Densities, Viscosities, Refractive Indices, and ... - ACS Publications

Faculty of Chemistry, University of Mazandaran, P. O. Box 453, Babolsar, Iran. J. Chem. Eng. Data , 2017, 62 (1), pp 156–168. DOI: 10.1021/acs.jced...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/jced

Densities, Viscosities, Refractive Indices, and Excess Properties of Aqueous 1,2-Etanediol, 1,3-Propanediol, 1,4-Butanediol, and 1,5-Pentanediol Binary Mixtures Mehrdad Moosavi and Abbas Ali Rostami* Faculty of Chemistry, University of Mazandaran, P. O. Box 453, Babolsar, Iran S Supporting Information *

ABSTRACT: The thermodynamic and transport properties of aqueous alkanediol systems at atmospheric pressure were studied over the whole concentration range using the experimental values of density and refractive index at temperatures of 288.15−318.15 K with 5 K interval and viscosity at T = 293.15, 298.15, and 303.15 K. The obtained data were compared with those available in literature. A detailed analysis regarding the behavior of aqueous alkanediols and interactions between the mixture components was performed using different excess thermodynamic properties. Moreover, McAllister model and Grunberg−Nissan and its new proposed models were applied to predict the viscosity data. The validity of the proposed equations was assessed by comparisons between the experimental values and those obtained by the models.

1. INTRODUCTION Study on the thermophysical properties of liquid solutions has been carried out in many engineering research investigations, which is crucial in industrial design and extension of modeling.1−3 These properties are essential throughout the chemical production and transport processes to optimize the process and avoid the plant collapse.4 The correct values of the transport properties, such as viscosity and thermal conductivity of fluids at a given temperature, pressure, and composition, are required for the calculations of heat transfer, mass transfer, and contactor efficiency. When the experimental data are not accessible, it is necessary to have more accurate tools to predict the thermophysical properties. Different approaches, such as temperature and composition dependence of viscosity models,5 activity models,6,7 and equations of state,8 can be taken into account to model the thermophysical behavior. The corresponding predicted and correlated equations are dependent on numerous constant and adjustable parameters. Many developments have been performed to improve these models, and it is not easy to select the appropriate model for a particular case. To survey the inclination of proposed models from real behavior, comparison between experimental and theoretical results is indispensable. Indeed, the accuracy of the calculation is obviously not only dependent on the selection of an equation of state or the mixing rules but also on an adequately accurate illustration of pure data. Therefore, the direct measurement of thermodynamic and phase equilibrium data remains an important source of information. This work represents the study on the thermophysical properties of aqueous 1,2-ethanediol (ethylene glycol), 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol binary mixtures. © XXXX American Chemical Society

Its objectives are investigation of the effect of structural variation on different thermophysical properties; determination of the molecular interactions between the studied aqueous systems; and examination of different equations of state and models to predict the calculated thermophysical data. Additionally, the used alkanediols have huge applications in the automotive, aviation, explosive, textile, surface coating, food, cosmetic, pharmaceutical, tobacco, petroleum, and polymer industries.9,10 A survey of literature shows that numerous studies have been carried out on different thermophysical properties of alkanediol systems. Density measurements of pure 1,2-ethanediol, 1,2-propanediol, and 1,2-butanediol at temperatures from 278.15 to 358.15 K and pressures up to 60 MPa were reported by Atilhan and Aparicio.11 For all studied 1,2-alkanediols, densities increase with an increase of pressure and decrease with an increase of temperature. The pressure− temperature effects on the speeds of sound, densities, and related thermodynamic properties of 1,4-butanediol were studied by Zorebski and Dzida.12 The results show that the calculated values of thermal expansion coefficient and isothermal compressibility decrease with the increase of pressure and decrease of temperature. The heat capacity of 1,2-ethanediol, 1,3-propanediol. 1,4-butanediol, and 1,5-pentanediol were measured by Goralski et al.13 at different temperatures. For all alkanediols, the heat capacity increased when temperature increased, and also the slope for the plots of Received: June 26, 2016 Accepted: December 5, 2016

A

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Specifications of the Used Chemicals

a

compd

CAS no.

supplier

purity/% (supplier)

water content/% (supplier)

water content/% (K.F.)a

1,2-ethanediol 1,3-propanediol 1,4-butanediol 1,5-pentanediol

107-21-1 504-63-2 110-63-4 111-29-5

Merck109621 Merck807481 Merck801532 Merck807061

≥99.5 ≥98 ≥99 ≥98

≤0.1 ≤0.2 ≤0.3

≤0.15 ≤0.22 ≤0.34 ≤0.32

Water content determined by Karl Fischer method.

Table 2. Densities (ρ), Viscosities (η), and Refractive Indices (nD) of Pure Liquids with the Available Corresponding Literature Values at Different Temperatures and Pressure of 0.1 MPaa T = 288.15 K exptl lit. (NIST)15

T = 293.15 K

0.9992 0.9991

0.9983 0.9982

exptl lit. (NIST)16

0.949 1.001

exptl lit. (NIST)17,18 1,2-ethanediol T/K

exptl

lit.

