Sie sind auf Seite 1von 9

View Article Online / Journal Homepage / Table of Contents for this issue

PCCP
INVITED ARTICLE
The hydrophobic eect

B. Widom, P. Bhimalapuram and Kenichiro Kogay

Department of Chemistry, Baker Laboratory, Cornell University, Ithaca,


New York 14853-1301, USA

Received 10th April 2003, Accepted 4th June 2003


First published as an Advance Article on the web 24th June 2003
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

The thermodynamics of the hydrophobic eect, as measured primarily through the temperature
dependence of solubility, is reviewed, and then a class of models that incorporate the basic mechanism of
hydrophobicity is described. These models predict a quantitative relation between the free energy of
hydrophobic hydration and the strength of the solvent-mediated attraction between pairs of solute molecules. It
is remarked that the free energy of attraction being just of the order of the thermal energy kT may be important
for the eective operation of the hydrophobic eect in proteins. Deviations from pairwise additivity of
hydrophobic forces are also briey discussed.

I. Introduction in the solvent around each solute molecule. This is the phe-
nomenon of hydrophobic hydration. The total volume of
Hydrocarbons are only slightly soluble in water: they are hydro- solvent so aected by a pair of solute molecules is less when
phobic. The accommodation of a hydrocarbon molecule in the two are close together than when they are far apart, as illu-
water is accompanied by an increase in an associated free strated schematically in Fig. 2. The result is an eective,
energy. Hydrocarbons are not the only hydrophobes but they solvent-mediated attraction between the two. This is the
are typical of the class. Characteristically, their solubility dec- hydrophobic attraction.
reases with increasing temperature at low temperatures, which These eects have long been recognized to be important in
provides an important clue to the mechanism of hydrophobi- physical chemistry and biochemistry. The subject thus has an
city. At higher temperatures the solubility, after reaching a min- enormous literature, ranging from works that are now clas-
imum, often then increases with further increase of temperature. sic217 to those more nearly current,1885 many of these quite
These eects are illustrated in Fig. 1, which shows the Ost- sophisticated. A recent authoritative assessment of the status
wald absorption coecient S of methane in water (the ratio of the eld with emphases dierent from those in the present
of the number density of dissolved methane to that in the equi- account is in a review by Pratt.29 A beautiful earlier review
librium vapor), as a function of temperature.1 This is at a par- by Scheraga86 with an account of experimental results and
tial pressure of methane of 1 atm, although there is almost no emphasis on the role of hydrophobicity in biochemistry,
dependence on the pressure as long as the concentration of should also be noted.
methane in both phases is low (Henrys law). The thermodynamics of transfer of a molecule from one
The unfavorable free energy change accompanying the dis- phase to another is outlined in Section II, and then the thermo-
solution of the hydrocarbon results from structural changes dynamics of hydrophobic hydration as inferred from solubility
measurements such as those in Fig. 1 is presented in Section
III. What is seen there, among other principles, is that the dis-
solution of a hydrophobe in water is energetically favorable,

Fig. 1 Ostwald absorption coecient S of methane as a function of Fig. 2 Two hydrophobic molecules, (a) far apart, and (b) close
temperature T (from compilation of Battino1). together. The regions within the dashed curves represent schematically
the volumes of solvent that are signicantly aected by the presence of
y Permanent address: Department of Chemistry, Okayama University, the solutes. The total volume so aected by the pair is smaller in (b)
3-1-1 Tsushimanaka, Okayama 700-8530, Japan. than in (a).

DOI: 10.1039/b304038k Phys. Chem. Chem. Phys., 2003, 5, 30853093 3085


This journal is # The Owner Societies 2003
View Article Online

perhaps contrary to nave expectation, but is entropically unfa- either phase is dilute in A the subtracted term accounts for the
vorable to such an extent that the equilibrium solubility is very whole of the dependence of DFV and DGp on the concentration
low. This was the historic observation that led to our present of A in that phase, and the resulting DFV and DGp are then
understanding of hydrophobicity and its implications.2 Those independent of that concentration. This remains true, and
two sections are an expanded version of an earlier treatment;87 DFV and DGp remain nite, even when cbA ! 0; i.e., in the limit
the principles are mostly well known,11 but it is useful to bring in which there is no A in the b phase prior to the transfer; while
them together here for reference. the original DFV and DGp diverge to 1 in that limit. When
Section IV describes a class of models that incorporate the the phases are not dilute in A there is a residual dependence
main mechanism of hydrophobicity as revealed by the thermo- of DFV and DGp on the concentrations, reecting the fact that
dynamic studies. The models make detailed predictions about the other A molecules are then a signicant part of the envir-
hydrophobic attraction including a quantitative connection onment from or to which the contemplated transfer of the A
between the strength of the attraction and the free energy molecule occurs.
of hydrophobic hydration. These models, too, have been Let DSV and DSp be the corresponding changes in the com-
described earlier;87 the present account includes some more posite systems entropy in the two dierent transfer processes,
recent results. and dene
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

