You are on page 1of 10

Fluid Phase Equilibria 194–197 (2002) 859–868

High-pressure NMR studies on solvation structure in


supercritical carbon dioxide
Mitsuhiro Kanakubo a,d,∗ , Tatsuya Umecky b,d , Chee C. Liew c , Takafumi Aizawa a,d ,
Kiyotaka Hatakeda a,d , Yutaka Ikushima a,d
a
Supercritical Fluid Research Center, National Institute of Advanced Industrial Science and Technology (AIST), 4-2-1
Nigatake, Miyagino-ku, Sendai-shi, Miyagi 983-8551, Japan
b
Department of Chemistry, Tohoku University, Aramaki, Aoba-ku, Sendai-shi, Miyagi 980-8578, Japan
c
Research Institute for Computational Sciences, Advanced Industrial Science and Technology (AIST), 1-1-1 Umezono,
Tsukuba-shi, Ibaraki 305-8568, Japan
d
CREST, Japan Science and Technology Corporation (JST), 4-1-8 Honcho, Kawaguchi-shi, Saitama 332-0012, Japan
Received 20 May 2001; accepted 19 July 2001

Abstract
19 F
NMR chemical shifts of 2,3,4,5,6-pentafluorotoluene, chloropentafluorobenzene, and perfluoro(methyl-
cyclohexane) in dilute carbon dioxide solutions were precisely determined at a fixed temperature of 314.3 K over
a wide range of pressure from 1 to 35 MPa. We presented an analytical procedure to determine the local solvent
densities around the solutes by means of the solvent-induced chemical shifts. The local solvent densities around
the different fluorines of the solutes were discussed in comparison with the previous results of hexafluorobenzene
and octafluorotoluene in terms of the substitutional and aromatic ring effects. It has become apparent that the local
solvent densities are quite sensitive to the molecular structures of solutes. © 2002 Elsevier Science B.V. All rights
reserved.
Keywords: Supercritical carbon dioxide; High-pressure; NMR spectroscopy; Chemical shift; Solvation structure; Binary
mixture

1. Introduction

Supercritical fluids (SCFs), in particular consisting of carbon dioxide and water, are promising media for
environmentally acceptable chemical processes. Recently, SCFs have extensively been used for a variety
of applications, which are of great importance from not only scientific but also industrial viewpoints.
One of the specific features of SCFs, as compared with usual media, is to tune readily the solvent’s

*Corresponding author. Tel.: +81-22-237-5211; fax: +81-22-237-5224.
E-mail address: kanakubo@aist.go.jp (M. Kanakubo).

0378-3812/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 8 - 3 8 1 2 ( 0 1 ) 0 0 7 9 2 - 0
860 M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868

