You are on page 1of 6

Relationship between cohesive energy density and hydrophobicity

Giuseppe Graziano

Citation: The Journal of Chemical Physics 121, 1878 (2004); doi: 10.1063/1.1766291
View online: http://dx.doi.org/10.1063/1.1766291
View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/121/4?ver=pdfcov
Published by the AIP Publishing

Articles you may be interested in


The electrostatic response of water to neutral polar solutes: Implications for continuum solvent modeling
J. Chem. Phys. 138, 224504 (2013); 10.1063/1.4808376

Rotational dynamics of water and benzene controlled by anion field in ionic liquids: 1-butyl-3-methylimidazolium
chloride and hexafluorophosphate
J. Chem. Phys. 127, 104506 (2007); 10.1063/1.2768039

Nonperturbative vibrational energy relaxation effects on vibrational line shapes


J. Chem. Phys. 121, 11250 (2004); 10.1063/1.1812748

Properties of 2,2,2-trifluoroethanol and water mixtures


J. Chem. Phys. 114, 426 (2001); 10.1063/1.1330577

Solvent-induced interactions between hydrophobic and hydrophilic polyatomic sheets in water and hypothetical
nonpolar water
J. Chem. Phys. 106, 9781 (1997); 10.1063/1.473867

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18
JOURNAL OF CHEMICAL PHYSICS VOLUME 121, NUMBER 4 22 JULY 2004

Relationship between cohesive energy density and hydrophobicity


Giuseppe Grazianoa)
Dipartimento di Scienze Biologiche ed Ambientali, Universitá del Sannio, Via Port’Arsa,
11-82100 Benevento, Italy
共Received 1 April 2004; accepted 5 May 2004兲
It has been recently claimed that the large cohesive energy density of water is the ultimate cause of
the poor solubility of nonpolar compounds in water. In order to test the validity of this idea, we
analyze the difference in solubility between light water and heavy water of several nonpolar
compounds. Even though the cohesive energy density of D2 O is larger than that of H2 O, nonpolar
compounds are slightly more soluble in D2 O than H2 O. In such case experimental data do not
support the correctness of the large cohesive energy density as the ultimate cause of hydrophobicity.
We show that D2 O is a slightly better solvent than H2 O for nonpolar compounds because it is
slightly less costly to create a cavity in the former liquid. This is because there is slightly more void
volume in heavy water than in light water. © 2004 American Institute of Physics.
关DOI: 10.1063/1.1766291兴

