Gubbins Molecular Theory
Gubbins Molecular Theory
Thermodynamics
Keith E. Gubbins
Department of Chemical Engineering
Cornell Universiiy
Ithaca, New York
I. Introduction
For the past 50 years, research in chemical engineering thermodynamics has
been largely concerned with problems related to the oil and petrochemical
industries. The primary methods of attack have been direct experiment and
macroscopic thermodynamic treatments of the data. Examples of the lat-
ter are empirical equations of state or activity coefficient expressions, group
contribution methods, and macroscopic corresponding states correlations.
These methods provide a convenient way to correlate large amounts of
experimental data, interpolate between different state conditions or between
different but structurally similar molecules, and make minor extrapolations
from experimentally measured conditions. They are most successful for fluids
of relatively simple molecules, such as the constituents of natural gas or
of low-molecular-weight hydrocarbon mixtures. However, they offer little
insight into the relation between the desired properties and the underlying
intermolecular forces or molecular structure and so are of little predictive
value. They will be much less useful for many of the new technologies, such
as the processing or design of electronic, photonic, or ceramic materials,
for biochemical processes, or for predicting the behavior of matter at and
near surfaces (e.g., in micelles, porous materials, or thin films).
125
Copyright 0 1991 by Audcmic Press. Inc.
ADVANCEBIN CHEMICAL ENGINERING. VOL. 16 All rights of reproduction in m y form rcmved.
126 Thermodynamics
D MACROSCOPIC
CORRELATIONS
Correlation
I
EXPERIMENT
Figure 1. The four methods for studying physical properties and what may be learned
from a comparison of any two of them.
time can be wasted. The same difficulty arises with testing the macroscopic
correlations. An example is provided by the large amount of work done on
molecular theories of liquid mixtures (various I-, 2-, and 3-fluid confor-
mal solution theories, cell theories, random mixture theory, etc.) in the period
1936-1970, before the first simulations of mixtures were reported. Although
comparisons of theory and experiment had indicated good agreement, com-
parisons with computer simulations showed that most of the theories were
quite inaccurate, in some cases not falling on the same piece of graph paper
as the experimental results. A second important use of simulations is to evalu-
ate intermolecular potential models by making comparisons with experi-
mental data; in this case, the statistical mechanics is exact in both simulation
and experiment, so the only source of error is the potential model assumed
in the simulation. In some important applications, computer simulation may
provide the best way to obtain a detailed understanding of the molecular
behavior because suitable experiments cannot be devised at the present time;
examples are the breakdown of some classical thermodynamic equations
for small drops, the study of phase transitions in narrow pores, and the
dynamics and thermodynamics of protein folding.
MACROSCOPIC APPROACH
I I I 1
UNBRIDLED BWR EOS LOCAL CORRESPONDING
EMPIRICISM COMPOSITIONS STATES
ACTIVlTY
COEFFICENT GROUP THEORY-BASED EOS
EQNS CONTRIBUTIONS
MOLECULAR APPROACH
I
I I I I
CELL PERTURBATION DENSITY FULL
MODELS THEORY FUNCTIONAL AB
THEORY INlTIO
KINETIC
THEORY
Figure 2. Examples of macroscopic and statistical mechanical methods for studying
physical properties. The degree of molecular basis increases from left to right.
lation. Examples are equations of state of the modified van der Waals type
(Redlich-Kwong, etc.), local composition methods (which attempt to account
for the fact that the composition around a molecule of a particular species
differs from the mean composition of the mixture), group contributions
(which approximates the intermolecular forces as a sum of group-group
interactions), and corresponding states. A common feature of all of these
methods is that the desired macroscopic properties are not related explic-
itly to some expression for the intermolecular forces, in contrast to the mo-
lecular theories described in the following section. Because of this, it is not
possible to test the macroscopic methods against simulation results, but only
against actual experimental data. Since much fitting to this data has often
been involved, such comparisons are usually a weak test.