microscopic properties as well as macroscopic ones by adjusting pressure and temperature. The local
solvent environment around a solute molecule dissolved in SCF should be different from that expected
from the bulk solvent properties. In order to use and control SCF solvents without waste energy loss,
it is very important to understand the local solvent environment, which should be closely related to the
free energy of solvation, because a lot of chemical processes are influenced by the relatively short-ranged
solvent effects. Over one or two decades, many researchers have studied solvent-induced spectroscopic
shifts of probe molecules, for which solvent molecules in the vicinity of the solutes are responsible,
with varying SCF density by means of various experimental techniques [1–4]. At present, such solvent
inhomogeneities around solute molecules in SCFs have been accepted in general [5–8]: the local solvent
densities around solutes are usually greater than the bulk solvent densities, where the maximum of the
local solvent density augmentation was frequently observed at reduced densities of ρ/ρ c ≈ 0.6 within
reduced temperatures of T/T c < 1.1.
In order to investigate the local solvent environment in SCF solutions, we have developed several kinds
of high-pressure instruments to provide in situ observations of SCF solutions at molecular level [9–11].
Among the instruments, NMR spectroscopy is the most powerful tool to study solute–solvent interactions
because it has a wide application to any NMR active nuclei and the relaxation measurements give a lot of
information on translational and rotational dynamics of molecules [12–15]. And, moreover, the chemical
shifts are very sensitive to a change in local solvent structures [11,16–18]. Recently, with a precisely
temperature-controllable high-pressure NMR cell, we have proposed an analytical procedure to determine
the local solvent densities from solvent-induced chemical shifts [11,18]. The use of NMR chemical shifts
has several advantages as follows. The local solvent density can be directly obtained without any other
theoretical consideration of macroscopic parameters of solvent such as a relative permittivity and refractive
index, and only a single experiment can provide local solvent structures around different positions of a
molecule at once. In fact, by measuring solvent-induced 19 F NMR chemical shifts, we have found the
anisotropic solvation structure of octafluorotoluene molecule in supercritical carbon dioxide (sc-CO2 ),
where the substitution of CF3 group into C6 F6 clearly brings about the positive contribution to the local
solvent densities at any positions of the molecule and the local solvent densities at phenyl ring fluorines
slightly increase in the following order, p-F < m-F < o-F [18].
In the present work, we measure the solvent-induced 19 F NMR chemical shifts of 2,3,4,5,6-pentafluoro-
toluene (C6 F5 CH3 ), chloropentafluorobenzene (C6 F5 Cl), and perfluoro(methylcyclohexane) (C6 F11 CF3 ),
of which the molecular structures are shown in Fig. 1, in dilute carbon dioxide solutions over a wide
pressure range from 1 to 35 MPa. The local solvent densities around the solutes in sc-CO2 are discussed

Fig. 1. Molecular structures of solute molecules studied in the present work: (a) 2,3,4,5,6-pentafluorotoluene; (b) chloropentaflu-
orobenzene; (c) perfluoro(methylcyclohexane).
M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868 861

by comparing with the previous results [18] of hexafluorobenzene (C6 F6 ) and octafluorotoluene (C6 F5 CF3 )
in terms of the substitutional and aromatic ring effects.

2. Experimental section

2.1. High-pressure NMR experiments


19
F NMR chemical shifts of C6 F5 CH3 , C6 F5 Cl, and C6 F11 CF3 in CO2 were measured at 314.3 ± 0.2 K
over a pressure range of 1–35 MPa. A Varian Inova 500 spectrometer equipped with a standard 10 mm
probe was used to obtain high-resolution 19 F NMR spectra, where the resonance frequency was fixed
at approximately 470.3 MHz. A standard single pulse sequence of 45◦ pulse width was used for all
experiments. The number of scan times was 16–400, being dependent on the signal intensities. The
observed chemical shift (δ obs ) of each signal was expressed by setting the extrapolated zero-density value
as a reference, which was obtained by linearly regressing the data at low densities of ρ < 0.1 g cm−3 .
The pressure fluctuations of the system during the experiments were within ±0.007 MPa at pressures
of p < 20 MPa and within ±0.014 MPa at p ≥ 20 MPa. The absolute pressures were measured with a
precise digital indicator (Druck Ltd., DPI 145) that had been traceably calibrated. The geminal coupling
constants of 19 F signals of C6 F11 CF3 in CO2 remained a constant value of |285–300| Hz over the observed
pressure range. The vicinal and long-range coupling constants of 19 F signals of the solutes in CO2 were
not read out exactly because of the line broadening. The peak assignments were made on the basis of the
spectra in neat samples.
The sample solutions were prepared in the mixing vessel outside the magnet, and then were transported
into the high-pressure cell. The sample concentrations were less than 5 mmol dm−3 at p < 8.5 MPa
and less than 10 mmol dm−3 at p ≥ 8.5 MPa, which were determined from the signal integrals using
the preliminarily-made calibration curve. In the present experiments, no concentration dependence was
observed at any pressures under the above-mentioned conditions. It was assumed that the determined
chemical shifts can be regarded as the infinitely dilute values. The experimental procedures have previ-
ously been described in detail [11].

