Você está na página 1de 12

J. Chem.

Thermodynamics 35 (2003) 1413–1424


www.elsevier.com/locate/jct

Thermodynamics of transfer of naphthalene


and 2-naphthoic acid from water to
(water + ethanol) mixtures at T ¼ 298:15 K
a,b,1 a,b,* c
Baoxue Zhou , Weimin Cai , Lizhuang Zou
a
Shool of Environmental Science and Engineering, Shanghai Jiao Tong University,
Shanghai 200240, PR China
b
Harbin Institute of Technology, Harbin 15000, PR China
c
School of Chemical and Environmental Engineering, China University of Mining and Technology,
Beijing 100083, PR China

Received 14 January 2003; accepted 28 April 2003

Abstract

Standard thermodynamic functions of transfer of naphthalene and 2-naphthoic acid from


water to (water + ethanol) mixtures at T ¼ 298:15 K have been determined from solubility
measurements at different temperatures. Standard free energies of transfer of both naphthalene
and 2-naphthoic acid showed decreasing tendency with the increasing x(EtOH), and the stan-
dard entropy and enthalpy of transfer exhibited a change of double peaks with x(EtOH). The
Dtr G0 of 2-naphthoic acid decreased more rapidly than that of naphthalene when x(EtOH)
< 0.746 and lower than that of naphthalene when x(EtOH) >0.746 at T ¼ 298:15 K. The dou-
ble peaks in the curves of standard entropy and enthalpy of transfer illustrated that the micro-
structure of the series of mixed solvents of (water + ethanol) underwent a variable process from
ordered to disordered and then from disordered to ordered. The results mean that there is a
relatively ordered structure near x(EtOH) ¼ 0.13 in the (water + ethanol) solutions besides
the existence of a clathrate structure in the water-rich region.
Ó 2003 Elsevier Science Ltd. All rights reserved.

Keywords: Naphthalene; 2-Naphthoic acid; Ethanol; Thermodynamic Functions of transfer; Intermolec-


ular interactions

*
Corresponding author. Tel.: +86-21-5474-7461; fax: +86-21-5474-0825.
E-mail addresses: zhoubaoxue@sohu.com (B. Zhou), wmcai@mail.sjtu.edu.cn (W. Cai).
1
Also corresponding author.

0021-9614/$ - see front matter Ó 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0021-9614(03)00112-5
1414 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

1. Introduction

Naphthalene is a typical nonpolar polycyclic aromatic hydrocarbon compound


(PAH). It is widely used as important chemical engineering material. It received
much attention as it was selected by the US environmental protection agency (US
EPA) as one of 129 types of priority pollutants in an aquatic environment. 2-Naph-
thoic acid is a typical polar derivative of PAH. It is widely used as important inter-
mediate of organic synthesis and plantgrowth regulator. It is also the main
metabolite of anaerobic degradation of naphthalene and 2-methylnaphthalene in
natural aquatic environment.
The thermodynamic properties of transfer play an important role for under-
standing the characteristics and interactions of intermolecular, intramolecular of
solute–solvent in various different media including biological media. These data
are also important for people to optimize the chemical engineering process and
to recognize the migration and enrichment process of pollutants in natural aquatic
environment and in vivo. Some thermodynamic properties including thermody-
namic properties of transfer of a number of monoaromatics and their derivants
in (water + monohydroxy alcohol) mixture, such as benzoic acid, p-nitroaniline
in (water + ethanol) mixture [1–3], benzene, toluene, trimethylbenzene in (water-
+ t-butyl alcohol) mixture [4,5] and inorganic compounds such as KB(ph)4 in
(water + t-butyl alcohol), (water + isopropanol) and (water + propanol) mixtures
are reported in literatures and our previous works [6–8]. However, the thermody-
namic properties of transfer for PAHs and their derivatives have received little
attention. Although thermodynamic properties in water-rich region, x(EtOH)
60.08, and the solubility data in (water + alcohol) mixture for naphthalene have
been reported in literature [9–11]. The thermodynamic properties of transfer on
the whole of (water + ethanol) region and the related interactions of intermolecular
and intramolecular in (water + ethanol) mixtures for naphthalene have not been re-
ported in the literature. No such studies are reported for 2-naphthoic acid in
(water + ethanol) mixtures. The purpose of this work is to find thermodynamic
functions of transfer for naphthalene and 2-naphthoic acid by measuring the tem-
perature dependence of their relative solubilities in water and (water + ethanol)
mixed solvents at various concentrations. With the experimental results, the effect
of intermolecular interactions on the thermodynamic properties of transfer for
naphthalene and 2-naphthoic acid in (water + ethanol) solutions is evaluated and
compared.