1.3338

AAD%

T = 298.15 K

T = 303.15 K

T = 308.15 K

T = 313.15 K

T = 318.15 K

0.9957 0.9956

0.9940 0.9940

0.9922 0.9921

0.9908 0.9902

Water Density/(g·cm−3) 0.9971 0.9970 Water Viscosity/(mPa·s) 0.843 0.890 Water Refractive Index 1.3325

1.3331 1.3333 1,3-propanediol exptl

lit.

AAD%

exptl

0.755 0.797 1.3318 1.3312 1.3318 1,4-butanediol lit.

AAD%

1.3305 1.3299 1.3305 1,5-pentanediol exptl

lit.

AAD%

‑3

1.115819 1.116813 1.113510 1.113313 1.113221

0.089 0.000 0.017 0.000 0.008

1.0562

1.055920

ρ/(g·cm ) 0.028

1.0532

0.045 0.045 0.117 0.054 0.072 0.045 0.000 0.036 0.054 0.126

1.0499

0.037 0.056 0.018 0.007 0.028 0.028 0.000 0.019 0.000 0.019 0.028 0.019 0.009 0.000

1.0163

1.109810 1.109826 1.109019 1.109713 1.109521 1.109827 1.106810 1.106426 1.106213 1.105421

1.053610 1.053813 1.053022 1.352820 1.050210 1.049612 1.049922 1.049720 1.049928 1.049729 1.047010 1.046513 1.046620 1.046729

1.1029

1.103210 1.102926 1.102219 1.102521 1.102827

0.027 0.000 0.063 0.036 0.009

1.0435

1.0991

1.099810 1.099226 1.099013 1.098321

0.063 0.009 0.009 0.072

1.0404

0.038 0.325 0.009 0.009 0.019 0.000 0.019 0.009 0.019

1.0069

313.15

1.043910 1.040113 1.043622 1.043420 1.043728 1.043529 1.040610 1.040320 1.040229

318.15

1.0963

1.096210 1.095219

0.009 0.100

1.0379

1.037610 1.037120 1.037328 1.037529

0.028 0.077 0.057 0.038

1.0008

293.15

21.13

52.0120

η/(mPa·s) 0.892

17.24

0.28 1.41 1.27 0.58

51.55

298.15

21.1933 20.8334 20.8621 17.1433

40.80

41.1120

0.75

288.15

1.1168

293.15

1.1133

298.15

1.1103

303.15

1.1068

308.15

1.0467

B

1.0186

0.9922 1.016310 1.016113 1.015723 1.015624 1.013010 1.012723 1.012528 1.012624

0.000 0.019 0.059 0.068 0.009 0.019 0.039 0.029

0.9892

0.992725

0.353

0.9862

0.987830

0.162

1.009910 1.009913 1.009723 1.011331 1.009624 1.006710 1.006723 1.006428 1.006431 1.006724

0.000 0.000 0.019 0.138 0.029 0.019 0.019 0.049 0.049 0.019

0.9832

0.986725

0.355

0.9803

0.974332

0.612

1.003510 1.003513 1.003723 1.004431 1.003724 1.000710 1.000623 1.000728 1.000731 1.000724

0.029 0.029 0.049 0.119 0.049 0.009 0.019 0.009 0.009 0.009

0.9774

0.977314 0.981225

0.010 2.192

0.9748

0.974414

0.041

90.70

90.3235 97.2636

0.418 7.232

118.7

69.86

71.1435

1.83

86.79

1.0129

1.0099

1.0032

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 2. continued 1,2-ethanediol T/K

exptl

303.15

14.23

288.15

1.4320

293.15

1.4306

298.15

1.4294

303.15

1.4282

308.15

1.4269

313.15

1.4257

318.15

1.4244

lit. 18.6821 17.1233 14.0233 13.6434 13.8621

1,3-propanediol AAD% 8.35 0.69 1.47 4.14 2.60

exptl

32.74

1.4407 1.431538 1.431921 1.430438 1.429241 1.43042 1.428938 1.428241 1.428542 1.428721 1.427638 1.427441 1.425842 1.425421

0.062 0.090 0.069 0.013 0.041 0.049 0.00 0.021 0.035 0.049 0.035 0.007 0.021

1.4393 1.4380

1.4367

1.4353 1.4340 1.4327

lit.

1,4-butanediol AAD%

1.440720 1.440839 1.439320 1.439340 1.437920 1.437939 1.438229 1.436520 1.436739 1.436829

η/(mPa·s) 1.79 2.18 0.42 16.79 2.53 nD 0.00 0.006 0.00 0.00 0.006 0.006 0.013 0.013 0.00 0.006

1.435120 1.435429 1.433720 1.434029 1.433220 1.432729

0.013 0.006 0.020 0.00 0.034 0.00

40.0628 41.6937 32.8820 27.2428 33.5737

1,5-pentanediol

exptl

lit.