2.2. Bulk magnetic susceptibility corrections

The observed chemical shifts (δ obs ) were converted into purely solvent-induced chemical shifts (δ sol )
by using the following equation (expressed in cgs emu unit), which is applicable to the parallel external
magnetic field to the cylindrical sample [19,20]

δsol = δobs + {χ(ref) − χ(sam)} × 106 . (1)
3
Here, χ(ref) and χ(sam) are the volume magnetic susceptibilities of the reference and sample solutions,
respectively. In the present experiments, the reference was set to be the extrapolated zero-density value
and the infinitely dilute condition of the sample solutions was satisfied. Thus,
χ(ref) = 0, (2)
and
ρ
χ(sam) ≈ χ(CO2 ) = χm (CO2 ), (3)
M
862 M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868

where χ m (CO2 ) is the molar magnetic susceptibility of CO2 (−20.8 × 10−6 cm3 mol−1 ) [21], M the molar
mass of CO2 , and ρ is the density of CO2 [22].

2.3. Chemicals

C6 F5 CH3 , C6 F5 Cl, and C6 F11 CF3 purchased from Aldrich Chemical Co., Inc., were used without further
purification. Pure-grade carbon dioxide from Sumitomo Seika Chemicals Co., Ltd. was 99.9% in purity.

3. Analytical procedure

The solvent-induced chemical shift obtained is directly related to the shielding constant (σ sol ) arising
from solute–solvent interactions; δ sol = −106 σ sol . Usually, σ sol is expressed as a sum of three contribu-
tions independently arising from van der Waals and permanent electric interactions between solute and
solvent molecules, and magnetic anisotropy of solvent molecules [23]. However, in the present study,
we can simply relate σ sol to van der Waals contribution (σ W ) because the contribution from magnetic
anisotropy of solvent CO2 is expected to be negligibly small [24,25] and the permanent electric contri-
bution is less effective due to lack of dipole moment of CO2 molecule. Moreover, σ W for 19 F chemical
shifts is generally considered to be much larger than the other contributions [26].
The van der Waals contribution, σ w (r), in terms of a pair of interacting molecules is inversely propor-
tional to the intermolecular distance, r, to the sixth power [27,28]:
3αI
σW (r) = −B , (4)
(4πε0 )2 r 6
where α and I are the polarizability and the ionization energy of solvent molecule, respectively, ε 0 the
permittivity of a vacuum, and B is a constant. The intermolecular distance can be rewritten as the distance
between the solvent molecule and the nucleus of interest, if the observed nucleus is not located at the
center of molecule. Furthermore, we can give an explicit expression for δ sol with a radial distribution
function, g(r), by assuming that only a pairwise interaction is the predominant factor in σ sol [11,16,29]:

ρ
δsol = −10 σsol = −10 NA
6 6
σW (r)g(r)4πr 2 dr, (5)
M
where NA is the Avogadro’s constant.
In order to obtain the local solvent density (ρ l ) around a solute molecule from a solvent-induced
chemical shift, we express the density dependence of δ sol as a combination of the following two functions:
δsol = f (ρ) + g(ρ), (6)
where
f (ρ) = aρ, (7)
and 
    c−1
c − 1 −(c−1)/c ρ − ρmax c − 1 1/c
g(ρ) = b +
c d c
   1/c c   1/c
ρ − ρmax c−1 c−1 c−1
exp − + + with ρmax = d . (8)
d c c c
M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868 863