2. Experimental

Temperature control was carried out by means of a DZK-2 thermostatted water


bath specially made by Shanghai Jinghong Experimental Instrument. The fluctua-
tion of temperature was 0.05 K. The molar absorbance was determined with
a PCS UV spectrophotometer made by Shanghai Unico. Naphthalene (Shanghai
Yuanhang Reagent Factory, AR) and 2-naphthoic acid (US Acros Organics, grade
B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424 1415

puriss. P0.990 mass fraction) were recrystallized three times from absolute ethyl al-
cohol, respectively, before use. Ethanol (Shanghai Chemical Reagent Corporation,
China Medicine (Group), GR) was purified by standard methods of fractional dis-
tillation. The experimental values of the refractive index at T ¼ 293:15 K (20 °C)
(n20
D ) was 1.3614, in good agreement with literature data [12]. The solvents were pre-
pared by mixing weighed quantities of water and ethanol, and the mole fraction of
ethanol x(EtOH) was in the range of 0 to 0.8. Double-distilled water, treated with
ion exchange resin before distillation, was used and its conductivity was 1.18  104
S  m1 .
Saturated solutions of naphthalene and 2-naphthoic acid were obtained by fol-
lowing procedures: 30 mL of mixed solvents of different composition and a slight
excess amount of naphthalene and 2-naphthoic acid were added to different glass
tubes each with a ground mouth. Then the tubes were sealed with a tight-fitting
ground stopper and parafilm, followed by equilibration in a shaking thermostat
at T ¼ (298.15, 303.15, 308.15 and 313.15) K, respectively. The equilibrium time
was 10 to 14 d depending on the temperature. The selection of equilibrium time
was based on the condition test by following procedures: when maintained at con-
stant temperature for up to 3 to 4 days, the supernatant solution was analyzed by
spectrophotometry. This operation was repeated at intervals of 3 to 4 days until
there was no change of absorbance beyond the experimental error. After the condi-
tion test, the sealed glass tubes were not opened before the determination of solubil-
ities in order to avoid a tiny effect of volume change of solvent on experimental
results. The concentration of the saturated solutions of naphthalene in different
mixed solvents was measured by using ultraviolet visible spectrophotometry after
proper dilution with its respective solvents at 275 nm wavelength. The saturated sol-
ubilities of 2-naphthoic acid in series of mixtures were also analyzed by using ultra-
violet visible spectrophotometry just like the measurement of naphthalene at 280 nm
wavelength after an appropriate correction using a series of respective standard
curves. The method was also corrected by a different method of titrating with stan-
dard NaOH solution and phenolphthalein indicator. The two methods agree to
within 1%.

3. Results and discussion

Based on thermodynamic principles [1,3,6], the standard Gibbs free energies


change (DG0s ) of naphthalene or 2-naphthoic acid dissolving in a solvent at a certain
temperature can be calculated by using equation (1)

DG0s ¼ RT ln s; ð1Þ

where s is the saturated solubility of solute in mol  dm3 . The standard Gibbs free
energies of transfer (Dtr G0 ) for a solute from medium 1 to medium 2 can be calcu-
lated by using equation (2)
Dtr G0 ¼ RT lnðs2 =s1 Þ; ð2Þ
1416 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

where s1 and s2 in equation (2) are the saturated solubilities of solute in medium 1 to
medium 2. The functional relationship of Dtr G0 with the temperature T can be fitted
to a nonlinear equation. The selected function as follows [1,3]:
Dtr G0 ¼ a þ bT þ cT ln T ; ð3Þ
where a, b, c are constants independent of temperature, their values vary with the
composition and species etc., of mixed solvents.
Adding equations (2) and (3) produces equation (4)
RT lnðs2 =s1 Þ ¼ a þ bT þ cT ln T ; ð4Þ
then
lnðs2 =s1 Þ ¼ a=ðRT Þ  b=R  ðc=RÞ ln T : ð5Þ