In Section V the geometric idea expressed in Fig. 2 is


extended to the simultaneous interaction of three solute cbA
DSV DSV k ln 5
molecules, from which it is possible to make some speculative caA
inferences about the three-body contribution to the solvent-
mediated interaction energy. cbA
DSp DSp k ln : 6
The main results are briey summarized in the concluding caA
Section VI
If either phase is dilute in A there is no dependence of DSV and
DSp on the concentration of A in that phase, but otherwise
there is. Again, these remain nite in the limit cbA ! 0, while
II. Thermodynamics of transfer
DSV and DSp would be innite.
We ask here for the changes in various thermodynamic quan- Now consider rst the xed-volume transfer. Each number
tities accompanying the transfer of a molecule A from a phase density cA is NA/V with NA the number of A molecules in
a to a phase b. Each phase separately is in complete internal the phase and V its volume. During the transfer, NA increases
equilibrium both before and after the transfer but the two or decreases by 1 (in the b or a phase, respectively), while V is
phases need not be in equilibrium with each other. Later A will xed. Then from (1), (3), and (5),
be a hydrophobic molecule, a some reference phase containing @DFV @DFV
A, and b water or an aqueous solution, but for now A, a, and b DSV  ; DSV  7
@T @T
are general.
The cases of most interest are those in which the tempera- where the temperature dierentiations are at xed numbers of
tures T of the two phases are equal and are the same before molecules of all the substances in each phase and at xed
and after the transfer (isothermal transfer). We will specify volumes of both phases. The change in the composite systems
that, and then contemplate either of two further circumstances energy, DEV , accompanying the constant-volume transfer, is
in which the transfer is made: that in which the volume of each
phase separately is the same before and after the transfer, or DE V DF V TDS V DF V TDSV 8
that in which the pressure of each phase separately is the same @DFV =T @DFV =T
before and after. : 9
@1=T @1=T
Let maA and mbA be the chemical potentials of A in the respec-
tive phases. In the transfer with the volumes xed, the dier- The constant-volume heat capacity CV of the two phases
ence mbA  maA is the net change in the Helmholtz free energy together also changes as a result of the transfer, the net change
of the two phases together; call it DFV : being
DF V mbA  maA : 1 @DEV @DSV @DSV
DCV V T T 10
@T @T @T
In the transfer with xed pressures, the same dierence
mbA  maA is the change in the composite systems Gibbs free @ 2 DFV @ 2 DFV
T T ; 11
energy; call it DGp : @T 2 @T 2
DGp mbA  maA : 2 where the second subscript V in (DCV)V species xed-volume
transfer.
Note that DFV DGp , but they are the DF and DG of dierent In the transfer at xed pressure the quantities of most inter-
processes. est, besides DGp and DSp , are the change DHp in the composite
Signicant parts of DFV and DGp result merely from the dif- systems enthalpy and the change (DCp)p in its constant-
fering concentrations of A in the two phases, but we wish to pressure heat capacity; but now the relations analogous to
know how much of DFV and DGp reect only the intrinsic ener- (7)(11) are slightly more complicated because the volumes V
getic and structural changes that occur in the phases as a result in cA NA/V are no longer xed when the temperature
of the transfer of the A molecule from one to the other. We changes at xed pressure and composition. Let ea and eb be
thus dene the coecients of thermal expansion of the two phases. Then
cbA from (1), (4), and (6), the analog of (7) is
DFV DFV  kT ln 3
caA @DGp @DGp
DSp  ; DSp  kTeb  ea ; 12
@T @T
cb
DGp DGp  kT ln A 4 where now the temperature dierentiations are at xed num-
caA
bers of molecules of all the substances in each phase and at
where k is Boltzmanns constant and caA and cbA are the number xed pressure of each phase. With the same understanding
densities of A in the respective phases; and now DFV DGp . If of what is now held xed in the temperature dierentiations,

3086 Phys. Chem. Chem. Phys., 2003, 5, 30853093


View Article Online

the analogs of (8)(11) are are those in which b is a condensed phase dilute in A (or from
which A is absent), while (i) a is a dilute gas containing A; or
DH p DGp TDS p DGp TDSp 13 (ii) a is a condensed phase dilute in A; or (iii) a is a condensed
@DGp =T @DGp =T phase that is pure or nearly pure A.
kT 2 eb  ea 14 In these situations the temperature derivatives of S may be
@1=T @1=T written as total derivatives, d/dT, and the thermodynamic
@ 2 DGp quantities of interest are obtainable from S, i.e., from the equi-
DCp p T librium partition coecient, by the formulas
@T 2
@ 2 DGp @eb  ea DFV DGp
T 2
2kTeb  ea kT 2 : 15  ln S 18
@T @T kT kT
A special case of the transfer process is that in which the DSV d DSp d
T ln S; T ln S Teb  ea 19
two phases are in equilibrium with respect to the transfer of k dT k dT
A. This is when maA mbA , so that DFV DGp 0. Then from DEV d DHp d
(1) and (2), T ln S; T ln S Teb  ea 20
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

kT dT kT dT
DF V DG p kT ln S equilibrium transfer; 16
DCV V d2 T ln S
T 21
where k dT 2

S cbA =caA eq ; 17
DCp p d2 T ln S b a
b a
2 @e  e
T 2Te  e T 22
k dT 2 @T
the ratio of the number densities of A in the two phases when as now follow from the more general formulas stated earlier.
they are in equilibrium with each other. This is the key to Applications of these formulas to the transfer of gaseous
obtaining the thermodynamics of transfer from measurements hydrocarbons into water are described in the next section,
of equilibrium solubility. where they illustrate the principles of hydrophobic hydration.
When b is a liquid or solid solution dilute in A and a is a
dilute gas, S is the Ostwald absorption coecient, a common
measure of gas solubility, as illustrated for methane in Fig. 1.
Since b is then a hardly compressible condensed phase its prop-
III. Hydrophobic hydration
erties as a solvent are insensitive to the pressure to moderate In Table 1 we give the results for the transfer of a molecule of
pressures; and the equilibrium ratio (cbA /caA )eq of the concentra- methane, ethane, propane, or n-butane from the gas phase into
tions of A in these dilute phases is independent of their sepa- water, at 300 K, as derived from (18)(22), with S taken from
rate values (Henrys law); so S is practically a function of the compilations of Battino.1,8890 The b phase is pure or prac-
the temperature alone (as remarked in Section I). This circum- tically pure water. For eb and its temperature derivative,
stance, in which both phases are dilute in A, is also, we saw, required for the correction terms in (19), (20) and (22), we took
that in which DFV , DGp , DSV , and DSp , and so also DEV , the data from the compilation by Rowlinson and Swinton.91
DHp , (DCV)V , and (DCp)p , are independent of the concentra- Those data are for water as it coexists with its vapor, so not
tions. Thus, their values are independent of whether or not at 1 atm, whereas the data for S are at 1 atm, but the distinc-
the transfer of the A molecule from a to b occurs with the ratio tion is of no importance since neither the coecient of thermal
cbA /caA equal to that for which the two phases are in equili- expansion of water nor the Ostwald absorption coecient S
brium. In this case, then, the signicance of (16) and (17) is varies signicantly with the pressure over this range. What
that DFV , DGp , and the quantities derivable from them are makes the pressure variation of eb even smaller is that it is pro-
the same as when the transfer occurs with the two phases in portional to the temperature variation of the isothermal com-
equilibrium, even when it does not; and these quantities are pressibility,92 and 300 K is not too far from 320 K, where the
then obtainable from the equilibrium solubility S, which is vir- compressibility of water has a temperature minimum.91 The a
tually a function of the temperature alone. phase, meanwhile, is a dilute, practically ideal gas, for which
When a and b are both condensed phases, but both again ea 1/T and dea/dT  1/T2. The correction terms in (19),
dilute in A, the same considerations apply. The equilibrium (20), and (22) are small; the dierence between DSV /k and
ratio S in (17) is then the partition coecient of A between DSp /k and between DEV/kT and DHp/kT at 300 K is only
the phases at equilibrium, and is virtually independent of the 0.9. By coincidence, the two correction terms on the right-hand
pressure and of the separate values of cbA and caA ; it is practi- side of (22), which are opposite in sign, nearly cancel each
cally a function of the temperature alone; and DFV and DGp , other, with the result that (DCV)V and (DCp)p are nearly equal
as given by (16) and (17), and the quantities obtainable from at 300 K.
them, are the same as when the transfer occurs with the two Table 1 has many important lessons. The rst and most
phases in equilibrium even when it does not. obvious is that the free energy of transfer is positive and is
Another case of practical interest is that in which the a several multiples of the thermal energy kT, which, via (18), just
phase is pure or nearly pure liquid or solid A while b is again reects the low solubility of the hydrocarbons in water. It is
a condensed phase dilute in A. We continue to dene DFV the next entries in the table that tell the hydrophobicity story:
and DGp by (3) and (4) and DSV and DSp by (5) and (6), that the low solubility is a consequence of an unfavorable
so the remaining relations (7)(17), which follow from (3)
(6), still all hold. Now caA in the denominator of the argument
of the logarithms in (3)(6) is practically a constant, and, Table 1 Thermodynamics of transfer of a hydrocarbon molecule
since we are supposing b to be dilute in A, both DFV and from the gas phase into water at 300 K
DGp are independent of the remaining variable cbA . Thus, DFV DGp DSV DSp DEV DHp DCV V DCp p
(16) with (17) hold in the case, too.
kT kT k k kT kT k k
Summarizing, there are three common situations in which
all the relations (7)(16) hold, with S, dened by (17), practi- CH4 3.4 7.5 8.4 4.1 5.0 29 29
C2H6 3.1 9.7 10.6 6.6 7.5 36 36
cally a function of the temperature alone; i.e., practically inde-
C3H8 3.4 11.0 11.9 7.6 8.5 41 41
pendent of pressure to moderate pressures and independent of n-C4H10 3.6 12.5 13.4 8.9 9.8 40 40
the concentration of A in the dilute phase(s). These situations

