a negative charge out of the molecular plane to give symmetrically ordered tetrahedral hydration [2418].
There may be a fundamental flaw in all these modeling studies (except ab initio) as different parameters are required for
different purposes. Thus, optimal parametric values used to describe the potential energy surface are not necessarily the
best to describe the dipole moment surface or the polarizability surface, which will be particularly relevant when modeling
water in the presence of electric fields [2340]. Altogether, it is clear that despite water appearing to be a very simple
molecule, it remains challenging to model realistically, probably because it does not consist of totally independent
molecules held together by weak electrostatic bonds (as modeled) but in 'flickering' clusters and microdomains. One of
the areas most studied, due to the practical difficulties, is within supercooled water. Unfortunately, this area is not well
modeled [4053]. At present, the best nonpolarizable model appears to be the point-charge model OPC [2168], and the
best polarizable models are iAMOEBA [2031] and BK3 [2080], which both seem to show good promise. Improvements to
existing models and novel optimization methods are being developed [2299]. A promising polarizable force field for water
(RexPoN) has been developed based solely on quantum mechanics calculations with no empirical data [3524]. Each
atom has an electronic core and a responding shell, and the water structure contains multi-branched polymer chains. A
many-body polarized (MB-pol) potential has emerged (2016) as a versatile molecular model for water simulations from
the gas to the condensed phase, and captures significant aspects of liquid water’s behavior over a range of temperatures
from the boiling point to deeply supercooled conditions at atmospheric pressure [4130, 4224]. The lack of success of
otherwise successful models within the supercooled regime has been used to support the 'two kinds of water' and liquid-
liquid critical point hypotheses [4228].

Some models have been developed to give particular physical properties. For example, the TIP4P/2005 model was
developed to reproduces the maximum density at T = 4 °C, whereas for TIP4P/ϵ, the model not only reproduces this
density but also the dielectric constant at T = 4 °C. The density and diffusion anomalies for the TIP4P/2005 and TIP4P/ϵ
models have been determined [3344]. Other properties of these models are not as well predicted. Artificial intelligence
and machine learning have been used to optimize water models using a selection of water's properties [4345]. Such
methods work well for the properties chosen but less well for other properties. The poor prediction (of known parameters)
of these models should be taken into account when such models are used to predict unknown parameters or processes
[2887]. This may be the cause of their failure to describe the experimentally suggested two-state mixture model. This has
been attributed to the underlying potential energy surface on which the simulations move being insufficiently corrugated,
leading to an averaged structure rather than a mixture of two families of structures [2189, 3607].a It has been claimed
that models generated for use within a zero electric field may fail when an electric field is present [2340].

In the light of these observations, it is unsurprising that contemporary water models are relatively poor predictors for the
conformation and hydration of biological molecules in solution (for example, [596, 2279]), and it may be helpful to develop
water models specifically for use with biomolecular solvation. The Franzese-Stanley model (FS) for hydration water
coarse-grains the configurations of water molecules and describes the water hydrogen bonding using two terms,
accounting for (i) the directional and (ii) the cooperative components of the hydrogen bond interaction. Interestingly, the
FS coarse-grained model predicts the two-state phase transition in supercooled water and the liquid-liquid critical point.
[3979]

Aqueous solutions

Although aqueous electrolytes are possibly the simplest of solutions, it has been proven difficult to model them well
[2502]. Perhaps this is expected due to the problems involved in modeling pure water (see above).
[Back to Top ]

Footnotes
a The Lennard-Jones relationship. The Lennard-Jones relationship [3808] is a useful isotropic approximation for the
energy of the non-bonding interaction between atoms or molecules. Note, however, that water is not a spherically
symmetrical molecule as judged by the variation in the van der Waals radii [206]. Also, in these models, the Lennard-
Jones interaction exerts a repulsive effect on hydrogen bonding, whereas some report that it is attractive [548] even at
this close contact. Also, in use, its effective distance is often truncated. The Lennard-Jones potential is made up of a
twelfth power repulsive term and a sixth power attractive term (rij = distance apart of the ith and jth atoms, with σ and ε
defined below):

Sometimes, this is written

where A=4εσ12, B=4εσ6, and

It is likely that the repulsive term is too repulsive, and in reality, the repulsion is somewhat softer, allowing somewhat
easier close molecular contact [1245].