AAD%

55.24

72.7436 72.3237 55.9835 56.8836 56.3337

4.12 3.52 1.33 2.96 1.97

1.4470

exptl

1.4503 1.445140

0.027

1.4490

1.4441

1.445743 1.443140

0.110 0.069

1.4478

1.4426

1.441140

0.103

1.4466

1.4412

1.442243 1.439140 1.437240

0.069 0.145 0.180

1.4454

1.438643 1.435240

0.013 0.222

1.4430

1.4384

AAD%

68.3

1.4455

1.4398

lit.

1.4441

a Standard uncertainties are u(T) = 0.01 K, u(p) = 10 kPa, and u(x1) = 0.0002; relative standard uncertainties are ur(ρ) = 0.002, ur(η) = 0.08, and ur(nD) = 0.01

excess results. Furthermore, some proposed models including the Redlich−Kister and McAlister were used to correlate and predict the obtained thermophysical data by driving adjustable parameters.

heat capacity against temperature increased with the increment of the number of carbon atoms in alkanediols. Zemankova et al.14 obtained the negative excess molar volume sand positive excess heat capacities for some aqueous alkanediol systems at different temperatures. Both excess molar volume and excess heat capacity increase with an increase of temperature. The volumetric and viscometric properties as well as excess enthalpy of mixing for ammonium-based ionic liquid + alkanediol systems were studied by Domanska and co-workers.15 The results indicated that the densities and viscosities increase with an increase of IL mole fraction and decrease of temperature. Moreover, they obtained the positive values of excess molar volume and excess enthalpy with the negative viscosity deviations. Kijevcanin et al.17 reported the densities, refractive indices, and dynamic viscosities of butyl lactate + 1,2-propanediol or 1,3-propanediol binary mixtures at different temperatures and at atmospheric pressure. For both binaries, all experimental values decrease with an increase of butyl lactate concentration and also the obtained negative excess molar volumes of butyl lactate + 1,3-propanediol mixtures and positive for butyl lactate + 1,2-propanediol indicated that the position of the hydroxyl group in alcohols has significant impact on the behavior of mixtures. In this work, densities, viscosities, and refractive indices for the binary liquid mixtures of water + 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol were measured at atmospheric pressure and the temperature range of 288.15−318.15 K over the whole composition range. Different thermodynamic and transport properties, such as excess molar volume, partial molar volume, excess refractive index, and molar refraction, were calculated from the experimental data. The intermolecular interactions can be discussed using the

2. EXPERIMENTAL SECTION 2.1. Chemicals. All the materials used in this study were supplied by Merck and used without further purification. The quality of these used substances was verified with a gas chromatograph (GC) equipped with a flame ionization detector (FID). The GC did not show significant impurities, and therefore, the resulting purity values for all the substances coincided with those indicated by the manufacturer. The GC spectra are indicated is Figure S1 in the Supporting Information. Moreover, the FTIR spectra of pure alkanediols are shown in Figure S2 in the Supporting Information. Furthermore, the specifications of the chemicals used are listed in Table 1. The comparison of experimental density, viscosity, and refractive index values of pure 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol with those reported in the literature, obtained from NIST archive at all studied temperatures, are given in Table 2. Accordingly, the average absolute percentage deviation of density values of 1,2ethanediol, 1,3-propanediol, and 1,4-butanediol from literature is less than 0.2. However, for 1,5-pentanediol valid reported data in literature are rare and only the density values reported by Domanska et al.15 are in agreement with our data with the maximum average absolute percentage deviation of 0.041. It should be mentioned that due to the low purity of 1,5-pentanediol (98%) the density values of this alkanediol are more uncertain than the other used diols. The calculated relative standard uncertainty for density is 0.002. A survey of C

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

a

D

1.1168 1.1147 1.1136 1.1119 1.1103 1.1083 1.1068 1.1044 1.1018 1.098 1.0945 1.0906 1.0857 1.0798 1.0726 1.0630 1.0526 1.0364 0.9992

1.0562 1.0562 1.0562 1.0564 1.0565 1.0566 1.0564 1.0560 1.0556 1.0547 1.0537 1.0521 1.0499 1.0463 1.0432 1.0383 1.0321 1.0235 0.9992

0.0000 0.1037 0.1593 0.2189 0.2669 0.3165 0.3496 0.3994 0.4449 0.5061 0.5563 0.6055 0.6518 0.7001 0.7482 0.8015 0.8444 0.9010 1.0000

0.0000 0.1625 0.1737 0.2386 0.2519 0.3216 0.3669 0.4238 0.4553 0.5075 0.5501 0.6009 0.6530 0.7146 0.7532 0.8040 0.8508 0.9021 1.0000