I. INTRODUCTION D2 O, 12,13 and v ⫽18.073 cm3 mol⫺1 for H2 O and 18.136


Hydrophobicity is the term used to indicate the poor cm3 mol⫺1 for D2 O. 14 Thus, at 25 °C, ced(D2 O)
solubility of nonpolar compounds in water. Even though this ⫽2365 J cm⫺3 and ced(H2 O)⫽2297 J cm⫺3 . These num-
phenomenon is well recognized, its molecular origin is still bers indicate that D2 O has a ced larger than that of H2 O:
strongly debated in the scientific community.1– 4 In the last ced(D2 O)/ced(H2 O)⫽1.0296. This is not a surprise since it
few years, independently, Lazaridis5 and Kodaka,6 by means is generally recognized that the hydrogen bonds are enthal-
of different arguments, proposed that the large cohesive en- pically stronger in D2 O than H2 O: by 0.96 kJ mol⫺1 accord-
ergy density of water is the real cause of the poor solubility ing to Nemethy and Scheraga,15 0.93 kJ mol⫺1 according to
of nonpolar compounds. Actually, this explanation was first Marcus and Ben-Naim,11 and 0.84 kJ mol⫺1 according to
advanced by Hildebrand7 in a small note published in 1968, quantum mechanical calculations by Scheiner and Cuma.16
and subsequently supported by Privalov and Gill,8 and by The larger ced of D2 O, on the basis of the claim by Hilde-
Honig and colleagues.9 The cohesive energy density10 is de- brand, Lazaridis, and Kodaka, should imply that D2 O is a
fined by ced⫽(⌬ vapH⫺RT)/ v , where ⌬ vapH is the molar poorer solvent for nonpolar compounds than H2 O. In other
vaporization enthalpy change at the temperature T and v is words, the solubility of a nonpolar compound should be
the molar volume of the liquid at the same temperature. The lower in D2 O than H2 O.
ced of water is by far the largest among all common solvents However, experimental solubility measurements state the
as a consequence of both the strength of hydrogen bonds and contrary: the solubility of simple nonpolar molecules is
the smallness of the molar volume of water. On this basis larger in D2 O than H2 O. 11,17–19 The values of the
Hildebrand stated: ‘‘The fact that octane is nearly insoluble Ben-Naim20 standard Gibbs energy change associated with
in water is merely the result of the fact that van der Waals the transfer from light to heavy water, ⌬G "(H2 O→D2 O),
octane-water attraction is not strong enough to penetrate the listed in Table I, are all negative at room temperature. The
high cohesion of water.’’ numbers are small in view of the strict similarity of the two
In order to be considered right, this claim should be solvents, but the difference goes systematically into one di-
tested in a clear-cut way. This is the aim of the present work. rection.
We would like to show that the ced argument is not of gen- If hydrophobicity were proportional to the ced of water,
eral validity because it fails to rationalize the well-known the Ben-Naim standard Gibbs energy change associated with
experimental finding that simple nonpolar compounds are the transfer from the gas phase to heavy water could be cal-
more soluble in heavy water than in light water.11 culated by means of the following relationship:
⌬G "共 D2 O兲 calc⫽⌬G "共 H2 O兲关 ced共 D2 O兲 /ced共 H2 O兲兴 , 共1兲
II. RESULTS
where ⌬G "(H2 O) is the experimental Ben-Naim standard
In order to clarify the role played by ced in the hydro- Gibbs energy of hydration, whose values at 25 °C for the
phobicity phenomenon, we compare the solubility of several nonpolar solutes considered are listed in the fourth column of
nonpolar compounds in H2 O with that in D2 O, because light Table I. According to Eq. 共1兲, the values of ⌬G "(D2 O) calc are
and heavy water have different values of ced. At 25 °C, larger than those of ⌬G "(H2 O), so that their difference is
⌬ vapH⫽44.0 kJ mol⫺1 for H2 O and 45.4 kJ mol⫺1 for always a positive quantity 共see the numbers listed in brackets
in the fifth column of Table I兲, in complete contrast with
FAX: ⫹39/0824/23013; Electronic mail address: graziano@unisannio.it
a兲
experimental data 共compare the open squares with the filled

0021-9606/2004/121(4)/1878/5/$22.00 1878 © 2004 American Institute of Physics


This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18
J. Chem. Phys., Vol. 121, No. 4, 22 July 2004 Cohesive energy density and hydrophobicity 1879

TABLE I. Hard sphere diameter, polarizability, experimental Ben-Naim standard Gibbs energy of hydration at
25 °C, experimental Ben-Naim standard Gibbs energy of transfer from H2 O to D2 O at 25 °C for several
nonpolar compounds, and SPT estimates of ⌬G c (H2 O→D2 O). The numbers reported in brackets in the fifth
column are the ⌬G "(H2 O→D2 O) estimates obtained via Eq. 共1兲. The values of ␴ and ␣ come from Ref. 33,
except those of C3 H8 and n-C4 H10 , which come from Ref. 34. The values of ⌬G "(H2 O) and ⌬G "(H2 O
→D2 O) come from Table II of Ref. 11.

␴ ␣ ⌬G "(H2 O) ⌬G "(H2 O→D2 O) ⌬G c (H2 O→D2 O)


共Å兲 共Å3兲 共kJ mol⫺1兲 共J mol⫺1兲 共J mol⫺1兲

Ne 2.78 0.393 11 191a ⫺203a 共331兲 ⫺104


Ar 3.40 1.63 8400a ⫺245a 共248兲 ⫺145
8373b ⫺144b 共248兲 ⫺145
Kr 3.60 2.46 6922b ⫺138b 共204兲 ⫺160
CH4 3.70 2.70 8408b ⫺155b 共249兲 ⫺168
C2 H6 4.38 4.33 7682b ⫺129b 共227兲 ⫺225
C3 H8 5.06 6.29 8196c ⫺96c 共243兲 ⫺292
n-C4 H10 5.65 8.12 8715c ⫺84c 共258兲 ⫺357
O2 3.46 1.56 8639b ⫺196b 共256兲 ⫺149
N2 3.70 1.73 10 265b ⫺308b 共304兲 ⫺168
CH4 4.66 2.86 13 056b ⫺259b 共386兲 ⫺251
SF6 5.51 4.48 12 585b ⫺207b 共375兲 ⫺340
a
Reference 19.
b
Reference 18.
c
Reference 17.