The former term of f(ρ) represents a monotonous change in the local solvent density in proportion to the
bulk solvent density. In the range of r/ρ c ∼ 2, it has been well known that the local solvent density should
approach the bulk solvent density [5–8]; hence, a linear increase of δ sol at higher ρ should be mainly
attributable to f(ρ). And, f(ρ) can be used as a reference line of the bulk solvent density at low densities: the
deviation from the reference should be a measure of the local solvent density augmentation or diminution.
On the other hand, the latter term of g(ρ) is called the Weibull line shape function. Although there is no
theoretical meaning for this function form, it well represents an asymmetric density dependence of the
local solvent density augmentation, which has a maximum value at ρ = ρ max .
The local solvent density augmentation is hereby defined as the excess local solvent density (ρlex ) by
subtracting the bulk solvent density (ρ) from the local solvent density (ρ l ).
ρlex = ρl − ρ, (9)
Because ρ l at ρ is given by δ sol /a, which is rewritten as ρ + g(ρ)/a by substituting Eq. (6) into δ sol , we
can obtain a simple expression for ρlex :
g(ρ)
ρlex = . (10)
a
The present analytical procedure of solvent-induced chemical shift is very practical, where we no longer
have to determine the unknown parameter B in Eq. (4).

4. Results and discussion

The solvent-induced 19 F chemical shifts (δ sol ) of C6 F5 CH3 , C6 F5 Cl, and C6 F11 CF3 in CO2 are plotted
against the density (ρ) of CO2 in Figs. 2–4, respectively, where the nuclei at different positions are
identified in Fig. 1. The solid curve for each δ sol represents the least-square fit of the data to Eq. (6). As
is seen from the figures, it is clearly found that the above-mentioned equation can reproduce the density

Fig. 2. Density dependence of solvent-induced 19 F chemical shifts of C6 F5 CH3 ((䊊), F␣ ; ( ), F␤ ; (䊐), F␥ ) in carbon dioxide at
314.3 K. The solid curve for each δ sol represents the least-square fit of the data to Eq. (6).
864 M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868

Fig. 3. Density dependence of solvent-induced 19 F chemical shifts of C6 F5 Cl ((䊊), F␣ ; ( ), F␤ ; (䊐), F␥ ) in carbon dioxide at
314.3 K. The solid curve for each δ sol represents the least-square fit of the data to Eq. (6).

dependence of δ sol very well. Here, the coefficients a, b, c, d, and ρ max for each δ sol versus ρ are given in
Table 1, together with the previous data of C6 F6 and C6 F5 CF3 [18].
The density dependence of ρlex , corresponding to g(ρ)/a, of each nucleus observed for C6 F5 CH3 ,
C6 F5 Cl, and C6 F11 CF3 is given in Figs. 5–7, respectively. We first pay attention to the maximum positions
of ρlex , i.e. ρ max . Values of ρ max range from 0.22 to 0.34 g cm−3 , which are not generally so far from the
previous results obtained by other experimental techniques [4–8]. The ρ max values of different nuclei in
each C6 F5 –X solute (X = F, Cl, CF3 , and CH3 ) remain virtually unchanged, and the substitutional effect
on ρ max in different C6 F5 –X solutes is found to be slight as well (see Table 1). Wada et al. [4] reported

Fig. 4. Density dependence of solvent-induced 19 F chemical shifts of C6 F11 CF3 ((䊉), CF3 ; (䊊), F␣ ; ( ), F␤ ; (䊐), F␥ ; (䉫), F␦ ;
(䉱), F⭴ ; (䊏), F␨ ; (䉬), F␩ ) in carbon dioxide at 314.3 K. The solid curve for each δ sol represents the least-square fit of the data
to Eq. (6).
M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868 865