1.0532 1.0531 1.0531 1.0533 1.0534 1.0535 1.0533 1.0530 1.0526 1.0518 1.0508 1.0492 1.0471 1.0435 1.0405 1.0355 1.0296 1.0213 0.9983

1.1133 1.1111 1.1100 1.1083 1.1068 1.1048 1.1034 1.1011 1.0985 1.0948 1.0914 1.0874 1.0826 1.0768 1.0697 1.0601 1.0502 1.0346 0.9983

293.15 K

308.15 K 1.1029 1.1005 1.0993 1.0976 1.0960 1.0942 1.0929 1.0906 1.0884 1.0847 1.0813 1.0776 1.0729 1.0673 1.0606 1.0515 1.0426 1.0280 0.9940

1.0435 1.0431 1.0431 1.0432 1.0433 1.0435 1.0433 1.0431 1.0428 1.0421 1.0412 1.0397 1.0376 1.0340 1.0311 1.0266 1.0213 1.014 0.994

303.15 K

1,2-Ethanediol 1.1103 1.1068 1.1080 1.1044 1.1069 1.1033 1.1052 1.1016 1.1036 1.1001 1.1017 1.0982 1.1003 1.0968 1.0980 1.0944 1.0956 1.0922 1.0918 1.0885 1.0884 1.0851 1.0845 1.0813 1.0797 1.0765 1.0737 1.0706 1.0667 1.0637 1.0574 1.0545 1.0480 1.0455 1.0328 1.0307 0.9971 0.9957 1,3-Propanediol 1.0499 1.0467 1.0497 1.0464 1.0497 1.0464 1.0499 1.0466 1.0499 1.0466 1.0501 1.0468 1.0500 1.0467 1.0496 1.0464 1.0493 1.0461 1.0485 1.0453 1.0476 1.0444 1.0461 1.0429 1.0439 1.0408 1.0404 1.0372 1.0374 1.0343 1.0326 1.0296 1.0269 1.0241 1.0189 1.0166 0.9971 0.9957

298.15 K

1.0404 1.0399 1.0400 1.0401 1.0401 1.0402 1.0401 1.0399 1.0396 1.0389 1.0380 1.0365 1.0344 1.0309 1.0281 1.0235 1.0184 1.0114 0.9922

1.0991 1.0967 1.0954 1.0937 1.0921 1.0903 1.0891 1.0869 1.0846 1.0811 1.0778 1.0740 1.0695 1.0640 1.0574 1.0484 1.0396 1.0254 0.9922

313.15 K

1.0379 1.0374 1.0374 1.0375 1.0375 1.0376 1.0375 1.0373 1.0370 1.0362 1.0354 1.0339 1.0318 1.0284 1.0255 1.0211 1.0161 1.0092 0.9908

1.0963 1.0938 1.0924 1.0907 1.0889 1.0873 1.0860 1.0839 1.0816 1.0781 1.0746 1.0711 1.0663 1.0607 1.0543 1.0456 1.0371 1.0235 0.9908

318.15 K

0.0000 0.1326 0.2201 0.2345 0.3001 0.4060 0.4969 0.5262 0.5488 0.6032 0.7012 0.7745 0.8051 0.8998 1.0000

0.0000 0.0992 0.1520 0.2039 0.2513 0.3003 0.3534 0.4399 0.4482 0.5011 0.5515 0.6114 0.6527 0.7005 0.7501 0.7998 0.8492 0.9001 0.9488 1.0000

x1

0.9922 0.9952 0.9970 0.9973 0.9986 1.0007 1.0026 1.0032 1.0037 1.0048 1.0066 1.0071 1.0070 1.0053 0.9992

1.0186 1.0208 1.0218 1.0228 1.0236 1.0245 1.0253 1.0262 1.0266 1.0272 1.0276 1.0277 1.0276 1.0272 1.0263 1.0242 1.0213 1.0159 1.0081 0.9992

288.15 K

Standard uncertainties are u(T) = 0.01 K, u(p) = 10 kPa, and u(x1) = 0.0002; relative standard uncertainty is ur(ρ) = 0.002

288.15 K

x1

0.9892 0.9921 0.9939 0.9942 0.9955 0.9976 0.9995 1.0002 1.0007 1.0019 1.0038 1.0046 1.0046 1.0032 0.9983

1.0163 1.0183 1.0193 1.0203 1.0211 1.0219 1.0227 1.0236 1.0240 1.0245 1.0250 1.0251 1.0250 1.0245 1.0236 1.0215 1.0189 1.0141 1.0067 0.9983