Phys. Chem. Chem. Phys., 2003, 5, 30853093 3087


View Article Online

(negative) entropy change that overweighs a favorable (nega- the solubility passes through a minimum with increasing tem-
tive) energy change. The dierence between DSV /k and perature and thereafter increases is a characteristic also of the
DEV/kT (or, equivalently, between DSp /k and DHp/ solubility of most hydrocarbons in water1,8890(as seen for
kT ), which is DFV /kT ( DGp /kT ), is nevertheless smaller methane in Fig. 1).
than either one separately; i.e., DEV/kT, while less than For the transfer of a molecule between coexisting phases
DSV /k, is nevertheless more than half of it (DHp/kT is at temperature T, the enthalpy and entropy changes DH
more than half of DSp /k), so there is considerable cancella- and DS (6 DS*) are related by DH TDS (6 TDS*). Refer-
tion between them. That the entropy of transfer is negative ring still to Fig. 3, we will for now use the notation DHA
at 300 K means, from (7), that at this temperature the free ( TDSA) for the enthalpy of transfer of an A molecule
energy of transfer becomes more positive (more unfavorable) from the phase a relatively rich in A to the phase b poorer
with increasing temperature; and that the energy of transfer in A, and DHB ( TDSB) for the enthalpy of transfer of a
is negative at 300 K means, from (20), that at this temperature molecule B from b to a. One may then show that DHA and
the solubility decreases with increasing temperature. Thus, at DSA are negative for temperatures below some temperature
300 K, these hydrocarbonsand hydrophobes generally T2 between T1 and T4 , they vanish at T2 , and are positive
become more hydrophobic with increasing temperature. above T2 ; while DHB and DSB are negative below some tem-
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

The heat capacity changes on transfer, (DCV)V and (DCp)p , perature T3 between T1 and T4 , they vanish at T3 , and are
are positive and huge: many multiples of the Boltzmann positive above T3 ; with T3 < T2 if the coexistence curve is
constant k. This was one of the earliest observations of a oriented as in the gure.
hydrophobic eect .13,93 This and the negative energetic For the transfer of an A molecule from the phase a richer in
and entropic changes are ascribable to structural changes in A to the phase b poorer in A, the modied entropy change
the solvent, generally thought to be the strengthening of (or DSA , dened analogously to DSp in (6), is algebraically smaller
formation of additional) hydrogen bonds among the water (more negative or less positive) than DSA itself, and so is still
molecules neighboring the solute.2 negative when DSA vanishes at the temperature T2 .
A further observation from the table is the manifestation of Where the b phase is very dilute in A it hardly matters to the
enthalpyentropy compensation ,15,94,95 according to which values of the DH and DS* of transfer whether the transfer of
free-energy changes (free energies of transfer in this instance) the A molecule from the a phase is to the coexisting b phase
are nearly constant through a homologous series, although (where they are then DHA and DSA ) or to pure B (pure water,
the changes in the separate energetic and entropic components say, if b is an aqueous phase). In the latter case the unmodied
of the free energy vary substantially through the series. DS of transfer would be innite. Likewise, if the a phase is
Often when a solute can itself form hydrogen bonds with nearly pure A it hardly matters to the values of the DH and
water the result is a closed-loop solubility curve,96 illustrated DS of transfer whether the transfer of the A molecule to b is
here in Fig. 3. The gure shows the temperature dependence from the coexisting a phase (when they are then DHA and
of the mutual solubility of two hypothetical substances A DSA) or from pure A, and now in either case there is hardly
and B at a xed pressure (or while simultaneously in equili- any dierence between DS and DS*.
brium with vapor). The tilt of the solubility curve is exagger- This completes our review of the thermodynamics of hydro-
ated in the gure to make clear that in general the phobic hydration. We turn next to microscopic models and to
temperature T1 of minimum solubility of A in B is not the the phenomenon of hydrophobic attraction.
same as the temperature T4 of minimum solubility of B in A
(although in theoretical models with an articial symmetry it
may be97). One may think of the A-rich side of the loop as IV. Microscopic models
the phase earlier called a and the B-rich side as the aqueous
phase b. Since on the lower part of the loop the solubility of There are many and diverse microscopic models of solutions of
A in B (as well as that of B in A) is decreasing with increasing hydrophobes in water (see refs. 9, 19, 21, 29, 35, 40, 77, 78,
temperature, it is manifesting the hydrophobic eect.23 That 98 among many others), but we shall here describe only those
we have ourselves been studying, which is a class of lattice
models.87,99102
Fig. 4 is an example in two dimensions with lattice coordina-
tion number Z 4, but the same basic model may be in any
number of dimensions with a variety of lattice types. Solvent
molecules are represented by the larger circles, one at each lat-
tice site. Each may be in any of q ( 2) possible orientations
or internal states, of which one (state number 1, say) is distin-
guished from the q  1 others: neighboring solvent molecules
interact with an energy w when both are in that special state
1 but with a higher energy u ( > w) otherwise. Solute molecules
are represented by the smaller circles in Fig. 4. They can be
accommodated only on the bonds between lattice sites, i.e.,
between neighboring pairs of solvent molecules, and only then
if both solvent molecules of the neighboring pair are in the
special state or orientation 1. Thus, accommodation of a solute
restricts the possible orientations of the neighboring solvent
molecules and at the same time puts that solvent pair in its
low-energy conguration. Thus, the model has built into it
the basic mechanism of hydrophobicity: the forced accommo-
dation of the solute is energetically favorable but entropically
Fig. 3 Closed-loop solubility curve with components A and B. Con- unfavorable.
centration variable xB is the mole or mass fraction of B. Points C and
C0 mark the lower and upper critical solution points. Temperatures T1
Lest the models seem more articial than they are, it should
and T4 are those of the minimum solubility of A in B and of B in A, be emphasized that, as in any lattice uid model,97,103 the
respectively, while the intermediate temperature T2 is that at which molecules are not to be thought of as conned to lattice sites
DHA and DSA vanish, and T3 that at which DHB and DSB vanish. or bond centers. One should think of the volume divided into