ones in Fig. 1兲. Experimental ⌬G "(H2 O→D2 O) data do not two isotopic species of water have very close values of the
support the claim that the large ced of water is the ultimate dielectric constant, refractive index, and molecular dipole
cause of hydrophobicity. moment.15 Thus, one obtains
According to well-founded statistical mechanical
approaches,21–23 the Ben-Naim standard Gibbs energy ⌬G "共 H2 O→D2 O兲 ⬵⌬G c 共 D2 O兲 ⫺⌬G c 共 H2 O兲
change associated with the transfer from light to heavy water
⫽⌬G c 共 H2 O→D2 O兲 . 共3兲
of a nonpolar compound is given by
⌬G "共 H2 O→D2 O兲 ⫽ 关 ⌬G c 共 D2 O兲 ⫺⌬G c 共 H2 O兲兴 Equation 共3兲 means that the different solubility of a nonpolar
⫹ 关 E a 共 D2 O兲 ⫺E a 共 H2 O兲兴 , 共2兲 compound in D2 O with respect to H2 O should be due to the
difference in the work of cavity creation. The latter quantity
where ⌬G c is the work of cavity creation and E a is the can be reliably calculated by means of the analytical formula
energy of turning on solute-water van der Waals interactions. provided by scaled particle theory, SPT.24,25 In performing
As a first approximation, it is reliable to assume that the E a calculations we use the experimental density at 25 °C of
term is nearly the same in D2 O as well as H2 O, because the H2 O, 0.997 05 g cm⫺3, and D2 O, 1.104 45 g cm⫺3, as re-
ported by Kell;14 the pressure is fixed at 1 atm, as suggested
by several authors.25,26 For the size of both H2 O and D2 O
molecules, we use ␴ ⫽2.80 Å because this value is close to
the location of the first peak in the oxygen-oxygen radial
distribution function of both light and heavy water.27–29
Moreover, we30 have shown that, by assigning a diameter of
2.80 Å to water molecules, SPT is able to reproduce satis-
factorily both the cavity size distribution and the ⌬G c values
obtained from molecular dynamics simulations in TIP4P
water31 by Pohorille and Pratt.32
The calculated ⌬G c values are always large and positive
but, more importantly, prove to be slightly smaller in D2 O
than H2 O 共see the numbers listed in the last column of Table
I兲. The difference increases with the cavity size, passing from
⫺0.12 kJ mol⫺1 for a cavity of 3 Å diameter to ⫺0.40
kJ mol⫺1 for a cavity of 6 Å diameter. By plotting the experi-
FIG. 1. Experimental Ben-Naim standard Gibbs energy of transfer from mental ⌬G "(H2 O→D2 O) data versus the effective size33,34
H2 O to D2 O at 25 °C 共filled squares兲 of several nonpolar compounds 共from of the nonpolar compounds listed in Table I, and the SPT
left to right: Ne, Ar, O2 , Kr, CH4 , N2 , C2 H6 , CF4 , C3 H8 , SF6 , and estimates of ⌬G c (H2 O→D2 O) versus the cavity diameter, it
n-C4 H10) vs their effective diameter; SPT estimates of ⌬G c (H2 O→D2 O) at
25 °C as a function of the cavity diameter 共continuous line兲; calculated
is evident that the latter quantity more than qualitatively ac-
values of the Ben-Naim standard Gibbs energy of transfer 共open squares兲, counts for the experimental data 共compare the filled squares
where ⌬G(D2 O) calc has been obtained by means of Eq. 共1兲. with the continuous line in Fig. 1兲.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18
1880 J. Chem. Phys., Vol. 121, No. 4, 22 July 2004 Giuseppe Graziano