293.15 K

303.15 K

1,4-Butanediol 1.0129 1.0099 1.0148 1.0118 1.0158 1.0128 1.0167 1.0137 1.0176 1.0145 1.0183 1.0152 1.0191 1.0159 1.0200 1.0168 1.0203 1.0171 1.0208 1.0177 1.0211 1.0179 1.0213 1.0182 1.0213 1.0181 1.0210 1.0178 1.0201 1.0169 1.0183 1.0153 1.0158 1.0129 1.0117 1.0092 1.0047 1.0033 0.9971 0.9957 1,5-Pentanediol 0.9862 0.9832 0.989 0.9859 0.9908 0.9877 0.9910 0.9880 0.9924 0.9893 0.9945 0.9914 0.9964 0.9934 0.9971 0.9941 0.9977 0.9947 0.9989 0.9960 1.0009 0.9980 1.0019 0.9991 1.0021 0.9995 1.0009 0.9986 0.9971 0.9957

298.15 K

0.9803 0.9829 0.9847 0.9850 0.9863 0.9885 0.9905 0.9912 0.9918 0.9931 0.9952 0.9964 0.9967 0.9963 0.9940

1.0069 1.0089 1.0098 1.0106 1.0114 1.0121 1.0128 1.0136 1.0139 1.0145 1.0148 1.0151 1.0151 1.0148 1.0139 1.0125 1.0102 1.0067 1.0011 0.9940

308.15 K

0.9774 0.9799 0.9817 0.9820 0.9833 0.9855 0.9876 0.9883 0.9889 0.9902 0.9923 0.9937 0.9941 0.9939 0.9922

1.0032 1.0052 1.0061 1.0069 1.0076 1.0083 1.0090 1.0098 1.0101 1.0106 1.0109 1.0112 1.0112 1.0110 1.0103 1.0089 1.0071 1.0036 0.9987 0.9922

313.15 K

0.9748 0.9773 0.9790 0.9793 0.9806 0.9828 0.9850 0.9857 0.9863 0.9876 0.9898 0.9912 0.9916 0.9920 0.9908

1.0008 1.0026 1.0034 1.0042 1.0049 1.0056 1.0063 1.0071 1.0073 1.0078 1.0081 1.0083 1.0084 1.0083 1.0074 1.0062 1.0040 1.0009 0.9962 0.9908

318.15 K

Table 3. Densities (ρ/(g·cm−3)) of Water + 1,2-Ethanediol, 1,3-Propanediol, 1,4-Butanediol, and 1,5-Pentanediol Binary Mixtures at Different Temperatures (T/K) and Pressure of 0.1 MPaa

Journal of Chemical & Engineering Data Article

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. Dependence of excess molar volume (VEm) as a function of water composition (x1) for binary mixtures of water + 1,n-alkanediol in comparison with obtained data from literature at T = 298.15 K. The solid lines represent the corresponding correlations by the Redlich− Kister equation.

uncertainty for the values of refractive index in this work is 0.001. 2.2. Experimental Procedures. All measurements of mass were performed on an electronic single pan balance with a standard uncertainty of 0.1 mg (Mettler Toledo AG204). Doubly distilled water was used for the preparation of solutions and apparatus calibration. Densities of pure liquids and their mixtures were measured using an Anton Paar vibrating U-tube densitometer (model, DMA 500) with the standard uncertainty claimed by the manufacture of 0.001 g·cm−3 and temperature standard uncertainty of 0.1 K, calibrated at T = 293.15 K with ambient air and pure water. Dynamic viscosities, η, were obtained using an Anton Paar viscometer Lovis 2000 M rolling-ball automated viscometer with temperature standard uncertainty of 0.02 K. Different capillaries with different diameters (1.59, 1.8, and 2.5 mm) with a standard uncertainty claimed by the manufacture up to 0.5% were selected to allow measurement of viscosities from 0.7 to 1700 mPa·s. The calibration was carried out using pure water and different references standards of known viscosity corresponding to the core range of the capillaries, and the instrument constants were adjusted in a way that the known correct results are found by the instrument including the level adjustment, adjustment of capillary configuration, and temperature adjustment. Refractive indices nD measurements of pure components and mixtures were carried out using an Abbe refractometer 2WAJ coupled with transparent thermostat water bath (Julabo F34, Seelbach, Germany), which allows temperature stabilization with standard uncertainty of 0.01 K. The uncertainty claimed by the manufacture of refractive indices measurements was less than ±0.001 units. The reliability of refractive index measurements was checked by comparison between measurements of different pure liquids with data that were reported in the literature at different temperatures.

viscosity data reported for pure alkanediols shows that there is a notable scattering in comparison to the average values. Based on Table 2, the deviation of our data from literature varies from 0.2818 to 16.7919 for 1,3-propanediol at the temperature of 308.15 K. For our measurement the calculated relative standard uncertainty for viscosity is about 0.08. As can be seen, the agreement between the experimental and literature values was found to be satisfactory. Moreover, the comparison of the reported values of refractive index with those available in literature shows that the average absolute percentage deviation is less than 0.1, but the data reported by Tsierkezos and Molinou20 for 1,2-ethanediol at T = 298.15 K and by Nain21 for 1,4-butanediol at the temperatures of 298.15−318.15 K show an average absolute percentage deviation of more than 0.1. The relative standard