3088 Phys. Chem. Chem. Phys., 2003, 5, 30853093


View Article Online

cubic lattices, respectively, in three dimensions, etc.). In the


equivalent one-component lattice gas106 in which the cells are
of volume v0 , the activity is z, and the near-neighbor inter-
action energy is e, the equivalences are
e u  w; v0 z 1=q  1: 26
The probability P1 that any given solvent molecule in the
hydrophobic-interaction model be in the special state 1 rather
than in any of the q  1 other states is equivalently the prob-
ability that any given cell in the lattice gas model be occupied
rather than empty or that any given spin in the Ising model be
in the direction of the eld rather than opposed.
The model in any number of dimensions and for any
coordination number may be treated in BetheGuggenheim
approximation107,108 as well as by simulation. (For the applica-
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

tion of the BetheGuggenheim approximation to dierent lat-


tice models of water and hydrophobic solvation see Besseling
Fig. 4 The lattice model. Large open circles at the lattice sites repre- and Lyklema35 and Eads.78) The approximation is exact on a
sent the solvent molecules, small lled circles on the bonds between Bethe lattice (Cayley tree), on which there are no closed loops,
sites represent the solute molecules. so that, as in one dimension, between any two sites there
is only a single, unique path. Then W(r1 ,r2) may be written
W(r), where r is the distance between bond centers measured
cells (only as a coordinate system, not with impenetrable as a number of lattice steps; and then via (24),
walls), so that every molecule has access to any point in the
whole volume. The only articiality is that the interactions W r kT ln gr  1 r  2 27
between molecules now depend on discrete rather than contin-
uous distance variables. where g(r) is the pair distribution function (radial distribution
The two parameters u  w and q  1 are properties of the function) of the equivalent one-component lattice gas in one
model solvent alone. The model has a third parameter, v, dimension. For r 1,
which is the energy of interaction between an accommodated W 1 kT ln y; 28
solute molecule and the two solvent molecules that neighbor
it. All three parameters enter into the solubility S, dened in with y (0  y  1) the dimensionless density of the one-compo-
(17), while the solvent-mediated part of the potential of mean nent lattice gas. As remarked earlier, this lattice-gas density is
force between pairs of solutes is determined by the solvent also the probability P1 that any solvent molecule in our origi-
parameters u  w and q  1 alone. nal model be in the special state 1. Further, the solubility S in
These follow from the expressions87,99,100 our model, via (23), is given in Bethe-Guggenheim approxima-
tion by
S P11 ev=kT 23
Pr1 ; r2 S y2 g1ev=kT : 29
W r1 ; r2 kT ln ; 24
P211 In this approximation, the lattice-gas g(r) and density y
for the rst of which it is assumed that the phase that is in required in (27)(29) are expressible in terms of the solvent
equilibrium with the saturated solution is a dilute gas. Here parameters u  w and q  1 of the hydrophobic-interaction
P11 is the probability that any neighboring pair of solvent model by
molecules in the pure solvent will both be found to be in   
1 Q1 r
the special orientation 1; P(r1 ,r2) is the probability that gr 1 1 r  1 30
the solvent molecules that are the neighbors of the bonds cen- y Q1
tered at r1 and r2 will all be found to be in that special orien- 1 1a
tation; and W(r1 , r2) is the solvent-mediated part of the 1 31
y a1 ca
potential of mean force between a pair of solute molecules
at r1 and r2 in the innitely dilute solution; i.e., the potential where
of mean force from which the direct solutesolute interaction, p
whatever that might be, has been subtracted. The parameters Q 1 4c  11  yy 32
u  w and q  1 are implicit in P11 and P(r1 , r2) while v c euw=kT 33
appears explicitly in (23). The nearest-neighbor pair correla-
 
tion P11 and the pairpair correlation function P(r1 ,r2) are 1 ca Z1
obtained analytically in the one-dimensional version of the q  1a; 34
1a
model99 and with sucient accuracy by Monte Carlo simula-
tion in two or three dimensions.87,100 with Z again the coordination number of the lattice. The lat-
For the simulations it is convenient to recognize the equiva- tice-gas density y is thus given parametrically (parameter a)
lence of the model to a related Ising model or one-component in terms of the model parameters q  1 and u  w (and Z
lattice gas.102 With conventional Ising-model notation ,104,105 and the temperature T ) by (31) and (34), with c dened by (33).
the spinspin interaction-energy parameter J and magnetic With Z 2 the BetheGuggenheim approximation becomes
eld H in the equivalent Ising model are related to u  w and the exact one-dimensional version of the model.99 In the limit
q  1 by Z ! 1 and u  w ! 0 with xed f Z(u  w) it becomes the
mean-eld approximation, in which g(r)1 for all r  1 and
1 1
J u  w; 2H Zu  w  kT lnq  1 25 for which y is given implicitly in terms of q  1 and f by q  1
4 2 (1/y  1)exp(yf/kT ).
where Z is the coordination number of the lattice (Z 2 in one One object in the study of these models is to nd a connec-
dimension, Z 6 or 8 for the simple cubic or body-centered tion between the free energy of hydrophobic hydration (DFV