Equation (5) can be regressed by a method of least squares based on the experi-
mental points of ln (s2 /s1 ) versus T. Designing a programme for equation (5) by using
Origin 6.0, the a, b, c and nonlinear correlation coefficient c can be obtained.
The standard free energy of transfer of solute from water to (water + ethanol)
mixtures is then calculated using equation (6)
Dtr G0 ¼ DG0s  DG0w ¼ ðas  aw Þ þ ðbs  bw ÞT þ ðcs  cw ÞT ln T ; ð6Þ
where subscripts s and w refer to the solvent mixtures and water, respectively. Ac-
cording to the functional relationship of Gibbs–Helmholtz, the standard entropy of
transfer (DStr0 ) can be derived with equation (7)
Dtr S 0 ¼ ðoDtr G0 =oT Þp : ð7Þ
Combination of equation (6) with (7) gives (8)
Dtr S 0 ¼ ðbw  bs Þ þ ðcw  cs Þð1 þ ln T Þ: ð8Þ
The standard enthalpy of transfer (Dtr H 0 ) can be calculated using equation (9) based
on the thermodynamic relationship DH 0 ¼ DG0 þ T DS 0
Dtr H 0 ¼ Dtr G0 þ T Dtr S 0 : ð9Þ
The solubilities of naphthalene and 2-naphthoic acid in various mixtures of
(water + ethanol) at T ¼ (298.15, 303.15, 308.15, and 313.15) K were presented in
table 1, figures 1 and 2, respectively. As shown in table 1, figures 1 and 2, on the
whole, the solubilities of both naphthalene and 2-naphthoic acid in (water + ethanol)
solution increase with increasing x(EtOH) very slowly at low x(EtOH) and sharply
at x(EtOH) P0.08. This implies that the effect of ethanol concentrations on the
solubilities (s) of naphthalene and 2-naphthoic acid is not remarkable at low con-
centration. However, when the concentration of ethanol exceeds a certain level of
0.08 of x(EtOH), the effect becomes much stronger. A similar result has been ob-
served in the (water + t-butyl alcohol) mixtures for benzene, toluene, trimethylben-
zene in our previous papers [4,5]. It also can be seen from table 1, figures 1 and 2 that
the replacement of the H atom of naphthalene in the 2 position by the COOH group
changes the solubilities considerably. The solubilities of 2-naphthoic acid are always
B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424 1417

TABLE 1
Solubilities of naphthalene and 2-naphthoic acid in water and (water + ethanol) solutions
xEtOH Naphthalene s/(104 mol  dm3 ) 2-Naphthoic acid s/(104 moldm3 )
T/K: 298.15 303.15 308.15 313.15 298.15 303.15 308.15 313.15
0 2.4198 2.9245 3.5028 4.1466 2.4345 2.9430 3.7366 4.8383
0.01 2.7325 3.3100 3.9607 4.6089 2.8133 3.5198 4.4948 5.8243
0.02 3.1019 3.8117 4.5675 5.4585 3.2203 4.0629 5.2146 6.7517
0.03 3.6152 4.4059 5.3799 6.2445 3.7222 4.7307 6.0896 7.8738
0.04 4.0213 5.1014 6.3422 7.7668 4.3644 5.5374 7.4315 9.6766
0.045 4.3537 5.7335 7.2100 8.8775 4.6365 6.1006 8.1203 10.474
0.05 4.5978 6.2369 7.8428 9.7683 4.9585 6.5699 8.8166 11.434
0.06 5.4583 7.4933 9.5581 11.679 5.9946 8.1059 10.630 13.798
0.08 7.5317 11.107 14.495 17.604 9.4992 13.373 18.248 23.875
0.10 11.277 16.092 20.608 25.694 17.396 22.253 30.065 40.478
0.20 99.241 142.52 176.39 214.03 194.18 238.55 297.20 348.62
0.40 691.99 964.22 1279.1 1626.6 1251.3 1400.5 1671.0 1873.6
0.60 1601.94 2090.0 2833.5 3572.8 2294.3 2534.9 2917.8 3200.0
0.80 3290.7 4331.2 5229.9 6344.9 2942.3 3312.7 3745.8 4056.4