3. RESULTS AND DISCUSSION 3.1. Volumetric Properties. Experimental values of densities of water + 1,2-ethanediol (1,2-ED), 1,3-propanediol (1,3-PD), 1,4-butanediol (1,4-BD), and 1,5-pentanediol (1,5-PD) binary liquid systems under atmospheric pressure at temperatures of 288.15−318.15 K with 5 K interval over the whole composition range are reported in Table 3. Dependence of density to composition (as the function of water mole fraction) for the studied aqueous alkanediol systems at T = 298.15 K is shown in Figure 1. Accordingly, the isothermal density values of aqueous 1,2-ethanediol mixtures decreased nonlinearly with the increase of water concentration, whereas densities of aqueous 1,3-propanediol (1,3-PD), 1,4-butanediol (1,4-BD), and 1,5-pentanediol (1,5-PD) pass over a maximum

Figure 1. Dependence of density (ρ/(g·cm−3)) as a function of water composition (x1) for water + 1,n-alkanediol binary mixtures at T = 298.15 K.

Table 4. Coefficients (A0, A1, A2, and A3) of the Redlich−Kister Equation along with Their Standard Deviations (σ) for Excess Molar Volume (VEm) and Excess Refractive Index (nED) for Binary Mixtures of Water + 1,2-Ethanediol (1,2-ED), 1,3-Propanediol (1,3-PD), 1,4-Butanediol (1,4-BD), and 1,5-Pentanediol (1,5-PD) at T = 298.15 K excess molar volume 1,2-ED 1,3-PD 1,4-BD 1,5-PD

excess refractive index

A0

A1

A2

A3

σ

A0

A1

A2

A3

σ

−1.309 −1.604 −2.205 −2.167

−0.642 −0.456 −0.426 −0.573

−0.441 0.480 −0.609 −0.702

0.331 −0.035 0.231 0.536

0.0031 0.0014 0.0060 0.0021

0.0160 0.0084 0.0069 0.0113

0.0155 0.0143 0.0156 0.0056

0.0091 0.0059 0.0354 0.0056

−0.0152 −0.0111 −0.0041 −0.0073

0.0007 0.0001 0.0001 0.0001

E

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 3. Dependence of partial molar volume of alkanediol (V̅ mi/(cm3·mol−1)) as a function of water mole fraction (x1) for binary mixtures of water + alkanediols: (a) 1,2-ethanediol, (b) 1,3-propanediol, (c) 1,4-butanediol, and (d) 1,5-pentanediol at different temperatures.

in x1 = 0.3, x1 = 0.6, and x1 = 0.8, respectively. Indeed, with the increment of the distance between hydroxyl groups, the density variations show a maximum that shifts to a water-rich region in the order of water + 1,3-PD (0.3) < water + 1,4-BD (0.6) < water +1,5-PD (0.7). With considering ρid = x1ρ1 + x2ρ2 as the ideal density, the maximum deviation takes place at water mole fraction of more than 0.5 and shifts to the water-rich region along with a decrement in deviation from ideality when the size of the diols increases. Furthermore, the densities of the binary mixtures vary as follows: water + 1,2-ED > water + 1,3-PD > water + 1,4-BD > water + 1,5-PD. The packing effect and molecular weight determine the density variation in the mixtures. The study of excess molar volume and excess refractive index, discussed below, indicated that small water molecules can be better packed with larger diols; hence, it should be expected that the aqueous larger diols have greater density values as well as greater density deviation between mixtures. However, in this case, the mass effects are more dominant than packing effects. The excess molar volumes of water + diols systems were calculated from density measurements using the following equation: VmE =

xiMi + xjMj ρij

⎛ xjMj ⎞ xM ⎟ − ⎜⎜ i i + ρj ⎟⎠ ⎝ ρi

obtained results, it can be seen that the excess molar volumes are negative for all studied aqueous systems at all temperatures over the whole composition range. Close inspection of the excess molar volume curves shows that for all binary systems the minimum sets at water mole fraction of about 0.5−0.6, and these positions do not change with temperature. These results are in agreement with those available in the literature. Moreover, the maximum negative value varies in the order of water + 1,2-ED < water + 1,3-PD < water + 1,4-BD ≈ water + 1,5-PD. The magnitude of excess molar volume is the result of several effects including the constituent of strong H-bonding between water and diols (negative contribution), interstitial accommodation of unlike molecules due to difference in molar volume and free volume between components (mostly negative contribution), and dissociation of identical molecules on mixing with positive effect on excess volume. The difference in negative excess molar volume can be attributed to the structural effects because the functional groups of all the used alkanediols are similar. In other words, with increasing the carbon chain length from 1,2-ethanediol to 1,5-pentanediol, the water molecules meet less structural hindrance to embed in diols, and therefore, more numbers of water molecules can be embedded in alkanediol structures. Excess molar volume of all aqueous diols mixtures shifts to zero value when temperature increases. The increasing of VEm values with temperature rising can be attributed to the structural disorder and H-bond cleavage between unlike components. The negative excess molar enthalpy of water + 1,2-ethanediol, 1,3-propanediol, and 1,4-butanediol was obtained in the works of Matsumoto et al.43 and Nagamachi and Francesconi44 (Figure S4). The results show that the values of excess molar enthalpy are negative for all aqueous systems in the order of