Table 1
Values of the coefficients a, b, c, d, and ρ max and ρlex at ρ = ρ max for δ sol vs. ρ of C6 F5 CH3 , C6 F5 Cl, and C6 F11 CF3 in CO2 at
314.3 Ka
Nucleus Data point a (g−1 cm3 ) b c d (g cm−3 ) ρ max (g cm−3 ) ρlex (g cm−3 ) at ρ = ρ max
C6 F5 CH3 F␣ 35 4.14(3) 0.883(23) 2.36(7) 0.408(8) 0.323 0.213
F␤ 35 4.03(2) 0.807(19) 2.38(6) 0.401(7) 0.319 0.200
F␥ 35 3.72(2) 0.710(18) 2.40(7) 0.397(7) 0.317 0.191
C6 F5 Cl F␣ 36 3.73(3) 0.767(22) 2.33(7) 0.411(9) 0.322 0.206
F␤ 36 4.55(3) 0.968(27) 2.35(7) 0.418(9) 0.330 0.213
F␥ 36 4.54(3) 0.947(27) 2.40(7) 0.418(8) 0.334 0.208
C6 F11 CF3 CF3 35 2.62(2) 0.358(15) 2.20(10) 0.361(12) 0.274 0.137
F␣ 35 2.08(2) 0.313(23) 2.27(19) 0.293(18) 0.227 0.151
F␤ 35 2.55(2) 0.364(16) 2.16(11) 0.353(14) 0.265 0.143
F␥ 35 2.04(2) 0.295(22) 2.36(20) 0.283(16) 0.224 0.145
F␦ 35 3.08(2) 0.462(15) 2.26(8) 0.382(10) 0.295 0.150
F⭴ 35 2.39(2) 0.341(18) 2.15(13) 0.339(16) 0.253 0.143
F␨ 35 3.13(2) 0.462(15) 2.30(8) 0.383(9) 0.299 0.147
F␩ 35 2.42(2) 0.338(17) 2.17(13) 0.340(15) 0.256 0.140
C 6 F6 b 38 4.32(2) 0.656(14) 2.20(5) 0.407(8) 0.308 0.152
b
C6 F5 CF3 CF3 36 3.08(9) 0.506(6) 2.13(3) 0.429(5) 0.319 0.164
F␣ 36 3.16(2) 0.531(11) 2.09(5) 0.420(9) 0.308 0.168
F␤ 36 4.45(2) 0.741(13) 2.14(5) 0.423(8) 0.315 0.166
F␥ 36 4.90(2) 0.805(14) 2.15(5) 0.425(8) 0.317 0.164
a
The 95% confidence limit in the least significant digits is given in parentheses.
b
Cited from [18].

Fig. 5. Density dependence of excess local solvent density of C6 F5 CH3 in carbon dioxide at 314.3 K.
866 M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868

Fig. 6. Density dependence of excess local solvent density of C6 F5 Cl in carbon dioxide at 314.3 K.

the excess local solvent densities of mono-substituted benzene molecules in sc-CO2 by means of FT-IR
spectroscopy. They observed that the excess local solvent densities reach the maximum at almost the
same bulk density of 0.25 g cm−3 for any molecules. Although the absolute values of ρ max in the present
work are larger than that in Wada’s work [4], in both the studies ρ max is not significantly influenced
by the insertion of substituent. On the other hand, the ρ max values of C6 F11 CF3 , ranging from 0.22 to
0.30 g cm−3 , are smaller than those of C6 F5 –X solutes, and are uniquely dependent on the positions of
the nuclei. It is found that ρ max at axial position has a smaller value than that at equatorial one does in
each geminal pair of fluorines. This is the first experimental evidence, to our knowledge, to demonstrate
that the excess local solvent densities at different positions in a single molecule have the distinct maxima.
The degree of ρlex is discussed in terms of the substitutional effect, where the ρlex values at ρ = ρ max
are calculated in Table 1. In C6 F5 –X solutes, the insertion of substituent apparently makes a positive

Fig. 7. Density dependence of excess local solvent density of C6 F11 CF3 in carbon dioxide at 314.3 K.
M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868 867