FIGURE 1. Plot of solubility of naphthalene in (water + ethanol) solutions against mole fraction ethanol.
(––) T ¼ 298:15 K; (–s–) T ¼ 303:15 K; (–M–) T ¼ 308:15 K; (–O–) T ¼ 313:15 K.

higher than those of naphthalene in the ranges of xEtOH ¼0 to 0.746 for T ¼ 298:15
K, 0 to 0.680 for T ¼ 313:15 K, 0 to 0.616 for T ¼ 318:15 K, 0 to 0.526 for
T ¼ 323:15 K, respectively. However, over the ranges, lower solubilities for
2-naphthoic acid than those of naphthalene are observed. In addition, it also can be
seen from the above results that the maximum value of xEtOH for the solubility of
2-naphthoic acid is higher than that of naphthalene.
Coefficients a, b, c and nonlinear correlation coefficient c in equation (5) are given
in table 2. The data of nonlinear correlation coefficients c show that the nonlinear
regressions of ln (s2 /s1 ) versus T in various (water + ethanol) solvents using equation
(5) are reliable as reflected by the values of c > 0:99. Table 3 and figure 3 show the
variations of Dtr G0 of naphthalene and 2-naphthoic acid with x(EtOH) from water to
(water + ethanol) mixtures. From figure 3, it can be observed that, on the whole,
1418 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

FIGURE 2. Plot of solubility of 2-naphthoic acid in (water + ethanol) solutions against mole fraction
ethanol. (––) T ¼ 298:15 K; (–s–) T ¼ 303:15 K; (–M–) T ¼ 308:15 K; (–O–) T ¼ 313:15 K.

Dtr G0 of naphthalene and 2-naphthoic acid decrease with the increase of x(EtOH).
According to thermodynamic principles, the value of Dtr G0 indicates the ease of
spontaneous transfer of solutes from water to the mixed solvent. The larger the value
of Dtr G0 , the easier is the transfer of solutes from water to the mixed solvent. Hence
the larger the value of Dtr G0 , the more favourable is the distribution for solutes in
the mixed solvent. It can be concluded that the distribution of naphthalene and
2-naphthoic acid are more and more favourable, in general, with the increase of
x(EtOH) in the mixed solvent. From figure 3, it also can be seen that the curves of
Dtr G0 for both naphthalene and 2-naphthoic acid exhibit several different downward
trends. At the initial stages, Dtr G0 decreases slowly. When x(EtOH) rises above 0.08,
the Dtr G0 decreases more readily. When x(EtOH) rises to about 0.27, the decrease of
Dtr G0 slows again. This indicates that the intermolecular interactions among the sol-
ute, water and ethanol in the series of mixed solvent are not similar to each other.
As a comparison of naphthalene with 2-naphthoic acid, the changing trend of
Dtr G0 is slightly different. The Dtr G0 of 2-naphthoic acid decreases more rapidly than
that of naphthalene when x(EtOH) < 0.746, but more slowly for x(EtOH) > 0.746 at
T ¼ 298:15 K. This implies that the substitution product of naphthalene by the
COOH group is easier to transfer than is naphthalene when x(EtOH) < 0.746 and
more difficult to transfer than is naphthalene for x(EtOH) > 0.746 for changes from
low concentrations to high concentrations of ethanol at T ¼ 298:15 K. This phenom-
enon should be attributed to the result of direct formation of H-bonds of the hydro-
philic COOH group with water molecules in 2-naphthoic acid in low concentrations
of ethanol following Kundu [1], a formation of hydrogen-bonded solvent cage
referred as ‘‘skin phase’’, the formed H-bonds cause a greater stabilization of the
2-naphthoic acid molecule than that of naphthalene and more negative Dtr G0 values
than those of naphthalene in water and (water + EtOH) mixtures. However, along
with the decrease of mole fraction of water and the increase of x(EtOH), the
H-bonds between the carboxylic group and water become weak and the hydrophobic
interaction between ethanol molecules increase gradually. The existence of the polar
TABLE 2