Before leaving these methods, we should note that it is often possible to
make a compromise between the molecular and macroscopic approaches by
starting from a sound molecular theory and adopting approximations to obtain
a semiempirical correlation that does not involve the intermolecular potential
or molecular correlation functions. Many examples occur in the chemical
engineering literature, and I mention only one, taken from recent work by
Bryan and Prausnitz [ l ] on developing an equation of state for polar flu-
ids. They write the equation for the compressibility factor Z = PVIRT in
the usual way as Z = Z,,f + Zpert, but in place of the hard-sphere reference
term they take Zref to be the equation of state for polar hard spheres, for
which an accurate statistical mechanical theory (i.e.. one that agrees closely
with the computer simulation results for polar hard spheres) exists. Such
a reference fluid is much closer to the real fluid of interest than a hard-sphere
one, so that the perturbation term Zpea is much smaller. Bryan and Prausnitz
used the mean field term of van der Waals, Zpert = -a/RTv, where a is the
van der Waals attraction constant and v is the molar volume. This approach
gives good results for polar fluids (see Fig. 3), in contrast to the many at-
tempts to doctor the perturbation term while retaining the traditional hard-
sphere reference.
B . Molecular Theory
The configurational part (the part involving the intermolecular forces) of
the Helmholtz free energy for a system of N molecules in volume V at
temperature T is given by
7 I 1 1 8 1 1 I I I
6-
-Exp.
- .&C
5-
Acetonitrile 13
6
n -
z'a
-c
0-
-I - -
-2
I h i i'17 8 b i o i i i2
2 3
10001K-I
T
Figure 3. Calculated and observed vapor pressures for argon and several polar fluids
from an equation of state that uses a reference fluid of polar hard spheres, for
which an accurate statistical mechanical theory exists. Reprinted from Bryan
and Prausnitz [ 11 with the permission of Elsevier Science Publishers.
rj are over the volume V of the system. Here rj is the position of the cen-
ter of molecule i, and Oj (= 6 , p, for linear or 0, 6, for nonlinear mol- x
ecules) is its orientation relative to some space-fixed set of axes. Equation
(1) is valid for nonflexible molecules in which translational quantum ef-
fects are negligible. The integrations can be easily carried out for
noninteracting molecules (e.g., ideal gases), moderately dense gases, or
crystals, but for most other cases approximations are necessary. The prin-
cipal theories are [2] corresponding states theory, perturbation and cluster
expansions, density functional theory, lattice models, and integral equation
theory.
The first three of these theories are of particular interest in chemical en-
gineering. The molecular principle of corresponding states is based on the
idea that one can identify a group of substances, all of which obey a single
intermolecular potential law; they differ only in the values of the poten-
tial parameters. It provides the foundation for many existing correlations
of thermodynamic and transport properties, but these can be expected to
work only for groups of similar substances such as simple inorganics or low-
molecular-weight hydrocarbons.
Keith E. Gubbins 131
ence fluid can be obtained from experimental data (or from simulation data
for model fluids such as hard spheres) or corresponding states correlations,
while the perturbation corrections are calculated from the statistical mechani-
cal expressions, which involve only reference fluid properties and the per-
turbing potential. Cluster expansions involve a series in molecular clusters
and are closely related to the perturbation theories; they have proved par-
ticularly useful for moderately dense gases, dilute solutions, hydrogen-bonded
liquids, and ionic solutions.
Density functional theories [2,4] are similar in spirit to the perturbation
theories, but are of particular value for problems involving nonuniform
systems in which the density (or molecular orientation) varies with posi-
tion (direction) in the system-surface phenomena, solidification and melting,
thin films, liquid crystals, polydisperse systems, and so on. In this approach,
one starts from the fact that the free energy density a ( r ) at some point r
in the system is a functional of the density profile p(r'). Using variational
methods, one finds the global minimum of the free energy density with
respect to the density profile and so determines the density profile itself.
The success of the method hinges on the accuracy of the expression used
for the free energy density functional a ( r ) .Usually this is written as a sum
of repulsive and attractive force contributions, the two terms being treated
by different approximations. Several versions of the theory exist, differing
in the approximations used for the repulsive term in a ( r ) , and these have
been successfully applied to problems in adsorption, micelles, fluids near
charged walls, and melting. This approach is likely to prove valuable for
the study of many of the interfacial problems to be met in the new tech-
nologies associated with electronic and microstructured materials, thin films,
etc.