The SPT formula shows that the magnitude of ⌬G c , at a III. DISCUSSION


given temperature for a given cavity diameter, is determined The experimental finding that simple nonpolar com-
by two geometric properties of the solvent:35 the size of its pounds are slightly more soluble in D2 O than H2 O cannot be
molecules ␴, and its volume packing density ␰ reconciled with the idea that the large ced of water is the
⫽ ␲␴ 3 N Av/6v , 共i.e., the ratio of the physical volume of a ultimate cause of hydrophobicity. The ced of D2 O is larger
mole of solvent molecules over the molar volume of the than that of H2 O, so one should expect that the former liquid
solvent兲. The two variables work in opposite directions: for a would be a poorer solvent for nonpolar compounds than the
given ␴, ⌬G c increases on increasing ␰, and for a given ␰, latter. This expectation, however, is wrong. We already
⌬G c increases on decreasing ␴. This behavior can be readily pointed out that the ced argument cannot explain the trend of
explained. The quantity (1⫺ ␰ ) is the fraction of the total xenon solubility in n-alkanols at room temperature.43 In ad-
volume not occupied by solvent molecules. Thus, when ␰ dition, experimental data show that the volume packing den-
increases, the fraction of void volume diminishes and ⌬G c sity decreases while the ced increases on passing from light
increases. On the other hand, the diameter of the solvent to heavy water. Therefore, Kodaka’s suggestion6 of a propor-
molecules determines the size of the pieces in which the void tionality between the volume packing density of a solvent
volume is partitioned. Keeping fixed ␰, if the diameter of the and its ced does not seem to hold.
solvent molecules decreases, it diminishes the average size A qualitative rationalization of ⌬G "(H2 O→D2 O) data is
of the pieces in which the void volume is partitioned. This provided by approaches that dissect the hydration process in
implies that the number of molecular-sized cavities, those two consecutive steps: creation of the cavity and turning on
solute-water van der Waals interactions. The latter contribu-
important for the solvation of real solutes, becomes smaller
tion, as a first approximation, can be considered to be the
and the corresponding ⌬G c values increase. It is the small-
same in light and heavy water, at least for solutes having a
ness of water molecules to render ⌬G c extremely large and
small polarizability. Assuming that there is no difference in
positive in water with respect to the other common liquids,
size between D2 O and H2 O molecules, SPT calculations
and eventually to cause hydrophobicity.22,23,35–38 show that the work of cavity creation is slightly smaller in
For the pair H2 O-D2 O there is no difference in the size D2 O than H2 O because the former liquid is characterized by
of the molecules, but there is a small difference in the vol- a slightly smaller value of the volume packing density.
ume packing density: at 25 °C, ␰ (D2 O)⫽0.381 66 while Freely speaking, there is slightly more void volume in D2 O
␰ (H2 O)⫽0.382 98, because the molar volume of D2 O is than H2 O and this favors the insertion of nonpolar solutes.
slightly larger. As a consequence ⌬G c is smaller in D2 O than The assumption that there is no difference in size be-
H2 O, and the solubility of nonpolar compounds is slightly tween H2 O and D2 O molecules is an important point. Ac-
larger in heavy water than in light water. A similar conclu- cording to crystallographic data, the molecular dimensions
sion emerged in previous SPT analyses39– 41 of solubility dif- and the length of a hydrogen bond are identical in H2 O and
ferences between the two isotopic species of water. Since we D2 O ice within a few thousandths of an angstrom unit.13,15
are interested in the difference between the work of cavity Similarly, the first peak of the oxygen-oxygen radial distri-
creation in two liquids having molecules of identical size, the bution function in H2 O and D2 O occurs at the same
SPT results have to be considered robust and reliable. position.27 From a theoretical point of view a difference in
According to Fig. 1, the difference in the work of cavity size between H2 O and D2 O could be expected,13,44 because
creation fully accounts for the experimental ⌬G "(H2 O the amplitude of vibrations involving deuterium is smaller
→D2 O) data in the case of solutes having a small polariz- than that of vibrations involving protium, as a result of the
ability. On the other hand, Fig. 1 shows that the experimental larger mass of deuterium. However, the difference in size
⌬G "(H2 O→D2 O) values become less negative on increasing caused by this effect is very small13 共i.e., it is a second order
the size of the solute molecule and so its polarizability, at effect affecting the vibrational fluctuation about the average
odds with the trend of ⌬G c (H2 O→D2 O). This should be length of C-D and C-H bonds兲.
reliably caused by the attractive solute-water van der Waals On the basis of information theory results, Pratt and
colleagues45,46 suggested the existence of a proportionality
interactions. The volume packing density, the polarizability,
between the work of cavity creation in water and the recip-
and the dielectric constant of D2 O are slightly smaller than
rocal of its isothermal compressibility, ⌬G c ⬀1/␤ T . Since it
those of H2 O 共i.e., ␣ ⫽1.26 Å 3 for D2 O and 1.45 Å3 for H2 O
is generally assumed that ⌬G c ⬀ ␥ , where ␥ is the surface
in Ref. 42; see also Ref. 15兲, causing a reduction of the
tension of the liquid,47 and it has been recently shown that
strength of van der Waals interactions. This interpretation is the product ␥␤ T is a constant for several liquids at 25 °C,
supported by the finding11 that, at 25 °C, ⌬G "(H2 O→D2 O) including water,48 the relationship ⌬G c ⬀1/␤ T is expected.
⫽⫺129 J mol⫺1 for ethane ( ␣ ⫽4.3 Å 3 ), 0 for fluo- On this ground, Pratt and colleagues49 explained the slightly
romethane ( ␣ ⫽3.0 Å 3 and ␮ ⫽2.08 D), 70 J mol⫺1 for chlo- larger solubility of nonpolar compounds in D2 O with respect
romethane ( ␣ ⫽5.4 Å 3 and ␮ ⫽1.89 D), and 250 J mol⫺1 for to H2 O by recognizing that the former liquid is slightly more
bromomethane ( ␣ ⫽6.0 Å 3 and ␮ ⫽1.82 D). Clearly, the compressible than the latter: at 25 °C, ␤ T ⫽46.50
contribution of van der Waals interactions becomes more im- ⫻10⫺6 bar⫺1 for D2 O, and 45.25⫻10⫺6 bar⫺1 for H2 O. 50
portant in the Gibbs energy balance of Eq. 共2兲 on increasing Clearly, ␤ T is a macroscopic thermodynamic quantity and a
both the polarizability and the dipole moment of the solute microscopic interpretation of the difference between D2 O
molecule. and H2 O should be provided.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18
J. Chem. Phys., Vol. 121, No. 4, 22 July 2004 Cohesive energy density and hydrophobicity 1881