B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424


Regression coefficients a, b, c of regression equation (5) and correlation coefficient c
xEtOH Naphthalene 2-Naphthoic acid
a b c c a b c c
1 1 1 1 1 1 1 1 1 1
(J  mol ) (J  mol K ) (J  mol K ) (J  mol ) (J  mol K ) (J  mol K )
0 70368.9 )958.924 139.03 1.000 )362195.8 8703.97 )1302.29 0.9998
0.01 162464.1 )3000.19 442.91 0.9999 )203943.7 5258.01 )790.86 1.0000
0.02 103747.6 )1671.98 244.17 0.9999 )162082.0 4348.32 )656.03 0.9999
0.03 201406.6 )3832.63 565.68 0.9991 )121030.3 3452.58 )523.20 0.9999
0.04 133288.6 )2232.88 324.85 1.0000 )166713.0 4509.26 )681.99 0.9990
0.045 281963.8 )5454.12 802.58 0.9999 51313.3 )274.42 29.16 0.9998
0.05 368730.2 )7328.00 1080.31 0.9994 56674.4 )372.79 43.17 0.9997
0.06 514560.8 )10527.60 1555.78 0.9999 153213.7 )2506.83 360.61 0.9999
0.08 878895.7 )18462.90 2733.60 0.9998 450697.1 )8939.67 1313.91 0.9999
0.10 600562.0 )12370.10 1827.47 0.9992 )287276.7 7194.30 )1084.30 0.9996
0.20 724061.8 )15165.40 2242.21 0.9965 185543.2 )3477.39 506.86 0.9986
0.40 390846.0 )7752.60 1134.49 1.0000 )34203.3 1171.66 )182.46 0.9918
0.60 116153.2 )1755.25 242.37 0.9983 15599.6 )1.36 )6.79 0.9938
0.80 356067.0 )7200.00 1055.70 0.9984 166315.6 )3334.66 489.16 0.9980

1419
1420 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

TABLE 3
Thermodynamic functions for transfer of naphthalene and 2-naphthoic acid from water to (water +
ethanol) solutions at T ¼ 298:15 K
xEtOH Naphthalene 2-Naphthoic acid
0 0 0
Dtr G Dtr S Dtr H Dtr G0 Dtr S 0 Dtr H 0
(kJ  mol1 ) (J  K1  mol1 ) (kJ  mol1 ) (kJ  mol1 ) (J K1 mol1 ) (kJ  mol1 )
0 0 0 0 0 0 0
0.01 )0.3013 6.0169 1.4926 )0.4442 20.5527 5.6836
0.02 )0.6157 8.8885 2.0344 )0.6935 27.2590 7.4338
0.03 )0.9953 16.1678 3.8251 )1.0526 33.3140 8.8800
0.04 )1.2593 29.4463 7.5201 )1.4472 40.1811 10.5328
0.045 )1.4563 51.0429 13.7622 )1.5008 60.8065 16.6287
0.05 )1.5887 64.7990 17.7311 )1.5926 65.3331 17.8865
0.06 )2.0168 79.8521 21.7911 )2.2341 73.2931 19.6182
0.08 )2.8150 126.7244 34.9679 )3.3754 121.3191 32.7959
0.10 )3.8158 102.6638 26.7934 )4.8755 49.6245 9.9200
0.20 )9.2077 120.2604 26.6480 )10.8568 64.3645 8.3335
0.40 )14.0194 126.4570 23.6839 )15.4760 32.1277 )5.8971
0.60 )16.1052 104.1860 14.9578 )16.9791 28.5490 )8.4672
0.80 )17.8883 101.6013 12.4041 )17.5959 40.1888 )5.6136

FIGURE 3. Plot of Dtr G0 from water to (water + EtOH) against mole fraction ethanol. (––) naphtha-
lene; (–s–) 2-naphthoic acid.