contribution to ρlex at any positions of the phenyl fluorines. Moreover, the enhancement of ρlex in C6 F5 CH3
and C6 F5 Cl is larger than that in C6 F5 CF3 . This fact suggests that the substitutional effect on ρlex is
exerted with decreasing the symmetry of molecule in addition to a change in the electronic state caused
by the insertion of substituent. And, a comparison between C6 F5 CH3 and C6 F5 CF3 shows no evidence
of specific interaction between fluorine and CO2 , even though several authors insisted on the peculiarity
of fluorine–CO2 interaction [17,30].
The positional dependence of ρlex at different ␣, ␤, and ␥ fluorines can be also observed in C6 F5 CH3 ,
C6 F5 Cl, and C6 F5 CF3 molecules. Among them, ρlex of phenyl fluorines in C6 F5 CH3 is particularly sen-
sitive to the positions and increases in the following order, F␣ < F␤ < F␥ . This may be attributable to the
fact that the replacement of F by CH3 group produces more asymmetric environment in phenyl fluorines
than those by CF3 and Cl do. The similar tendency in ρlex at different positions is found in C6 F5 CF3 ;
however, in C6 F5 Cl ρlex at ␤ fluorine seems to be largest. At present, this discrepancy cannot be ratio-
nalized straightforwardly because the substitution concurrently brings about several changes in solute’s
properties in view of solute–solvent interactions: the insertion of substituent disturbs the electronic state
of molecule, gives rise to the dipole moment of entire molecule, and of course, leads to a change in the
geometrical structure.
The ρlex values of C6 F11 CF3 are relatively smaller than those of the similar analogue, C6 F5 CF3 . This
can be interpreted by a lack of aromaticity in C6 F11 CF3 . In fact, on the basis of molecular dynamics
simulations for infinitely dilute solutions of C6 H6 in sc-CO2 , it was found that CO2 molecules favorably
tend to be oriented parallel to the benzene plane [31]. On the other hand, the positional dependence
of ρlex in C6 F11 CF3 appears to be more remarkable than that in C6 F5 CF3 . The ρlex of axial fluorine at
ρ = ρ max has a smaller value than that of equatorial one does in ␦–ε and ␨–␩ geminal pairs, whereas the
difference of ρlex between ␤ and ␥ fluorines is not very significant. We found, as stated above, that ρlex of
axial position in C6 F11 CF3 approaches the maximum at lower densities than that of equatorial one does
in each geminal pair of fluorines. These two facts indicate that in the lower density range of ρ < ρ max
solvent CO2 molecules are preferentially solvated to a C6 F11 CF3 molecule in the axial direction, where the
fluorine atoms are overcrowded. In view of the geometrical structure of C6 F11 CF3 molecule, however, the
solvation space around the axial fluorines is exceedingly limited due to the overcrowding in comparison
with the case of equatorial fluorines, which results in inversion in ρlex at higher ρ (Fig. 7).

5. Conclusion

We demonstrated that solvent-induced NMR chemical shifts are very useful for studies of local solvent
structures in SCF solutions. With our developed analytical procedure of solvent-induced chemical shift,
the local solvent densities around C6 F5 CH3 , C6 F5 Cl, and C6 F11 CF3 in CO2 were determined at different
positions of fluorines. The excess local solvent densities around the solutes were discussed in comparison
with the previous results [18] of hexafluorobenzene and octafluorotoluene in terms of the substitutional
and aromatic ring effects. It has become apparent that the local solvent densities are quite sensitive
to molecular structures of solutes. A combination of other experimental and theoretical studies, for
example, molecular dynamics simulations, can provide more reliable pictures of solvation structures in
SCF solutions. In order to achieve environmentally acceptable SCF processes for a variety of applications,
a further systematic examination of solvation structures for various kinds of solute molecules is of great
importance.
868 M. Kanakubo et al. / Fluid Phase Equilibria 194–197 (2002) 859–868