Lattice models were used extensively to describe fluids from the 1930s
to the 1970s, but really describe solids rather than fluids, and they have been
superseded for most applications by the more sophisticated theories described
above. An exception is the study of the critical point, a singular point in
the phase diagram where conventional mean field and perturbation theo-
ries fail. The Ising model, a lattice theory that can be solved essentially
exactly, can be used successfully in this region and, together with some
related cell theories that can be mapped onto the king model (e.g., the deco-
rated lattice gas model), gives valuable information on the equation of state
in the critical region.
132 Thermodynamics
Integral equation methods provide another approach, but their use is limited
to potential models that are usually too simple for engineering use and are
moreover numerically difficult to solve. They are useful in providing equa-
tions of state for certain simple reference fluids (e.g., hard spheres, dipo-
lar hard spheres, charged hard spheres) that can then be used in the
perturbation theories or density functional theories.
C. Computer Simulation
In molecular simulation [ 5 , 61, one starts from a molecular model and an
equation for the intermolecular forces and calculates the macroscopic prop-
erties by a numerical solution of the equations. Two methods are in com-
mon use, the Monte Carlo and molecular dynamics techniques (Fig. 4).The
Monte Carlo (MC) method makes use of a random number generator to
“move” the molecules in a random fashion. Statistical mechanics tells us
that, for a fixed temperature and density, the probability of a particular
arrangement of the molecules is proportional to exp(-UlkT), where U is the
total energy of the collection of molecules, k is the Boltzmann constant, and
T is the temperature. In MC, the random moves are accepted or rejected
according to a recipe that ensures that the various molecular arrangements
that are generated appear with probabilities given by this law. After gen-
erating a long series of such arrangements, they can be averaged to obtain
the various equilibrium properties of the system of molecules.
These two techniques have several features in common. Accurate results
can be expected, provided that the simulation runs are carried on long enough
and that the number of molecules is large enough. In practice, the results
are limited by the speed and storage capacity of current supercomputers.
Typically, the number of molecules in the sample simulated can range up
to a few thousand or tens of thousands; for small molecules, the real time
simulated in MD is of the order of a nanosecond.
In order to minimize boundary effects in such small samples, it is cus-
tomary to use periodic boundary conditions; that is, the sample is surrounded
on all sides by replicas of itself, so that when a molecule moves through a
boundary and so out of the sample, it is automatically replaced by a mol-
ecule moving into the sample through the opposite face of the box. (Any-
one who has played Pacman, Asteroids, or similar video games is familiar
with this periodic boundaries trick.) Although molecular simulation can be
successfully applied to a wide range of problems, difficulties can arise with
some applications because of storage or speed limitations. Examples include
ionic fluids, such as plasmas and electrolytes, in which the range of the
intermolecular forces is very large, necessitating a large number of mol-
ecules. Difficulties also arise with substances in which long-range fluctuations
Keith E. Gubbins 133
N - -
100 10,000
Periodic Boundaries
Prescribed Intermolecular Potential
1
Generate random moves
1
Solve Newton's equations
5 = miq
occur; these are associated with substances very near critical points and those
exhibiting certain surface phenomena. Long-time phenomena are also apt
to make simulation difficult.
There are not only common features but also significant differences
between the MC and MD methods. MC is easy to program and can be easily
adapted to different conditions; it is adaptable, for example, to mixture studies
at constant pressure or to adsorption at constant chemical potential. MD is
more difficult to program and, in its conventional form, energy must be
conserved, providing a less convenient set of state variables to work with
for some applications. However, it is now possible to overcome this prob-
lem and to carry out MD calculations at constant temperature or pressure
[ 5 , 61. MD has two important advantages over MC; it can be used to study
time-dependent phenomena and transport processes, and the molecular
motions are natural and therefore can be observed and photographed eas-
ily using computer graphics.
Several specialized simulation techniques exist for particular applications
[ 5 ] : nonequilibrium MD for the study of transport properties and nonlinear
response, Brownian dynamics for the study of large molecules (e.g., pro-
teins) in solution, and quantum simulations for the study of non-classical
fluids and solids. Simulated annealing is a Monte Carlo technique for
optimization subject to a set of constraints and is finding widespread use
in design of chemical processes, circuit design (especially VLSI), image
processing, and protein engineering (see Sect. IV A).