Pratt and colleagues45 suggested that the small value of At 25 °C, P int⫽141.8 J cm⫺3 for D2 O ( ␣ ⫽0.221
␤ T for H2 O with respect to the other common solvents ⫻10⫺3 K⫺1 ), and 169.5 J cm⫺3 for H2 O ( ␣ ⫽0.257
should be a reflection of the stiffness of its hydrogen bonding ⫻10⫺3 K⫺1 ). Thus, even though ced(D2 O)⬎ced(H2 O), it
network 共i.e., the strength of hydrogen bonds兲. This interpre- results P int(D2 O)⬍ P int(H2 O), confirming that ced and P int
tation, however, does not seem to hold. Even though the account for different properties of the liquids, especially
hydrogen bonds in D2 O are enthalpically stronger than those when hydrogen bonds are present.52 Following Dack’s pro-
in H2 O11,15,16 共see also Appendix A兲, ␤ T is larger for D2 O posal, one obtains ced(D2 O)⫺ P int(D2 O)⫽2223.2 J cm⫺3 ,
than for H2 O. and ced(H2 O)⫺ P int(H2 O)⫽2127.5 J cm⫺3 ; these numbers
The isothermal compressibility of a system is a measure indicate that hydrogen bonds are stronger in D2 O than H2 O
of the ensemble fluctuations in local number density:51 ␤ T by 95.7 J cm⫺3. If the latter number is multiplied by the
⫽( v /RT) 关 ( 具 ␳ 2 典 ⫺ 具 ␳ 典 2 )/ 具 ␳ 典 2 兴 , where the number density mean molar volume, 18.1 cm⫺3 mol⫺1, and divided by 2, the
␳ ⫽N Av / v , the angled brackets denote ensemble average val- mean number of hydrogen bonds per water molecule, one
ues, v is the molar volume, and R is the universal gas con- obtains that the difference in strength amounts to 866 J per
stant. The small value of ␤ T in the case of H2 O should be a mole of hydrogen bond at room temperature. This estimate
reflection of the fact that the molar volume of H2 O is signifi- agrees with those obtained by more rigorous
cantly smaller than that of common solvents: at 25 °C, v treatments.11,15,16 It is worth noting that, at 25 °C, a fraction
(in cm3 mol⫺1 )⫽18.07 for water, 89.41 for benzene, 108.75 of hydrogen bonds is broken;53 this fraction should be differ-
for c-hexane, 131.62 for n-hexane, and 97.09 for carbon ent in the two isotopic species of water,41,54 but such differ-
tetrachloride.14,33 The molecular origin of this is the fact that ence should be so small to be neglected in a simple calcula-
the size of water molecules is the smallest among all other tion as the present one.
solvents.25,33
We suggest that the isothermal compressibility of D2 O is APPENDIX B: RELATIONSHIP BETWEEN MOLAR
slightly larger than that of H2 O mainly because the molar VOLUME AND STRENGTH OF HYDROGEN BONDS
volume of D2 O is slightly larger than that of H2 O 共see also
Appendix B兲. Assuming that the molecular sizes of H2 O and In general, the molar volume of a liquid is determined by
D2 O are the same, this means that the volume packing den- the size of its molecules and the strength of intermolecular
sity of D2 O is slightly smaller than that of H2 O. In other interactions. In the case of water, however, characterized by
words, D2 O has a slightly larger void volume than H2 O and the presence of strongly directional interactions such as the
so it proves to be more compressible. It is evident that this hydrogen bonds, one has to make different considerations. A
explanation corresponds to that used to rationalize the find- perfect tetrahedral arrangement leads to an increase of molar
ing that ⌬G c is slightly smaller in D2 O than H2 O. volume. The transition from liquid water to ice at 0 °C is
The present analysis, even though rudimentary in some accompanied by an increase of molar volume,13 because
respects, is able to provide a rationalization of the experi- there is an increase in the number of molecules with a perfect
mental ⌬G "(H2 O→D2 O) data for several nonpolar com- tetrahedral coordination. The coordination number is 4 in ice
pounds. The latter represent a robust test to distinguish a whereas its average value in liquid water is ⬇5, due to the
physically correct interpretation of hydrophobicity from the presence of an ‘‘interstitial’’ molecule that is not hydrogen
noncorrect ones. Since the ⌬G "(H2 O→D2 O) data cannot be bonded to the central one.27–29 On this basis, the increase in
rationalized using the idea that the large ced of water is the molar volume passing from H2 O to D2 O and T2 O13 should
cause of hydrophobicity, this idea should not be considered reflect the increase in the strength of hydrogen bonds that, in
correct. The large unfavorable entropy change determining turn, leads to an increase in the number of molecules with
hydrophobicity around room temperature has nothing to do perfect tetrahedral coordination. In other words, the struc-
with the large ced of water. The process of cavity formation tural arrangement of D2 O and T2 O molecules in the liquid
provides the correct explanation: the negative entropy state is more icelike than that of H2 O molecules. Thus, the
change due to the excluded volume effect is exaggerated in increased strength of hydrogen bonds in D2 O with respect to
water by the small size of individual water molecules. H2 O manifests itself not only in the increase of ced but also
in the increase of molar volume. Since the molecular dimen-
sions are identical for H2 O and D2 O molecules at room tem-
ACKNOWLEDGMENTS
perature, the slightly larger molar volume of D2 O means that
This work was supported by Grant COFIN-2002 from there is slightly more void volume in D2 O than in H2 O. This
the Italian Ministry for Instruction, University and Research fact renders less costly the cavity creation in heavy water,
共M.I.U.R., Rome兲. favoring the solubility of nonpolar compounds.