Acknowledgement

M. Kanakubo is much indebted to Dr. H. Kawanami for his helpful discussion.

References

[1] Y. Sun, M.A. Fox, K.P. Johnston, J. Am. Chem. Soc. 114 (1992) 1187.
[2] R.S. Urdahl, D.J. Myers, K.D. Rector, P.H. Davis, M.D. Fayer, J. Chem. Phys. 107 (1997) 3747.
[3] J. Zhang, D.P. Roek, J.E. Chateauneuf, J.F. Brennecke, J. Am. Chem. Soc. 119 (1997) 9980.
[4] N. Wada, M. Saito, D. Kitada, R.L. Smith Jr., H. Inomata, K. Arai, S. Saito, J. Phys. Chem. B 101 (1997) 10918.
[5] S.C. Tucker, M.W. Maddox, J. Phys. Chem. B 102 (1998) 2437.
[6] O. Kajimoto, Chem. Rev. 99 (1999) 355.
[7] S.C. Tucker, Chem. Rev. 99 (1999) 391.
[8] W. Song, R. Biswas, M. Marocelli, J. Phys. Chem. B 104 (2000) 6924.
[9] Y. Ikushima, N. Saito, M. Arai, H.W. Blanch, J. Phys. Chem. 99 (1995) 8941.
[10] Y. Ikushima, K. Hatakeda, N. Saito, M. Arai, J. Chem. Phys. 108 (1998) 5855.
[11] M. Kanakubo, T. Aizawa, T. Kawakami, O. Sato, Y. Ikushima, K. Hatakeda, N. Saito, J. Phys. Chem. B 104 (2000) 2749.
[12] M. Kanakubo, H. Ikeuchi, G.P. Satô, H. Yokoyama, J. Phys. Chem. B 101 (1997) 3827.
[13] M. Kanakubo, H. Ikeuchi, G.P. Satô, J. Chem. Soc. Faraday Trans. 94 (1998) 3237.
[14] M. Kanakubo, C.C. Liew, T. Aizawa, T. Kawakami, O. Sato, Y. Ikushima, K. Hatakeda, N. Saito, Chem. Lett. (2000) 1320.
[15] C.R. Yonker, J. Phys. Chem. A 104 (2000) 685.
[16] D.M. Pfund, T.S. Zemanian, J.C. Linehan, J.L. Fulton, C.R. Yonker, J. Phys. Chem. 98 (1994) 11846.
[17] A. Dardin, J.M. DeSimone, E.T. Samulski, J. Phys. Chem. B 102 (1998) 1775.
[18] M. Kanakubo, T. Umecky, H. Kawanami, T. Aizawa, Y. Ikushima, Y. Masuda, Chem. Phys. Lett., in press.
[19] W.C. Dickinson, Phys. Rev. 81 (1951) 717.
[20] E.D. Becker, High-Resolution NMR Theory and Chemical Applications, 2nd Edition, Academic Press, New York, 1980
(Chapter 4).
[21] H.A. Landolt, R. Börnstein, Zahlenwerte und Functionen, 6th Edition, Vol. II, Springer, Berlin, Part 10, 1967.
[22] S. Angus, B. Armstrong, K.M. de Reuck, International Thermodynamic Table of the Fluid State-3 Carbon Dioxide, IUPAC,
Blackwell Scientific Publication, Oxford, 1976.
[23] A.D. Buckingham, T. Schaefer, W.G. Schneider, J. Chem. Phys. 32 (1960) 1227.
[24] Y. Lim, N.E. Nugara, A.D. King Jr., J. Phys. Chem. 97 (1993) 8816.
[25] J.C. Linehan, S.L. Wallen, C.R. Yonker, T.E. Bitterwolf, J.T. Bays, J. Am. Chem. Soc. 119 (1997) 10170.
[26] R. K. Harris, Nuclear Magnetic Resonance Spectroscopy A Physicochemical View, Pitman, London, 1983 (Chapter 8).
[27] A.A. Bothner-By, J. Mol. Spectrosc. 5 (1960) 52.
[28] W.T. Raynes, A.D. Buckingham, H.J. Bernstein, J. Chem. Phys. 36 (1962) 3481.
[29] N.J. Tappeniers, J.G. Oldenziel, Physica 82A (1976) 581.
[30] A. Cece, S.H. Jureller, J.L. Kerscher, K.F. Moschner, J. Phys. Chem. 100 (1996) 7435.
[31] H. Inomata, S. Saito, P. Debenedetti, Fluid Phase Equilib. 116 (1996) 282.

You might also like