D. Computer Graphics
A single molecular simulation provides a vast amount of detailed informa-
tion-typically lo8 coordinate positions and an equal number of orienta-
tions and linear and angular velocities. Even after averaging to obtain
macroscopic properties, the amount of molecular data (spatial, angular, and
time correlation functions, diffusion rates, and so on) is difficult to assimilate
from graphs or tables. Computer graphics provides a way to present such
large arrays of data and is particularly useful in visualizing such physical
processes as nucleation and phase separation, molecular motion near a sur-
face, adsorption and hysteresis in porous materials, and the folding of pro-
teins. In the simulation of very large molecules or materials, computer
graphics can help decide which structures or motions are important and which
functions would best characterize them. The ability to rotate the figure and
to zoom onto regions of particular interest is a great help in viewing such
processes. Although computer graphics is a rapidly advancing area, several
problems remain. The graphics workstations commonly available to uni-
versity research groups are still too slow to represent molecular motions
or rotations of the system in real time for most problems, and one must be
Keith E. Gubbins 135
HCONH, 25 pyrrole 25
HCON(CH,), 25,100 pyridine 25
CH,CONHCH, 100 CH4 -161
CH,OH 25 C2H6 -89
C,H,OH 25 C3H8 -42, 25
,-C,H,OH 25 n-C4H 10 -0.5, 25
i-C,H,OH 25 I-C4H 25
t-C,H,OH 25 ,-'SH12 25
CH,SH 6 i-C5H12 25
C,H,SH 25 neo-CSHI , 25
(CH3)$ 25 C-CSHIO 25
C,H,SCH 25 n-C6H 14 25
(C,H,),S 25 CH,CH,CH=CH, 25
CH3SSCH3 25 I-CH,CH=CHCH, 25
(CH3)2O -25 c-CH,CH=CHCH, 25
C,H,OCH, 25 (CH,),C=CH, 25
(C2Hd20 25 benzene 25
THF 25 CH,CO,CH, 25
a From Jorgensen and Tirad+Rives [8].
Keith E. Gubbins 137
230
I95
160
u)
E! 125
0
90
55
20
20 55 90 125 160 195 230
EXPERIMENTAL
Figure 5. Comparisons of calculated and experimental volumes per molecule in A3 for
the liquids in Table 1 and TIP4P water. Calculated values are from computer
simulation using the OPLS potential method. Reprinted with permission from
W. L. Jorgensen and J. Tirado-Rives, J . Am. Chem. SOC.110, 1657 (1988) [8].
Copyright 1988 American Chemical Society.
138 Thermodynamics
15
12
v)
J 9
n
0
0
0 3 6 9 12 15 18
EXPER I M ENTAL
Figure 6. Comparison of calculated (computer simulation results for the OPLS model)
and experimental heats of vaporization in kcal mo1-l for the liquids in Table 1
and TIP4P water. Reprinted with permission from W. L. Jorgensen and J.
Tirado-Rives, J . Am. Chem. SOC.110, 1657 (1988) [8]. Copyright 1988
American Chemical Society.
bons, alcohols, amines, sulfur compounds, ethers, and so on, and has been
successfully applied to protein crystal structures and energy minimization
[81.
The models described so far use isotropic site-site interactions. They give
a good description of the molecular shape, but in most cases neglect the
rearrangement of the valence electrons that occurs on bonding. This shift
of the electrons into bonds, x orbitals, and lone pairs has a significant effect
on the intermolecular forces, and its description requires the use of aniso-
tropic site-site interactions [ 131. The electrostatic interactions between
molecules can be qu ite accurately described by using a series of sites within
the molecule (i.e., the nuclei of the atoms or CH, groups), each of which
interacts with sites on neighboring molecules with multipole forces (point
charge, dipole, quadrupole, and so on); an approach called distributed mul-
tipole analysis [13, 141 can be used to determine the best location of the
sites and the multipole moments needed from ab initio calculations. Such
Keith E. Gubbins 139
Continuous mixtures
Phase equilibria of mixtures
Polar and associating liquids
Electrolyte solutions
Solvation, folding of biological molecules
Gas hydrates (clathrates)
Polymers, advanced materials
Micelles, colloids, vesicles
Adsorption problems
Surfactants
Fluid behavior in porous materials
Nucleation
Thin films (Langmuir-Blodgett, etc.)