APPENDIX A: AN ESTIMATE OF THE DIFFERENT 1


W. Blokzijl and J. B. F. N. Engberts, Angew. Chem., Int. Ed. Engl. 32,
STRENGTH OF HYDROGEN BONDS IN D2 O AND H2 O 1545 共1993兲.
2
L. R. Pratt, Annu. Rev. Phys. Chem. 53, 409 共2002兲.
Dack52 proposed that, in the case of liquids characterized 3
N. T. Southall, K. A. Dill, and A. D. J. Haymet, J. Phys. Chem. B 106, 521
by the presence of hydrogen bonds, the quantitative differ- 共2002兲.
4
B. Widom, P. Bhimalapuram, and K. Koga, Phys. Chem. Chem. Phys. 5,
ence between ced and P int⫽( ⳵ U/ ⳵ V) T ⬅ ␣ T/ ␤ T , where U is 3085 共2003兲.
the internal energy and ␣ is the coefficient of thermal expan- 5
T. Lazaridis, Acc. Chem. Res. 34, 931 共2001兲.
sion, should be a measure of the strength of hydrogen bonds. 6
M. Kodaka, J. Phys. Chem. B 105, 5592 共2001兲; 108, 1160 共2004兲.
This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18
1882 J. Chem. Phys., Vol. 121, No. 4, 22 July 2004 Giuseppe Graziano