(1)

where ρij is the measured density of the binary mixture and xi, Mi, and ρi are the mole fraction, molar mass, and density of component i, respectively. The obtained excess molar volume data are reported in Tables S1−S4 and shown in Figure 2 at T = 298.15 K and Figure S3 at all studied temperatures in comparison with the available data in the literature. From the F

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

a

G

1.4320 1.4278 1.4239 1.4202 1.4152 1.4080 1.3996 1.3881 1.3724 1.3338

1.4407 1.4361 1.4335 1.4290 1.4248 1.4190 1.4107 1.3985 1.3810 1.3338

0.0000 0.1593 0.2669 0.3496 0.4449 0.5563 0.6518 0.7482 0.8444 1.0000

0.0000 0.1737 0.2519 0.3669 0.4553 0.5501 0.6509 0.7532 0.8508 1.0000

1.4393 1.4347 1.4321 1.4276 1.4234 1.4177 1.4095 1.3973 1.3800 1.3331

1.4306 1.4263 1.4225 1.4188 1.4138 1.4067 1.3984 1.3870 1.3714 1.3331

293.15 K

308.15 K 1.4269 1.4225 1.4187 1.4150 1.4102 1.4031 1.3949 1.3838 1.3687 1.3312

1.4353 1.4306 1.4281 1.4236 1.4194 1.4139 1.4059 1.3939 1.3771 1.3312

303.15 K

Water + 1,2-Ethanediol 1.4294 1.4282 1.4251 1.4239 1.4212 1.4200 1.4176 1.4163 1.4126 1.4114 1.4055 1.4043 1.3973 1.3961 1.3860 1.3849 1.3705 1.3696 1.3325 1.3318

Water + 1,3-Propanediol 1.4380 1.4367 1.4334 1.4320 1.4308 1.4295 1.4263 1.4250 1.4221 1.4208 1.4165 1.4152 1.4083 1.4071 1.3962 1.3951 1.3790 1.3780 1.3325 1.3318

298.15 K

1.4340 1.4294 1.4268 1.4223 1.4181 1.4127 1.4047 1.3929 1.3761 1.3305

1.4257 1.4213 1.4174 1.4138 1.4090 1.4019 1.3938 1.3827 1.3677 1.3305

313.15 K

1.4327 1.4281 1.4255 1.4210 1.4169 1.4115 1.4036 1.3918 1.3752 1.3299

1.4244 1.4200 1.4161 1.4125 1.4077 1.4006 1.3926 1.3816 1.3668 1.3299

318.15 K

0.0000 0.1235 0.2586 0.3455 0.4864 0.5755 0.6777 0.7611 0.8474 0.9205 1.0000

0.0000 0.152 0.2513 0.3534 0.4482 0.5515 0.6527 0.7501 0.8492 0.9488 1.0000

x1

1.4503 1.4493 1.4477 1.4463 1.4425 1.4401 1.4342 1.4253 1.4085 1.3851 1.3338

1.4470 1.4441 1.4413 1.4376 1.4332 1.4277 1.4203 1.4097 1.3913 1.3596 1.3338

288.15 K

Standard uncertainties are u(T) = 0.01 K, u(p) = 10 kPa, and u(x1) = 0.0002; relative standard uncertainty is ur(nD) = 0.01

288.15 K

x1

1.4490 1.4477 1.4462 1.4447 1.4411 1.4386 1.4328 1.4239 1.4073 1.3840 1.3331

1.4455 1.4425 1.4397 1.4360 1.4317 1.4262 1.4188 1.4082 1.390 1.3586 1.3331

293.15 K

303.15 K

Water + 1,4-Butanediol 1.4441 1.4426 1.4411 1.4395 1.4383 1.4368 1.4346 1.4331 1.4304 1.4289 1.4247 1.4232 1.4174 1.4160 1.4069 1.4056 1.3888 1.3876 1.3577 1.3569 1.3325 1.3318 Water + 1,5-Pentanediol 1.4478 1.4466 1.4464 1.4448 1.4447 1.4431 1.4432 1.4416 1.4396 1.4380 1.4371 1.4355 1.4315 1.4300 1.4225 1.4211 1.4060 1.4048 1.3828 1.3816 1.3325 1.3318