The Lennard-Jones potential for the SPC/E model (solid red line) Lennard-Jones potential for the SPC/E model
is shown right. The σ parameter gives the molecular separation
for zero interaction energy. The minimum energy (-ε) lies 12.25 %
further at σ x 21/6 Å. ε is the energy-well depth and a measure of
how strongly the two particles attract each other.

Also shown (dotted blue line) is an equivalent Buckingham


potential (σ 3.55 Å, ε 0.65 kJ ˣ mol−1, γ 12.75); the σ parameter in
the Buckingham potential gives the σ x 21/6 position in the
Lennard-Jones potential.

[859] [Back]

b Models may be checked for agreement with gas-phase clusters (for example, water dimers) before use in liquid water
simulations. However, such compliance should not be a necessary prerequisite for accurate liquid water predictions as
they tend to be biased towards internal hydrogen bond maximization, and surface unconnected ('dangling') hydrogen-
bonding capability minimization, due to their relatively large surface area. Thus, they are not representative of real bulk
liquid water structuring. [Back]

c Molecular polarization may be electronic (caused by the redistribution of its electrons), geometric (caused by changes
in the bond lengths and angles), or orientational (caused by the rotation of the whole molecule) [867]. Polarization is an
effective way to include many-body effects. This paper [867] describes Charge-On-Spring polarizable force fields (for
example, COS/G3) as most suitable for aqueous solutions of proteins (although such COS models are very poor on
physical properties such as freezing point [1952]). Alternatively, a model possessing out-of-plane polarization and
fluctuating charges (POL5/TZ) is proposed best for comparison with experimental vibrational data [878]. [Back]

d One model describes the water molecule solely in terms of dipoles and polarizabilities on the atoms and a quadrupole
on the oxygen atom [736]. [Back]

e As well as the charge (q) and dipole (µ), it may be that the quadrupole (Θ), octupole (Ω), and hexadecapole (H)

interactions are also very important [1228]. The interactions of these multipoles are q1ˣ q2/r, q1ˣ µ2/r2, q1ˣ Θ2/r3, q1ˣ
Ω2/r4, q1ˣ H2/r5, µ1ˣ µ2/r3, µ1ˣ Θ2/r4, µ1ˣ Ω2/r5, Θ1ˣ Θ2/r5, etc., so lessening in importance with distance (r) [2775]. These
multipole moments of the models are generally far lower than the calculated values for liquid water.

Calculated multipole moments for some water models

Quadrupole moments1 Octupole moments2


Model
Θxx, D Å Θyy, D Å Θzz, DÅ Ωxxz, D Å2 Ωyyz, D Å2 Ωzzz, D Å2
SPC/E -2.71 0.00 -1.36 -1.39 0.00 -0.55
TIP3P -2.30 0.00 -1.38 -1.19 0.00 -0.57
TIP4P -2.86 0.00 -1.60 -1.49 0.00 -0.70
TIP5P -1.33 0.76 -0.42 -0.69 -0.09 -0.57
Other -4.27 [453] -7.99 [453] -5.94 [453] -1.754 [452] -0.554 [452] -1.981 [452]
1 Directions as given elsewhere, zero position at the oxygen atom; Θxx= Σi cixi2 where c = charge, x = distance in x-direction
and the summation is over all (i) charges.
2 Directions as given elsewhere, zero position at the center of mass; Ωxxz= Σi cixi2zi where c = charge, x and z = distance in x-
and z-directions and the summation is over all (i) charges. [Back]

f As an example, the TIP3P model of n atoms makes use of the following equation

where ke is the electrostatic constant, rij= distance apart of the ith and jth atoms, with σ and ε defined above. [Back]

g An 'averaged' structure may not exist at all. Consider an 'average' person; neither man nor woman. [Back]

Home | Site Index | Icosahedral water clusters | The evidence | LSBU | Top

This page was established in 2001 and last updated by Martin Chaplin on 1 June, 2022

This work is licensed under a Creative Commons Attribution


-Noncommercial-No Derivative Works 2.0 UK: England & Wales License

You might also like