Phys. Chem. Chem. Phys., 2003, 5, 30853093 3089


View Article Online

or DGp , Section III), and the solvent-mediated part, W(r1 , r2),


of the potential of mean force between hydrophobic solutes.
To this end we choose the parameters u  w and q  1 to match
the experimental solubility S of methane as closely as possible
over the 60 temperature interval 273 K  T  333 K, and then
with these values of the parameters calculate W.
We continue to dene f Z(u  w). In Fig. 5 we show the
nearly perfect t to the experimental S obtained with the
model in BetheGuggenheim approximation with Z 8 and
the model parameters
f=k 1008 K; q  1 6:16; v=k 464:3 K: 35
The corresponding plot when the properties of the model on
the body-centered cubic lattice (Z 8) are obtained by Monte
Carlo simulation87 and the solubility then t with the para-
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

meter values
Fig. 6 W(r)/kT for the model in BetheGuggenheim approximation
f=k 1026 K; q  1 6:7; v=k 406:7 K 36 with Z 8.
is almost indistinguishable from that in Fig. 5.
It is noteworthy that the parameter values in (35) and (36), energetically favorable change in the solventsolvent inter-
where Z 8 in both, are not very dierent. It should also be action that accompanies the dissolution of the solute, which
noted that it is not a forgone conclusion that the experimental overweighs what in this model (not necessarily in reality) is
S even over this restricted temperature range could have been the unfavorable solutesolvent interaction.
t by these three-parameter functions. Only the two para- With the parameters in (35) or (36) that t the experimental
meters u  w and q  1 occur in the factor P11 in (23), so the S, one may now calculate the solvent-mediated part of the
slope and curvature and thus the whole shape of a plot of potential of mean force, W, between methane molecules in
kT ln S as a function of T have to be t by those alone, while the innitely dilute solution. This follows from (24) in general
the third parameter, v, corresponds merely to a constant dis- for the models of this class, or in particular from (27) and (28)
placement of such a plot without a change in its shape. In with (30)(34) in BetheGuggenheim approximation. In Fig. 6
BetheGuggenheim approximation the t is better the greater we show W(r)/kT in this approximation with Z 8, as a func-
the coordination number Z, and so, although already nearly tion of r at each of the three temperatures 273 K, 298 K, and
perfect at Z 8, is even better in mean-eld approximation 333 K. These curves are strongly reminiscent of the theoretical
(Z ! 1), where the t is with the parameter values curves in Marcelja et al.,17 although those are indexed by
f=k 912:4 K; q  1 4:95; v=k 498:2 K; 37 solute size rather than temperature. In Fig. 7, from ref. 87,
we show the corresponding results from the Monte Carlo
while the t is poorest, although still acceptable, in one dimen- simulations of the model on the body-centered cubic lattice.
sion (Z 2).87 In the latter plot W(r1 ,r2) is again called W(r), with r again
We note in (35)(37) that the model parameter v is positive. in multiples of the shortest distance between bond centers, p
One must make a distinction between the interaction energy but this shortest distance is now the fraction 1/ 3
between solute and solvent, which here is v, and the energy 0.577   of the bond length itself. In both gures, then, r 1
or enthalpy change DEV or DHp of the whole system that is the closest the centers of two hydrophobic molecules may
accompanies the forced accommodation of the solute. The lat- come if both molecules are imagined located at the lattices
ter are negative here, as well as in reality, because they follow bond centers.
from the experimental S, which we have t nearly exactly with Not visible in Fig. 6 on the scale of that plot, and scarcely
the parameter values in (35)(37). But we note also in (35)(37) visible in Fig. 7, is the crossing of the curves for the various
that f, which is Z(u  w), is greater than v; so the favorable temperatures: the higher the temperature the stronger the
energy and enthalpy of hydration is a consequence of the hydrophobic attraction as measured by W(1) but the shorter
is its range as measured by its rate of decay as r ! 1. For the
body-centered cubic lattice, and generally in three dimensions

Fig. 5 Fit of the model S in BetheGuggenheim approximation with


Z 8 to the experimental S for methane. The points are the theoreti-
cal values with the parameters given in (35); the curve is from Battinos Fig. 7 W(r)/kT by Monte Carlo simulation of the model on the
interpolation formula tting the experimental results.1 body-centered cubic lattice.

3090 Phys. Chem. Chem. Phys., 2003, 5, 30853093


View Article Online

with the direct interactions of short range, W(r) decays propor- Let the resulting unfavorable free-energy density in those
tionally to (1/r)exp(r/x), and in BetheGuggenheim approx- aected regions be some f > 0. Then DGp v1f > 0; while
imation proportionally to exp(r/x), where the decay lengths the solvent-mediated potential energy, W(1) in the notation
x decrease with increasing temperature. This may not be an of our model, is W(1) (v2  2v1)f, which is negative (attrac-
important feature of hydrophobic forces, however, because tive) because v2 < 2v1 . But also v2 > v1 , so v2  2v1 >  v1 ;
W(r) is already so small when the curves cross: the three curves i.e., W(1) < DGp . In this argument we have assumed that
in Fig. 6 taken in pairs cross at r 2.7, 4.0, and 5.9. the free energy density (what we have called f) in the lens-
This attraction, which is largely entropic in origin, has been shaped overlap region and in the non-overlap region are not
termed the hydrophobic bond .3,6 In both gures it is seen very dierent.
that the strength of the attraction, W(1), is only about kT. A related argument allows one to guess the sign of the three-
The direct van der Waals attraction between pairs of solutes body contribution to the hydrophobic attraction. This is the
at this distance is only about 15 or 20% of the solvent-mediated subject of the next section.
attraction,19,109 so the total is still only about kT. This means,
for example, that of the hydrophobic bonds in the native
structure of a protein, at any moment approximately the frac- V. Three-body potential
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