Chemical equilibria
Polydisperse fluids
Zeolites
140 Thermodynamics
geology, and the study of planetary atmospheres. Since the range of pos-
sible compositions, temperatures, and pressures that are met in practice is
enormous, it is not feasible to carry out experiments for more than a small
fraction of the systems of interest, and there is therefore much benefit to
be gained from developing computer simulation and theoretical prediction
methods. Some approximate estimates of the cost and time needed for such
calculations and experiments at the present time are shown in Table 3. Theo-
retical and empirical correlation methods are satisfactory only for rather
simple mixtures at the present time. For more complex fluids, simulation
or experiment is more reliable. Estimates of the cost and time for simula-
tions are strongly dependent on the complexity of the molecular model and
the accuracy desired and thus are difficult to make. At present, the simu-
lations are cheaper than experiments for simple mixtures but are still rela-
tively expensive for complex fluids, e.g., hydrogen-bonded ones; these costs
will decrease as faster machines become available. The simulations are in
general considerably faster than the experiments.
It is not straightforward to calculate phase equilibria or chemical poten-
tials in a simulation, and special techniques must be used. This has been
an active research area over the past few years [ 151. Several methods have
been proposed, and these can be divided into direct methods, in which the
coexistence properties of the phases are calculated directly, and indirect
methods, where the chemical potential is first calculated and then used to
determine the phase equilibrium conditions (Table 4).
The most straightforward direct method is to simulate a two-phase sys-
tem and allow it to equilibrate. This approach is valuable for studying the
properties of the interface but is less satisfactory for determining the prop-
erties of the bulk phases themselves because of slow diffusion across the
interface; the results are also sensitive to the interfacial area. An important
new development is the Gibbs ensemble Monte Carlo method [16], a di-
rect method that avoids the interfacial diffusion problem and is much faster
than simulating the two-phase system, especially for mixtures. The method
is illustrated in Fig. 7 and involves setting up two homogeneous phases
(I and 11) that are in thermodynamic equilibrium but not in physical con-
tact. Equilibration is achieved by allowing changes in the volumes and
number of molecules in each phase (keeping the total volume and number
of molecules for the two-phase system constant), together with the usual
Monte Carlo moves of molecules in each box to obtain equilibrium. Some
typical results for mixtures are shown in Fig. 8. Agreement with results from
other (indirect) methods is good, and the Gibbs method offers a great
improvement in speed because no interface is involved. Typically, for a binary
mixture of 600 molecules, the time required is five CPU minutes per mil-
Keith E. Gubbins 141
\KIc
0 . 0 0
E'+AE', E'+AE",
N1+ I , V ' N"- I,V"
Figure 7. Possible steps in the Gibbs method for simulating the properties of fluids. The
schematic illustrates the initial system configuration and three steps: (a)
particle displacement, (b) volume change, and ( c )particle transfer. Variables
are defined as in Fig. 4.Reprinted with permission from W. L. Jorgensen and
J. Tirado-Rives, J . Am. Chern. SOC.110, 1657 (1988) [8]. Copyright 1988
American Chemical Society.
i t permits large density fluctuations in the fluid. The test particle, grand
canonical, and Gibbs ensemble methods work well at moderate densities
but become difficult to use at high densities, e.g., a dense liquid near its
triple point or a solid. The reason is most easily seen for the test particle
method, since the random insertion of a test molecule in a dense fluid will
almost certainly result in molecular overlap and consequently a large positive
value of Ut and a very small contribution to the integral that determines
p.The same problem occurs in the Gibbs and grand canonical methods with
the addition of new molecules to the system. What one needs to do is develop
some way of guiding the test particles or molecules toward any hole that
may be present in the fluid or solid. Several "biased sampling" methods have
Keith E. Gubbins 143
I I I I I
0 0.2 0.4 0.6 0.8 1.0
x2
(b)
1201 I I I I 1
EX PER1MENT
I00
P/bar
80
60
40
20
0
0 0.2 0.4 0.6 0.8 1.0
x2
Figure 8. (a) Vapor-liquid coexistence curves for a mixture of Lennard-Jones molecules
with parameters chosen to approximate acetone ( I ) - carbon dioxide (2):
0 = Gibbs method at constant pressure; A = test particle method at constant
volume. Horizontal and vertical lines are error bars. (b) Experimental (x) and
empirical equation of state (-) results for acetone-carbon dioxide mixtures.