carboxylic group can enervate the hydrophobic interaction between the naphthalene
and ethanol molecules. Thus, the solubility of 2-naphthoic acid increase slowly and
Dtr G0 decreases gently at high x(EtOH). Particularly when x(EtOH) > 0.746 at
T ¼ 298:15 K, the solubility of 2-naphthoic acid is lower and Dtr G0 is higher than
that of naphthalene.
The standard entropy (Dtr S 0 ) and enthalpy (Dtr H 0 ) of transfer of naphthalene
and 2-naphthoic acid from water to (water + ethanol) mixtures are given in table 3
and figure 4. As illustrated in figure 4, the T Dtr S 0 curves for both naphthalene and
2-naphthoic acid show a change in the double peak curves with the increase
B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424 1421

FIGURE 4. Plot of T Dtr S 0 and Dtr H 0 from water to (water + EtOH) against mole fraction ethanol. (––)
T Dtr S 0 , naphthalene; (–s–) T Dtr S 0 , 2-naphthoic acid; (–M–) Dtr H 0 , naphthalene; (–O–) Dtr H 0 , 2-naphthoic
acid.

of x(EtOH). The first peaks appear at x(EtOH) ¼ 0.0838 for naphthalene and
x(EtOH) ¼ 0.0801 for 2-naphthoic acid. The second peaks appeared at
x(EtOH) ¼ 0.285 for naphthalene and x(EtOH) ¼ 0.267 for 2-naphthoic acid.
The change of thermodynamic functions of transfer depends on the change of
structure of mixed solvents. When a solute is introduced into solvent systems, a per-
turbation of the solvent structure and the solventÕs physical and chemical properties
would occur. However, this perturbation and change determines the variation of
macroscopic thermodynamic properties in the systems. In general, the transfer of
a solute from water to mixed solvent may be conceived to occur in four successive
steps. Firstly, the solute is removed from the water molecules. Secondly, the water
molecules reissue. (The cavity in the water is closed.) Thirdly, a new cavity is formed
in the co-solvent solution. Finally, the solute is solvated again by mixed solvents. It is
well known that the (water + monohydroxy alcohol) mixture has a high degree of
structure in the water-rich region [2,13–15], which makes it easy to form the clathrate
structure of the alcohol molecule surrounded by water molecules. For example, the
clathrate structure of ethanol and t-butyl alcohol surrounded by several water mol-
ecules in water-rich region is suggested in the literature [4,5,16]. However, as the size
of the alcohol molecule is different, the ratio of aqua alcohol in clathrate is different.
The smaller the size of alcohol molecule, the more difficult it is to determine the ex-
istence directly and to characterize the structure of clathrate. This perhaps is the
main reason why there are no reports in the literature where the structure of clath-
rate has been determined directly. Nevertheless, from the information of this exper-
iment, it is concluded that there is a clathrate structure in the (ethanol + water)
mixture in the water-rich region. Das [1] and Hu et al. [2] also suggested that there
is a clathrate structure (or solvent cage) in the (ethanol + water) mixture in the water-
rich region based on the studies of thermodynamics of the transfer of benzoic acid in
(water + ethanol) systems.
Ethanol is characterized by the presence of the hydrophilic and hydrophobic
groups leading to two different types of hydration. Thus both hydrophilic and hydro-
1422 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