tion 1/(1 + e) 1/4 of them are broken . This may be


important for the dynamical functioning of the protein.110 In an analysis of the one-dimensional version of this lattice
This has illustrated the calculation of the strength W(1) of model101 it was found that the solvent-mediated part of the
the hydrophobic attraction in the model, using parameters that potential of mean force among n solute molecules on the bonds
closely t the experimental S; or, equivalently, from (16), that centered at x1 , x2 ,. . ., xn with x1 < x2 <   < xn is the sum of
t the experimental hydration free energy DFV or DGp . Fig. 8 the pair potentials between nearest neighbors only:
now shows how W(1)/kT varies with DGp /kT. Shown are the
W r12 ; r23 ;    ; rn1;n W r12 W r23    W rn1;n
BetheGuggenheim approximation with Z 2, 3, 8, and 1
(Z 2 is the one-dimensional model and Z 1 the mean- 38
eld approximation), and the results of the simulation on the where r12 x2  x1 , etc. With three solute molecules at x1 , x2 ,
body-centered cubic lattice. We see a steady progression with x3 , this is
Z. Over this restricted temperature range all the plots are
nearly linear: almost exactly linear from the simulations but W r12 ; r23 W r12 W r23 : 39
with slight positive curvature for Z 2 and 3 and slight nega-
This is greater than the sum of the three pair potentials,
tive curvature for Z 8 and 1 in the BetheGuggenheim
approximation. They are nearly parallel with a common slope W r12 ; r23  W r12  W r23  W r13 W r13 > 0
near 0.7. 40
These plots answer the question, at least for this class of
models, of how the strength of the solvent-mediated attraction because the pair potential W(r) is negative (attractive); i.e.,
between a pair of hydrophobic solute molecules depends on the three-body contribution to the energy of interaction of
the hydration free energy of a single one; i.e., they are a quan- three such solute molecules is positive: the three-body force
titative answer to the question implicit in Fig. 2. We note again is repulsive.
that the hydrophobic eect, i.e., DGp > 0 and W(1) < 0, One may guess that the same will be true in near linear con-
becomes stronger with increasing temperature. We also see gurations of three hydrophobic solutes in two or three dimen-
again that W(1) is only about kT, while DGp , from the gure sions as well, while an extension of the argument at the end of
or from Table 1, is about 3kT. One can understand why Section IV indicates that it may also be true in triangular con-
W(1) < DGp .87 Suppose in Fig. 2 that the volume of solvent gurations of the three. An equilateral conguration of three is
structurally and energetically aected by the presence of an shown schematically in Fig. 9. The volume of solvent aected
isolated solute molecule is v1 and that the total volume so by the presence of a single solute molecule is called u1 ; it is the
aected by a closely spaced pair of solute molecules is v2 . same as what was called v1 in the argument at the end of Sec-
tion IV. The volume of the lens-shaped region of overlap of
two such regions is u2 , previously called 2v1  v2 . The volume
of the curvilinear region of triple overlap is u3 .
The total volume of solvent aected by the presence of the
solutes in this conguration is 3u1  3u2 + u3 . But with W1 ,
W2 , and W3 the one-, two-, and three-body eective potentials,
the total excess free energy due to the presence of the three
solute molecules is 3W1 + 3W2 + W3 , because there are three
distinct single bodies, three distinct pairs, and one triple. The

Fig. 8 Variation of W(1)/kT with DGp /kT; W(1) is the strength of


the hydrophobic attraction between a pair of solute molecules while Fig. 9 Three solute molecules in an equilateral conguration. The
DGp is the hydration free energy of a single one. The solid curves are volume of solvent aected by the presence of a single solute molecule
the BetheGuggenheim approximation with the indicated coordination is u1 , the volume of the lens-shaped region of overlap of two such
numbers Z; the dashed curve is the result of Monte Carlo simulation of regions is u2 , and the volume of the curvilinear region of triple overlap
the model on the body-centered cubic lattice, for which Z 8. is u3 .

Phys. Chem. Chem. Phys., 2003, 5, 30853093 3091


View Article Online

one-body potential W1 is the earlier hydration free energy DGp , References


which was v1f, now called u1f; the pair potential W2 is what
was earlier called W(1), the eective potential between a closely 1 R. Battino, in IUPAC Solubility Data Series, ed. H. L. Clever
and C. L. Young, Pergamon, Oxford, 1987, vol. 27/28, pp. 16.
spaced pair of solutes, which was (v2  2v1)f, now called u2 f.
2 H. S. Frank and M. W. Evans, J. Chem. Phys., 1945, 13,
Thus, the three-body potential W3 , which is our present 507532.
object, is here seen to be W3 u3f, which is positive. Thus, 3 W. Kauzmann, Adv. Protein Chem., 1959, 14, 163.
in this conguration, as in the linear conguration in the 4 G. Nemethy and H. A. Scheraga, J. Phys. Chem., 1962, 66, 1773
one-dimensional model, the three-body force is repulsive, 1789.
which is what we wished to determine. 5 G. Nemethy, Angew. Chem., Int. Ed., 1967, 6, 195206.
6 G. Nemethy, H. A. Scheraga and W. Kauzmann, J. Phys. Chem.,
But for application to real hydrophobic interactions this
1968, 72, 1842.
argument can be no more than suggestive; it is not a substitute 7 F. H. Stillinger, J. Solution Chem., 1973, 2, 141158.
for detailed calculation with realistic models.53,76 8 J. C. Owicki and H. A. Scheraga, J. Am. Chem. Soc., 1977, 99,
74137418.
9 L. R. Pratt and D. Chandler, J. Chem. Phys., 1977, 67, 3683
3704.
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