Reprinted with the permission of Taylor & Francis Ltd. from Panagiotopoulos
et al. [ 161; and with permission from A. Z. Panagiotopoulos, U. W. Suter, and
R. C. Reid, Ind. Eng. Chem. Fundam. 25,525 (1986). Copyright 1986
American Chemical Society.
144 Thermodynamics
been developed for doing this; they extend the density range in which these
methods can be used. However, they still usually fail for solids or dense
fluids of highly nonspherical molecules (e.g., liquid crystals). For these more
difficult situations, one must resort to thermodynamic integration.
0.4
L
1st. LAYER
0.2
d
o k I I I I
The principal tools have been density functional theory and computer
simulation, especially grand canonical Monte Carlo and molecular dynamics
[17-191. Typical phase diagrams for a simple Lennard-Jones fluid and for
a binary mixture of Lennard-Jones fluids confined within cylindrical pores
of various diameters are shown in Figs. 9 and 10, respectively. Also shown
in Fig. 10 is the vapor-liquid phase diagram for the bulk fluid (i.e., a pore
of infinite radius). In these examples, the walls are inert and exert only weak
forces on the molecules, which themselves interact weakly. Nevertheless,
146 Thermodynamics
'A Or YA
Figure 10. Vapor-liquid equilibria for an argon-krypton mixture (modeled as a Lennard-
Jones mixture) for the bulk fluid (R* = m) and for a cylindrical pore of radius
R* = R/o,, = 2.5. The dotted and dashed lines are from a crude form of
density functional theory (the local density approximation, LDA). The points
and solid lines are molecular dynamics results for the pore. Reprinted with
permission from W. L. Jorgensen and J. Tirado-Rives, J . Am. Chem. SOC.
110, 1657 (1988) [8]. Copyright 1988 American Chemical Society.
condensation occurs at pressures far below the vapor pressure of the bulk
fluid (capillary condensation), and the critical temperature and pressure are
reduced substantially. For pores of intermediate size, e.g., a radius of 7.00
(0 is the molecular diameter), there are so-called layering transitions; these
are first-order phase transitions between adsorbed layers of one molecular
thickness and two, between two molecular layers and three, and so on. Such
transitions depend strongly on the strength of the wall forces, the temperature,
and the pore radius; for activated carbon pores, which interact more strongly
with the fluid molecules, the layering transitions would be more pronounced
than those shown in Fig. 9. For very small pores, they are inhibited by
molecular packing effects, while in very large pores they have less effect
on the overall phase diagram because of the large amount of bulk fluid
Keith E. Gubbins 147
present. When the pore radius becomes very small, a cylindrical pore ap-
proaches a one-dimensional limit in which the fluid molecules can only move
along a line. Exact statistical mechanics tells us that such a system cannot
show any phase transitions. This seems to occur at a radius between l a and
20. For mixtures, confinement within a pore will lead to large shifts in relative
adsorption and volatility, in addition to these effects (see Fig. 10). Much
remains to be done to understand these complex phase diagrams and ad-
sorption effects in terms of the underlying intermolecular forces, particu-
larly for the more complex fluids and porous materials of technological
interest.
B . Teaching
These developments will call for a restructuring and rethinking of the teaching
of chemical engineering thermodynamics. Many widely used texts concentrate
exclusively on the classial approach, the word molecule never appearing
in the course. Statistical mechanics and quantum mechanics are usually taught
as part of the chemistry or physics sequence, but the student rarely if ever
sees any applications of this material in chemical engineering courses.
Implicitly, these subjects are treated like Greek mythology-part of the
student’s general education, but not of great importance in chemical engi-
neering. Such an approach may prevent our graduates from involvement in
important technologies of the future. We need to introduce examples into
our existing chemical engineering courses that apply the fundamental knowl-
edge of these topics that the student has gained in chemistry or physics. A
problem at present is the lack of suitable textbooks at the undergraduate
level. Books on statistical mechanics written by chemists usually do not
progress past applications to the ideal gas and other systems of noninteracting
particles, so that the student gets little feel for the relevance of the subject
152 Thermodynamics
Acknowledgments
I am grateful to many colleagues for helpful discussions and for putting up with many ques-
tions on areas outside my own direct experience. In particular, I thank F. H. Arnold, B. J.
Berne, J. C. G. Calado, E. A. Carter, P. T. Cummings, W. A. Goddard, W. L. Jorgensen, F.