phobic interactions compete for organization of the water structure. In the low con-
centration region of the alcohol (water-rich region), ‘‘free’’ ethanol is scarce due to
the existence of clathrate. Naphthalene can not enter the clathrate easily, which re-
sults in its solubilities increasing slowly with x(EtOH). In this process, the standard
free energy of transfer exhibits a slow change. Although 2-naphthoic acid has polar
carboxylic group leading the hydrophilic interaction (skin phase), the large hydro-
phobic group in 2-naphthoic acid makes it difficult to enter the clathrate. Subse-
quently, the clathrate structure broke down gradually with increasing x(EtOH),
and the ethanol molecules are removed from the clathrate. The probability of hydro-
phobic interactions occurrence between hydrophobic groups of ethanol and naph-
thalene or hydrophobic groups of 2-naphthoic acid would increase. This made the
solubilities increase sharply and continuously. The corresponding free energy for
transfer decreases quickly when x(EtOH) is above 0.08. When x(EtOH) rises to
about 0.27, the decreasing tendency of Dtr G0 became slow once again. The above
phenomenon indicated that the microstructure and intermolecular interactions
change in a different way in the series of mixed solvents. From the mathematical
point of view, the changing characteristics of Dtr G0 can be ultimately expressed by
the change of Dtr S 0 since the Dtr S 0 is dependent on the slop of Dtr G0 .
It is well known that Dtr S 0 can be used as a probe for elucidating the structure of
mixed solvents. Therefore, the introduction of alcohol would enhance the stabiliza-
tion of the clusters corresponding to a gradual organization into a water clatharate
surrounding an ethanol molecule. Initially, the water structure is little affected but
becomes more organized until it reaches a maximum structure. Beyond that point,
further addition of ethanol initiates a rapid structural collapse. A peak in the
T Dtr S 0 curves at x(EtOH) ¼ 0.0838 for naphthalene and at x(EtOH) ¼ 0.0801 for
2-naphthoic acid accompanies this process. Along with increasing x(EtOH), for
x(EtOH) up to 0.134 for naphthalene and 0.139 for 2-naphthoic acid, corresponding
a low valley of T Dtr S 0 , a relatively ordered structure appears in the mixed solvents.
This relatively ordered structure should be attributed to the maximum formation of
H-bonds between ethanol and water after the rupture of H-bonds between water and
water. After that, H2 O–EtOH H-bonds are replaced gradually by EtOH–EtOH
H-bonds and a maximum of T Dtr S 0 is observed again at x(EtOH) ¼ 0.285 for naph-
thalene and at x(EtOH) ¼ 0.267 for 2-naphthoic acid. These correspond to the for-
mation of a new relatively disordered structure (accompanied by inflection in the
curves of Dtr G0 near 0.27 for x(EtOH) in mixed solvents). This phenomenon, to some
degree, should be attributed to the maximum rupture of H-bonds between ethanol
and water though Dtr G0 decreased continuously. Henceforth, the mixed solvent sys-
tem enters the region of rich ethanol, and then the degree of EtOH–EtOH H-bond-
ing increases. The ordered degree of the solvent system increases gradually
accompanied by the increasing hydrophobic interaction of ethanol molecules. It is
noted that both enthalpy and entropy of 2-naphthoic acid decrease up to about
x(EtOH) ¼ 0.483, corresponding to the second low valley of T Dtr S 0 for 2-naphthoic
acid. Beyond that mole fraction, an increasing trend of T Dtr S 0 is observed in the
curve of 2-naphthoic acid. This result can also be attributed to the effect of the
substituted carboxylic group in molecule evidently. A similar result has been
B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424 1423

observed in the (water + EtOH) mixtures of benzoic acid [17]. It should be pointed
out that various studies, including those for some thermodynamic properties of
(water + ethanol) mixed solvents, indicate the maximum peak of Dtr S 0 in water-rich
region. However, only one peak of Dtr S 0 was found and the reported values of
x(EtOH) were much different. Kundu et al. [1,3] suggested a peak of Dtr S 0 at about
0.061 of x(EtOH) for p-nitroaniline and at about 0.144 of x(EtOH) for benzoic acid
in (water + ethanol) mixed solvents. Hu et al. [13] suggested a maximum Dtr S 0 region
of x(EtOH) about 0.1 to 0.15 for amino acid and benzoic acid [2], x(EtOH) > 0.2 for
aminobenzoic acid and x(EtOH) > 0.25 for glycine [14,15]. In comparison our work,
it is found that the points of experimental data with described in the literature for the
water-rich region are fewer than that of our work. Some small change in the series of
solvent system may be compensated by using the deficient data. Therefore, it is nec-
essary to obtain a numbers of data to confirm the region of x(EtOH) for this change.
It can be seen from our studies that there were at least two peaks of T Dtr S 0 in
(water + ethanol) for both naphthalene at x(EtOH) ¼ 0.0838, x(EtOH) ¼ 0.285, re-
spectively, and for 2-naphthoic acid at x(EtOH) ¼ 0.0801, x(EtOH) ¼ 0.267, corre-
sponding to two maximum disordered structures in the series solvent system. The
low values of T Dtr S 0 at x(EtOH) ¼ 0.134 for naphthalene and at x(EtOH) ¼ 0.139,
0.483 for 2-naphthoic acid, correspond to the maximum formation of newly ordered
structures in our solvent system and are observed in our studies. All of these results
have not yet been reported in literature.
It can be seen from these studies that both naphthalene and 2-naphthoic acid are
easily transferred from water to (water + ethanol) mixtures. However, the trends of
transfer for both naphthalene and 2-naphthoic acid are not significant in the
water-rich region. Analysis from the microstructure of mixed solvents in the
water-rich region shows that this phenomenon may be attributed to the effect of
clathrate formation. However, the solubilities increase dramatically and free energies
of transfer decrease with x(EtOH) when x(EtOH) > 0.08. This is obvious due to the
increase of hydrophobic interactions between the hydrophobic group of ethanol and
naphthalene or 2-naphthoic acid after a rapid structural collapse of the clathrate.
Subsequently, the solvent system undergoes a series of changes from ordered to dis-
ordered, and ordered again, though the change in trend of solubility is not obvious.
The replacement by polar a carboxylic group in the naphthalene molecule can in-
crease its solubility at low x(EtOH) and decrease its solubility at high x(EtOH).