VI. Summary and conclusions 10 C. Tanford, The Hydrophobic Eect: Formation of Micelles and
Biological Membranes, Wiley, New York, 2nd edn., 1980.
The basic result of the thermodynamics as outlined in Section 11 A. Ben-Naim, Hydrophobic Interactions, Plenum, Oxford,
II and illustrated in Section III is contained in eqns. (18)(22), 1980.
12 R. L. Baldwin, Proc. Natl. Acad. Sci., 1986, 83, 80698072.
where the thermodynamics of transfer of a hydrophobic mole- 13 D. H. Everett, in Hydrogen-Bonded Solvent Systems, ed.
cule into water are seen to follow from measurements of the A. K. Covington and P. Jones, Taylor & Francis, London,
equilibrium partition coecient or Ostwald absorption coe- 1968, pp. 18.
cient. That the suitably dened free energy of transfer [denoted 14 F. Franks, Water, A Matrix of Life, Royal Society of Chemistry,
with an asterisk in eqns. (3) and (4)] is positive and equal to a Cambridge, 2nd edn., 2000.
few multiples of the thermal energy kT, and that the energy (or 15 R. Lumry and S. Rajender, Biopolymers, 1970, 9, 11251227.
16 P. J. Rossky and M. Karplus, J. Am. Chem. Soc., 1979, 101,
enthalpy) and suitably dened entropy of transfer are both 19131937.
negative, are characteristic of, and essentially dene, hydro- 17 S. Marcelja, D. J. Mitchell, B. W. Ninham and M. J. Sculley,
phobic solutes. Hydrophobicity increases with increasing tem- J. Chem. Soc., Faraday Trans. 2, 1977, 73, 630.
perature, reecting the corresponding decrease in solubility. 18 K. A. Dill, Biochemistry, 1990, 29, 71337155.
Many of the features of the hydrophobic eect are realisti- 19 D. E. Smith and A. D. J. Haymet, J. Chem. Phys., 1993, 98,
cally rendered by the lattice model described in Section IV. 64456454.
20 S. Garde, G. Hummer and M. E. Paulaitis, Faraday Discuss.,
The model was designed to make the accommodation of a 1996, 103, 125139.
solute in the solvent energetically favorable but entropically 21 B. Guillot and Y. Guissani, J. Chem. Phys., 1993, 99, 80758094.
unfavorable, and thus to incorporate the basic mechanism of 22 J. Forsman and B. Jonsson, J. Chem. Phys., 1994, 101, 5116
hydrophobicity as revealed by the thermodynamics. Once the 5125.
parameters in the model are chosen to reproduce the experi- 23 M. E. Paulaitis, S. Garde and H. S. Ashbaugh, Curr. Opin.
mentally measured solubility of methane and its temperature Colloid Interface Sci., 1996, 1, 376383.
24 M. Pellegrini, N. Grnbech-Jensen and S. Doniach, J. Chem.
dependence in the temperature interval 273 to 333 K, the Phys., 1996, 104, 86398648.
model leads to a realistic picture of the solvent-mediated part, 25 G. Hummer, S. Garde, A. E. Garca, A. Pohorille and L. R.
W(r), of the potential of mean force between a pair of solute Pratt, Proc. Natl. Acad. Sci., 1996, 93, 89518955.
molecules as a function of their separation r. It is found that 26 G. Hummer, S. Garde, A. E. Garca, M. E. Paulaitis and L. R.
W(1) , the potential at close approach of the pair, is nega- Pratt, J. Phys. Chem. B, 1998, 102, 10 46910 482.
tive, which is the hydrophobic attraction, and becomes more 27 S. Garde, A. E. Garca, L. R. Pratt and G. Hummer, Biophys.
Chem., 1999, 78, 2132.
negative with increasing temperature, in accord with the con- 28 M. A. Gomez, L. R. Pratt, G. Hummer and S. Garde, J. Phys.
clusion from the thermodynamic analysis that hydrophobicity Chem. B, 1999, 103, 35203523.
increases with increasing temperature. The strength W(1) of 29 L. R. Pratt, Annu. Rev. Phys. Chem., 2002, 53, 409436.
the hydrophobic attraction increases nearly linearly with the 30 M. E. Paulaitis and L. R. Pratt, Adv. Protein Chem., 2002,
free energy of hydrophobic hydration, but is less than the 62, 283.
latter: only about kT as compared with about 3kT. That 31 S. Ludemann, H. Schreiber, R. Abseher and O. Steinhauser,
J. Chem. Phys., 1996, 104, 286295.
the solvent-mediated attraction is as small as kT means that 32 R. M. Lynden-Bell and J. C. Rasaiah, J. Chem. Phys., 1997, 107,
hydrophobic bonds are easily broken , which may have 19811991.
implications for dynamical eects in protein structure. 33 D. L. Bergman and R. M. Lynden-Bell, Mol. Phys., 2001, 99,
It is remarked in Section V that in the one-dimensional ver- 10111021.
sion of the lattice model the three-body force in the solvent- 34 F. M. Floris, M. Selmi, A. Tani and J. Tomasi, J. Chem. Phys.,
mediated interaction among hydrophobic solutes is always 1997, 107, 63536365.
35 N. A. M. Besseling and J. Lyklema, J. Phys. Chem. B, 1997, 101,
repulsive. A simple argument is presented according to which 76047611.
that may still be so in various congurations in two and three 36 G. Graziano, J. Chem. Soc., Faraday Trans., 1998, 94, 3345
dimensions as well; but one is warned that the argument, while 3352.
suggestive, is not denitive. 37 R. D. Mountain and D. Thirumalai, Proc. Natl. Acad. Sci. USA,
1998, 95, 84368440.
38 K. A. T. Silverstein, A. D. J. Haymet and K. A. Dill, J. Am.
Chem. Soc., 1998, 120, 31663175.
39 K. A. T. Silverstein, K. A. Dill and A. D. J. Haymet, Fluid Phase
Acknowledgements Equilib., 1998, 151, 8390.
40 K. A. T. Silverstein, A. D. J. Haymet and K. A. Dill, J. Chem.
We thank Jonathan Widom and Harold Scheraga for helpful Phys., 1999, 111, 80008009.
discussions. KK acknowledges the award of a fellowship from 41 N. T. Southall and K. A. Dill, J. Phys. Chem. B, 2000, 104, 1326
1331.
the Japan Society for the Promotion of Science, which made 42 N. T. Southall, K. A. Dill and A. D. J. Haymet, J. Phys. Chem.
possible his participation in this work. The work was sup- B, 2002, 106, 521533.
ported by the U.S. National Science Foundation and the 43 K. Lum, D. Chandler and J. D. Weeks, J. Phys. Chem. B, 1999,
Cornell Center for Materials Research. 103, 45704577.

3092 Phys. Chem. Chem. Phys., 2003, 5, 30853093


View Article Online

44 E. Fois, A. Gamba and C. Redaelli, J. Chem. Phys., 1999, 110, 78 C. D. Eads, J. Phys. Chem. B, 2002, 106, 12 28212 290.
10251035. 79 M. Predota, I. Nezbeda and P. T. Cummings, Mol. Phys., 2002,
45 M. Predota and I. Nezbeda, Mol. Phys., 1999, 96, 12371248. 100, 21892200.
46 C. H. Cho, S. Singh and G. W. Robinson, Phys. Rev. Lett., 1996, 80 M. T. Stone, P. J. in t Veld, Y. Lu and I. C. Sanchez, Mol. Phys.,
76, 16511654. 2002, 100, 27732792.
47 P. M. Wiggins, Physica A, 1997, 238, 113128. 81 J. Hernandez-Cobos and L. F. Vega, J. Mol. Liq., 2002, 101,
48 S. W. Rick, J. Phys. Chem. B, 2000, 104, 68846888. 113125.
49 B. Madan and K. Sharp, J. Phys. Chem., 1996, 100, 77137721. 82 G. L. Pollack, Science, 1991, 251, 13231330.
50 L. E. S. de Souza and D. Ben-Amotz, J. Chem. Phys., 1994, 101, 83 W. Blokzijl and J. B. F. N. Engberts, Angew. Chem., Int. Ed.
98589863. Engl., 1993, 32, 1545.
51 P. Bruscolini and L. Casetti, Phys. Rev. E, 2001, 64, 051805. 84 T. P. Silverstein, J. Chem. Educ., 1998, 75, 116118.
52 S. Shimizu and H. S. Chan, J. Chem. Phys., 2000, 113, 46834700 85 N. Matubayasi and M. Nakahara, J. Phys. Chem. B, 2000, 104,
[Erratum 2002, 116, 8636]. 10 35210 358.
53 S. Shimizu and H. S. Chan, J. Chem. Phys., 2001, 115, 14141421. 86 H. A. Scheraga, J. Biomol. Struct. Dynam., 1998, 16, 447460.
54 S. Shimizu and H. S. Chan, Proteins: Struct., Funct. Genet., 2002, 87 K. Koga, P. Bhimalapuram and B. Widom, Mol. Phys., 2002,
48, 1530. 100, 37953801.
55 P. R. ten Wolde, S. X. Sun and D. Chandler, Phys. Rev. E, 2001, 88 R. Battino, in IUPAC Solubility Data Series, ed. W. Hayduk,
Published on 24 June 2003. Downloaded by University of Groningen on 05/07/2017 15:10:07.