Kohler, J. A. McCammon, A. Z. Panagiotopoulos, N. Quirke, H. A. Scheraga, W. C. Still,
and D. N. Theodorou. Part of this work was supported by grants from the National Science
Foundation and the Gas Research Institute.
References
1. Bryan, P. F., and Prausnitz, J. M., Fluid Phase Equilibria 38,201 (1987).
2. See, for example: Hansen, J. P., and McDonald, I. R., Theory of Simple Liquids, 2nd
Ed. Academic Press, London, 1986; Gray, C. G., and Gubbins, K. E., Theory of Mo-
lecular Fluids. Clarendon Press, Oxford, 1984; Lee, L. L., Molecular Thermodynam-
ics of Nonideal Fluids. Buttenvorths, Boston, 1987.
3. Chapman, W. G., Gubbins, K. E., J o s h , C. G., and Gray, C. G., Pure Appl. Chem. 59,
53 (1987).
4. Evans, R., Adv. Phys. 28,43 (1979).
5 . Allen, M. P., and Tildesley, D. J., Computer Simulation of Liquids. Clarendon Press,
Oxford, 1987.
6. Abraham, F. F., Adv. Phys. 35, 1 (1986).
7. Kataoka, Y., J. Chem. Phys. 87, 589 (1987).
8. Jorgensen, W. L., and Tirado-Rives, J., J. Am. Chem. SOC. 110, 1657 (1988), and ref-
erences therein.
9. Momany, F. A., McGuire, R. F., Burgess, A. W., and Scheraga, H. A., J. Phys. Chem.
79,2361 (1975); Nemethy, G., Pottle, M. S., and Scheraga, H. A., J. Phys. Chem. 87,
1883 (1983); Sippl, M. J., Nemethy, G., and Scheraga, H. A., J. Phys. Chem. 88,623 1
(1984).
10. Brooks, B. R., Bruccoleri, R.E., Olafson, B. D., States, D. J., Swaminathan, S., and
Karplus, M. J., J. Comp. Chem. 4, 187 (1983). See also Brooks et al. [30].
1 I. Weiner, S. J., Kollman, P. A., Nguyen, D. T., and Chase, D. A., J. Phys. Chem. 7,230
(1986).
Keith E. Gubbins 153
12. Dauber-Osguthorpe. P., Roberts, V. A., Osguthorpe, D. J., Wolff, J., Genest, M., and
Hagler, A. T., Proteins 4, 31 (1988).
13. Price, S. L., Mol. Simul. 1, 135 (1988), and references therein.
14. Stone, A. J., and Alderton, M., Mol. Phys. 56, 1047 (1988); Price, S. L., Stone, A. J.,
and Alderton, M., Mol. Phys. 52,987 (1984).
15. See Gubbins, K. E., Mol. Simul. 2,223 (1989), and references therein.
16. Panagiotopoulos, A. Z., Mol. Phys. 61.8 13 (1987); Panagiotopoulos, A. Z., Quirke, N.,
Stapleton, M., andTildesley, D. J., Mol. Phys. 63,527 (1988).
17. See, for example: Peterson, B. K., Gubbins, K. E., Heffelfinger, G. S., Marini Bettolo
Marconi, U., and van Swol, F., J. Chem. Phys. 88, 6487 (1988); Heffelfinger, G. S.,
Tan, Z., Gubbins, K. E., Marini Bettolo Marconi, U., and van Swol, F., Mol. Simul. 2,
393 (1989), and references therein.
18. Magda, J. J., Tirrell, M., and Davis, H. T., J. Chem. Phys. 83, 1888 (1985); Bitsanis,
I., Magda, J. J., Tirrell, M., and Davis, H. T., J. Chem. Phys. 87, 1733 (1987).
19. Walton, J. P. R. B., and Quirke, N. P., Mol. Simul. 2, 361 (1989).
20. National Research Council, Committee on Chemical Engineering Frontiers: Research
Needs and Opportunities. Frontiers in Chemical Engineering. Research Needs and
Opporrunities. National Academy Press, Washington, D. C., 1988; Krantz, W. B.,
Wasan, D. T., and Nerad, P. V., eds., Interfacial Phenomena in the New and Emerging
Technologies, Proceedings of workshop held at University of Colorado, May 29-3 1,
1986, National Science Foundation, Division of Engineering, Washington, D. C., 1987.