7
J. H. Hildebrand, J. Phys. Chem. 72, 1841 共1968兲. Pratt and A. Pohorille, Proc. Natl. Acad. Sci. U.S.A. 89, 2995 共1992兲.
8
P. L. Privalov and S. J. Gill, Pure Appl. Chem. 61, 1097 共1989兲. 33
E. Wilhelm and R. Battino, J. Chem. Phys. 55, 4012 共1971兲.
9
B. Honig, K. Sharp, and A. S. Yang, J. Phys. Chem. 97, 1101 共1993兲. 34
G. Graziano, J. Chem. Soc., Faraday Trans. 94, 3345 共1998兲.
10
J. H. Hildebrand and R. L. Scott, Regular Solutions 共Prentice-Hall, Engle- 35
G. Graziano, J. Phys. Chem. B 106, 7713 共2002兲.
wood Cliffs, NJ, 1962兲. 36
M. Lucas, J. Phys. Chem. 80, 359 共1976兲; K. Soda, J. Phys. Soc. Jpn. 62,
11
Y. Marcus and A. Ben-Naim, J. Chem. Phys. 83, 4744 共1985兲. 1782 共1993兲.
12
F. D. Rossini, J. W. Knowlton, and H. L. Johnston, J. Res. Natl. Bur. 37
A. Wallqvist and D. G. Covell, Biophys. J. 71, 600 共1996兲; M. Ikeguchi, S.
Stand. 24, 369 共1940兲. Shimizu, S. Nakamura, and K. Shimizu, J. Phys. Chem. B 102, 5891
13
D. Eisenberg and W. Kauzmann, The Structure and Properties of Water 共1998兲.
共Oxford University Press, Oxford, 1969兲. 38
E. Gallicchio, M. M. Kubo, and R. M. Levy, J. Phys. Chem. B 104, 6271
14
G. S. Kell, J. Chem. Eng. Data 12, 66 共1967兲; 20, 97 共1975兲. 共2000兲.
15
G. Nemethy and H. A. Scheraga, J. Chem. Phys. 41, 680 共1964兲. 39
C. Jolicoeur and G. Lacroix, Can. J. Chem. 51, 3051 共1973兲; P. R. Philip
16
S. Scheiner and M. Cuma, J. Am. Chem. Soc. 118, 1511 共1996兲. and C. Jolicoeur, J. Solution Chem. 4, 105 共1975兲.
17
W. Wilhelm, R. Battino, and R. J. Wilcock, Chem. Rev. 共Washington, 40
R. Fernandez-Prini, R. Crovetto, M. L. Japas, and D. Laria, Acc. Chem.
D.C.兲 77, 219 共1977兲.
Res. 18, 207 共1985兲.
18
B. A. Cosgrove and J. Walkley, J. Chromatogr. 216, 161 共1981兲.
19
41
G. Graziano, J. Phys. Chem. B 104, 9249 共2000兲.
R. Crovetto, R. Fernandez-Prini, and M. L. Japas, J. Chem. Phys. 76, 1077 42
D. R. Lide, Handbook of Chemistry and Physics, 77th ed. 共CRC, Boca
共1982兲.
Raton, FL, 1996兲.
20
A. Ben-Naim, Solvation Thermodynamics 共Plenum, New York, 1987兲. 43
G. Graziano, Biophys. Chem. 105, 371 共2003兲.
21
L. R. Pratt and D. Chandler, J. Chem. Phys. 67, 3683 共1977兲; S. Garde, A. 44
E. Garcia, L. R. Pratt, and G. Hummer, Biophys. Chem. 78, 21 共1999兲. M. Turowski, N. Yamakawa, J. Meller, K. Kimata, T. Ikegami, K. Hosoya,