65, 011201-1-9. Pergamon, Oxford, 1982, vol. 9, pp. 12.


56 P. R. ten Wolde and D. Chandler, Proc. Natl. Acad. Sci. USA, 89 R. Battino, in IUPAC Solubility Data Series, ed. W. Hayduk,
2002, 99, 6539. Pergamon, Oxford, 1986, vol. 24, pp. 12.
57 R. P. Bonifacio, A. A. H. Padua and M. F. Costa Gomes, 90 R. Battino, in IUPAC Solubility Data Series, ed. W. Hayduk,
J. Phys. Chem. B, 2001, 105, 84038409. Pergamon, Oxford, 1986, vol. 24, pp. 1617.
58 N. Guisoni and V. Bohomoletz Henriques, Braz. J. Phys., 2000, 91 J. S. Rowlinson and F. L. Swinton, Liquids and Liquid Mixtures,
30, 736740. Butterworths, London, 3rd edn., 1982, p. 46.
59 T. M. Raschke, J. Tsai and M. Levitt, Proc. Natl. Acad. Sci. 92 J. S. Rowlinson and F. L. Swinton, Liquids and Liquid Mixtures,
USA, 2001, 98, 59655969. Butterworths, London, 3rd edn., 1982, p. 14.
60 E. Gallicchio, M. M. Kubo and R. M. Levy, J. Phys. Chem. B, 93 J. T. Edsall, J. Am. Chem. Soc., 1935, 57, 15061507.
2000, 104, 62716285. 94 E. Grunwald and C. Steel, J. Am. Chem. Soc., 1995, 117, 5687
61 A. Wallqvist, E. Gallicchio and R. M. Levy, J. Phys. Chem. B, 5692.
2001, 105, 67456753. 95 H. Qian and J. J. Hopeld, J. Chem. Phys., 1996, 105, 9292
62 J. Hernandez-Cobos, A. D. Mackie and L. F. Vega, J. Chem. 9298.
Phys., 2001, 114, 75277535. 96 J. Hirschfelder, D. Stevenson and H. Eyring, J. Chem. Phys.,
63 D. Renzi, C. M. Carlevaro, C. Stoico and F. Vericat, Mol. Phys., 1937, 5, 896.
2001, 99, 913922. 97 J. C. Wheeler, Annu. Rev. Phys. Chem., 1977, 28, 411443.
64 A. Bakk and J. S. Hye, Physica A, 2002, 303, 286294. 98 H. A. Sallouta and G. M. Bell, Mol. Phys., 1976, 32, 839855.
65 A. Bakk, J. S. Hye and A. Hansen, Physica A, 2002, 304, 99 A. B. Kolomeisky and B. Widom, Faraday Discuss., 1999, 112,
355361. 8189.
66 D. M. Huang and D. Chandler, Proc. Natl. Acad. Sci. USA, 100 G. T. Barkema and B. Widom, J. Chem. Phys., 2000, 113, 2349
2000, 97, 83248327. 2353.
67 D. M. Huang and D. Chandler, J. Phys. Chem. B, 2002, 106, 101 D. Bedeaux, G. J. M. Koper, I. Ispolatov and B. Widom,
20472053. Physica A, 2001, 291, 3948.
68 H. S. Ashbaugh, T. M. Truskett and P. G. Debenedetti, J. Chem. 102 G. M. Schutz, I. Ispolatov, G. T. Barkema and B. Widom,
Phys., 2002, 116, 29072921. Physica A, 2001, 291, 2438.
69 T. Urbic, V. Vlachy, Y. V. Kalyuzhnyi, N. T. Southall and K. A. 103 B. Widom, in Proc. R. A. Welch Found. Conf. on Chem. Research
Dill, J. Chem. Phys., 2002, 116, 723729. XVI. Theoretical Chemistry, 1972, pp. 161202.
70 T. Ghosh, A. E. Garca and S. Garde, J. Chem. Phys., 2002, 116, 104 C. Domb, Adv. Phys., 1960, 9, 149361.
24802486. 105 R. Kubo, Statistical Mechanics, North-Holland, Amsterdam,
71 P. F. B. Goncalves and H. Stassen, J. Comput. Chem., 2002, 23, 1965, p. 303.
706714. 106 T. L. Hill, Statistical Mechanics, McGraw-Hill, New York, 1956,
72 G. Graziano, Can. J. Chem., 2002, 80, 401412. pp. 291293.
73 J. T. Wescott, L. R. Fisher and S. Hanna, J. Chem. Phys., 2002, 107 G. S. Rushbrooke, Introduction to Statistical Mechanics,
116, 23612369. Clarendon Press, Oxford, 1949, pp. 300304.
74 J. R. Henderson, J. Chem. Phys., 2002, 116, 50395045. 108 T. L. Hill, Statistical Mechanics, McGraw-Hill, New York, 1956,
75 N. Tsunekawa, H. Miyagawa, K. Kitamura and Y. Hiwatari, pp. 348353.
J. Chem. Phys., 2002, 116, 67256730. 109 J. N. Israelachvili, Intermolecular and Surface Forces: With
76 C. Czaplewski, S. Rodziewicz-Motowido, A. Liwo, D. R. Applications to Colloidal and Biological Systems, Academic Press,
Ripoll, R. J. Wawak and H. A. Scheraga, Protein Sci., 2000, 9, London, 1985, p. 105.
12351245. 110 A. E. Garca and G. Hummer, Proteins: Struct., Funct. Genet.,
77 C. D. Eads, J. Phys. Chem. B, 2000, 104, 66536661. 2000, 38, 261272.

Phys. Chem. Chem. Phys., 2003, 5, 30853093 3093

Das könnte Ihnen auch gefallen