21. McCammon, J. A., Science 238,486 (1987).
22. See for example: Proceedings of the Conference on Industrial Applications of Molecu-
lar Simulation, Mol. Simul., Vol. 2 and 3 (1989).
23. For some recent work in this area see: Theodorou, D. N., and Suter, U. W., Macromol-
ecules 19, 139,379 (1986); Mansfield, K. F., and Theodorou, D. N., “Atomistic Simu-
lation of Glassy Polymer Surfaces and Glassy Polymer Solid Interfaces,” Annual
AIChE Meeting, Washington, D. C., Nov. 1988.
24. For an application to the properties of crystalline silicon see: Carr, R., and Parrinello,
M., Phys. Rev. Lett. 55,2471 (1985). For such treatments for a variety of materials, see:
Proceedings of the CCPS-CCP9 Conference on Computer Modeling of New Materi-
als, University of Bristol, UK, Jan. 4-6, 1989; Mol. Simul. 4, Nos. 1-3 (1989).
25. Carter, E. A., and Goddard, W. A., J. Chem. Phys. 88,3132 (1988); Carter, E. A., and
Goddard, W. A., J. Cutul. 112.80 (1988); Carter, E. A., and Goddard, W. A., Surface
Sci. 209, 243 (1988).
26. Carter, E. A., private communication (1988).
27. Guo, Y., Langlois, J.-M., and Goddard, W. A., Science 239,896 (1988); Chen, G., and
Goddard, W. A., Science 239,899 (1988). See also Science 242,31 (1988). This theory
explains high-T superconductivity in terms of magnetic interactions of electrons and
pairs of copper atoms, which leads to electron pairing; it predicts that the highest critical
temperature for the copper oxide superconductors under development is likely to be
about 225 K, about 100 K higher than it is now. The theory may also point the way to
improved superconductors based on materials other than copper oxide. For a discus-
sion of other theories, see: Emery, V. J., Physics Today Jan. 1989, pp. 5-26; Little, W.
A., Science 242, 1390 (1988).
154 Thermodynamics
28. For an example of the use of computer simulation to study micelles, see: Woods, M.
C., Haile, J. M., and O’Connell, J. P., J . Phys. Chem. 90, 1875 (1986). See also Davis,
H. T., in Perspectives in Chemical Engineering: Research and Education (C.K. Colton,
ed.), p. 169. Academic Press, San Diego, Calif., 1991 (Adv. Chem. Eng. 16).
29. Bash, P. A., Singh, U. C., Langridge, R., and Kollman, P. A., Science 236,564 (1987);
Kollman, P. A,, Ann. Rev. Phys. Chem. 38,303 (1987).
30. For a comparison of these methods, see: Hall, D., and Pavitt, N., J . Comp. Chem. 5,
441 (1984); also, Brooks, C. L., 111, Karplus, M., and Pettitt, B. M.,Adv. Chem. Phys.
71 (1988).
3 1. Wuthrich, K., NMR of Proteins and Nucleic Acids. Wiley, New York, 1986.
32. van Gunsteren, W. F., and Berendsen, H. J. C., J . Cornput.-Aided Molec. Des. 1, 171
(1987).
33. McCamrnon, J. A., and Harvey, S . C., Dynamics of Proteins and Nucleic Acids. Cam-
bridge University Press, New York, 1987.
34. Wong, C. F., and McCamrnon, J. A., J. Amer. Chem. SOC. 108,3830 (1986).
35. Kirkpatrick, S., Gelatt, C. D., and Vecchi, M. P., Science 220,67 1 (1983).
36. See, for example: Bringer, A. T., Kuriyan, J., and Karplus, M., Science 235,458 (1987).
37. See Section 7 of the Proceedings of the 9th IUPAC Conference on Chemical Thermo-
dynamics, Lisbon, July 1986. Published in: Pure Appl. Chem. 59 (1987).
38. Berne, B. J., and Thirumalai, D., Ann. Rev. Phys. Chem. 37,401 (1986).
39. The last of these conferences was held in Banff, Alberta, in May 1989.The conference
proceedings appeared in Fluid Phase Equilibria 52 (1 989).
40. The most recent of these conferences was held in Prague, Aug. 29-Sept. 2, 1988. The
proceedings appeared in Pure Appl. Chem. 61 (1989).