22
B. Lee, Biopolymers 24, 813 共1985兲; 31, 993 共1991兲. N. Tanaka, and E. R. Thornton, J. Am. Chem. Soc. 125, 13836 共2003兲.
45
23
G. Graziano, Biophys. Chem. 82, 69 共1999兲; G. Graziano, J. Phys. Chem. S. Garde, G. Hummer, A. E. Garcia, M. E. Paulaitis, and L. R. Pratt, Phys.
B 105, 2632 共2001兲; G. Graziano and B. Lee, ibid. 105, 10367 共2001兲; G. Rev. Lett. 77, 4966 共1996兲.
46
Graziano, Can. J. Chem. 80, 401 共2002兲. G. Hummer, S. Garde, A. E. Garcia, M. E. Paulaitis, and L. R. Pratt, J.
24
H. Reiss, Adv. Chem. Phys. 9, 1 共1966兲. Phys. Chem. B 102, 10469 共1998兲.
47
25
R. A. Pierotti, Chem. Rev. 共Washington, D.C.兲 76, 717 共1976兲. H. S. Ashbaugh and M. E. Paulaitis, J. Am. Chem. Soc. 123, 10721
26
F. H. Stillinger, J. Solution Chem. 2, 141 共1973兲; S. Shimizu, M. Ikeguchi, 共2001兲.
S. Nakamura, and K. Shimizu, J. Chem. Phys. 110, 2971 共1999兲.
48
G. R. Freeman and N. H. March, J. Chem. Phys. 109, 10521 共1998兲.
49
27
A. H. Narten and H. A. Levy, Science 165, 447 共1969兲; W. E. Thiessen, G. Hummer, S. Garde, A. E. Garcia, and L. R. Pratt, Chem. Phys. 258, 349
and A. H. Narten, J. Chem. Phys. 77, 2656 共1982兲. 共2000兲.
28
A. K. Soper, F. Bruni, and M. A. Ricci, J. Chem. Phys. 106, 247 共1997兲; A.
50
G. S. Kell, in Water: A Comprehensive Treatise, edited by F. Franks 共Ple-
K. Soper, Chem. Phys. Lett. 258, 121 共2000兲. num, New York, 1972兲, Vol. 1, p. 363.
29
J. M. Sorenson, G. Hura, R. M. Glaser, and T. Head-Gordon, J. Chem.
51
T. L. Hill, An Introduction to Statistical Thermodynamics 共Addison-
Phys. 113, 9149 共2000兲; T. Head-Gordon and G. Hura, Chem. Rev. 共Wash- Wesley, Reading, 1960兲.
ington, D.C.兲 102, 2651 共2002兲.
52
M. R. J. Dack, Chem. Soc. Rev. 4, 211 共1975兲.
30
G. Graziano, Biophys. Chem. 104, 393 共2003兲. 53
N. Muller, Acc. Chem. Res. 23, 23 共1990兲; B. Lee and G. Graziano, J. Am.
31
W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L. Chem. Soc. 118, 5163 共1996兲; K. A. T. Silverstein, A. D. J. Haymet, and
Klein, J. Chem. Phys. 79, 926 共1983兲. K. A. Dill, ibid. 122, 8037 共2000兲.
32
A. Pohorille and L. R. Pratt, J. Am. Chem. Soc. 112, 5066 共1990兲; L. R. 54
N. Muller, J. Solution Chem. 20, 660 共1991兲.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
140.254.219.138 On: Thu, 04 Dec 2014 02:33:18

You might also like