Acknowledgement

This project was supported by China Postdoctoral Science Foundation ([2002]


No. 11).

References

[1] K. Das, A.K. Das, K. Bose, K.K. Kundu, J. Phys. Chem. 82 (1978) 1242–1245.
[2] X.G. Hu, R.S. Lin, H.S. Zong, Acta Phys.-Chim. Sin. 15 (9) (1999) 838–844.
1424 B. Zhou et al. / J. Chem. Thermodynamics 35 (2003) 1413–1424

[3] K.K. Bose, K. Kundu, Can J. Chem. 55 (1977) 3961–3966.


[4] L.Z. Zou, G.Y. Yang, B.X. Han, R.L. Liu, H.K. Yan, Sci. China Ser. B 42 (3) (1999) 225–235.
[5] L.Z. Zou, G.Y. Yang, B.X. Han, R.L. Liu, H.K. Yan, Sci. China Ser. B 42 (4) (1999) 400–410.
[6] X.L. Wang, L.Z. Zou, P.J. Zhang, F.G. Wang, Q.J. Yu, B.X. Zhou, J. Chem. Thermodyn. 31 (1999)
1609–1616.
[7] L.X. Cao, B.X. Zhou, P.F. Shi, L.Z. Zou, Acta Phys.-Chim. Sin. 17 (6) (2001) 565–569.
[8] L.X. Cao, B.X. Zhou, P.F. Shi, L.Z. Zou, Acta Phys.-Chim. Sin. 16 (11) (2000) 1053–1056.
[9] D. Bennett, W.J. Canady, J. Am. Chem. Soc. 106 (1984) 910–915.
[10] R.M. Dickhut, A.W. Andren, D.E. Armstrong, J. Chem. Eng. Data 34 (4) (1989) 438–443.
[11] P. Perez-Tejeda, C. Yanes, A. Meastre, J. Chem. Eng. Data 35 (1990) 244–246.
[12] R.C. Weast, CRC Handbook of Chemistry and Physics, 69th ed., CRC Press, Boca Raton, FL, 1988,
pp. E41–88.
[13] X.G. Hu, R.S. Lin, H.S. Zong, Acta Chim. Sin. 57 (5) (1999) 479–484.
[14] S.D. Chen, X.G. Hu, F.M. Pan, R.S. Lin, W.J. Fang, Acta Chim. Sin. 59 (7) (2001) 1089–1095.
[15] S.Q. Li, X.G. Hu, R.S. Lin, W.J. Fang, W.Q. Sang, Q.W. Liu, Acta Phys.-Chim. Sin. 17 (5) (2000)
400–405.
[16] S.K. Chakravorty, A. Pal, S.C. Lahiri, Thermochim. Acta 124 (1988) 43–52.
[17] A. Pal, S.C. Lahiri, Indian J. Chem. 28 (1989) 276–279.

JCT 03/011

Você também pode gostar