298.15 K

1.4454 1.4434 1.4416 1.4402 1.4367 1.4341 1.4286 1.4197 1.4035 1.3804 1.3312

1.4412 1.4381 1.4353 1.4316 1.4274 1.4219 1.4147 1.4044 1.3865 1.3560 1.3312

308.15 K

1.4441 1.4419 1.4400 1.4386 1.4352 1.4326 1.4271 1.4183 1.4022 1.3793 1.3305

1.4398 1.4367 1.4339 1.4302 1.4260 1.4205 1.4134 1.4031 1.3853 1.3552 1.3305

313.15 K

1.4430 1.4408 1.4386 1.4373 1.4340 1.4313 1.4259 1.4171 1.4010 1.3781 1.3299

1.4384 1.4352 1.4324 1.4288 1.4246 1.4191 1.4120 1.4018 1.3842 1.3543 1.3299

318.15 K

Table 5. Experimental Values of Refractive Index (nD) for Water + 1,2-Ethanediol, 1,3-Propanediol, 1,4-Butanediol, and 1,5-Pentanediol Binary Mixtures at Different Temperatures (T/K) and Pressure of 0.1 MPaa

Journal of Chemical & Engineering Data Article

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

where V*1 and V*2 are the molar volumes of pure components 1 and 2, respectively. The

water + 1,2-ED < water + 1,3-PD < water + 1,4-BD with the negative maximum at the water mole fraction of about 0.6. Also, with increasing temperature, the negative excess molar enthalpy values shift to zero.29,45 Comparison between Figures 2 and S4 indicates a specific interaction between components in all binary mixtures. To drive adjustable parameters and estimate the data point deviation, excess molar volumes of binary mixtures at T = 298.15 K were fitted to an isothermal Redlich−Kister type equation.46 The Redlich−Kister coefficients, Ai, at each temperature were determined by the least-squares method and are given in Table 4 along with the standard deviations of the fits. The number of Ai parameters depends on a molecular complexity of solution, on the quality of the data, and on the number of data points available. The partial molar volume of component i in the binary ∂V mixtures is defined by the Vi̅ = ∂n equation, in which ni

( )

T , P , nj

ε−1 4π = Nα ε+2 3

is the number of moles of component i that added to the system and nj is the number of moles of the other component. The partial molar volumes of water, V̅ m,1, and diols, V̅ m,2, were calculated using the following equations: ⎛ ∂V E ⎞ Vm̅ ,1 = VmE + V1* + (1 − x1)⎜ m ⎟ ⎝ ∂x1 ⎠T , P

⎛ ∂V E ⎞ Vm̅ ,2 = VmE + V 2* − x1⎜ m ⎟ ⎝ ∂x1 ⎠T , P

value was calculated by the

T ,P

differentiation of the polynomial fitting of the VEm equation with respect to x1. The calculated values of the partial molar volumes are given in Tables S1−S4 and graphically presented in Figures 3 and S5. From Figures 3 and S5, it can be observed that with increasing the partial molar volume of one component in the mixture, the partial molar volume of the other component decreases, which is based on the Gibbs−Duhem equation. Moreover, Figure 3 indicates that these aqueous diol systems show the anomaly thermodynamic behavior at the water-rich region (water mole fraction of >0.9). This anomaly has taken place at the water-rich region of water + small amphiphilic molecules, such as C1−C4 monohydric alcohols, amines, and ethers. Dynamic light scattering, small-angle neutron scattering, and X-ray diffraction as well as molecular dynamic studies on the anomaly behavior of water and small amphiphilic solution indicated that the formation of clathrate hydrate structure (>nanometer in size) at this water-rich region is the main reason for the anomaly behavior of these aqueous mixtures.47,48 3.2. Optical Properties. Experimental refractive index data nD for water + 1,2-ED, 1,3-PD, 1,4-BD, and 1,5-PD binary mixtures were measured as a function of the water concentration at T = 288.15, 293.15, 298.15, 303.15, 308.15, 313.15, and 318.15 K (Table 5). Figure S6 and the reported data in Table 5 show that refractive indices data of the aqueous alkanediols systems decrease nonlinearly with rising water concentration and decrease as temperature increases. Furthermore, the refractive index values of pure diols and their aqueous mixtures decrease in the order of 1,5-PD > 1,4-BD > 1,3-PD > 1,2-ED. Our results confirm that the longer is the length of the alkyl chain, the higher is the refractive index, and the results are also in agreement with the obtained density values. For these studied pure diols, density and refractive index has the opposite order. If we consider the variation plot of density and refractive index of these pure materials against the carbon number of diols, it can be observed that the density decreases in a concave trend with the equation of y = 0.0084x2 − 0.0831x + 1.1843 and the refractive index increases in a convex trend with the equation of y = −0.0011x2 + 0.0119x + 1.4187 (Figure S7 in the Supporting Information). The volumetric and optical properties connect together as molar refraction, which is defined as the measure of the total polarizability of a mole of a substance, dependent on the temperature, pressure, density, and refractive index. In other words, the molar refractivity reflects the deformation of the electron shells of the atoms or molecules under the influence of the electric fields of neighboring ions. This thermodynamic property was initially calculated using the Clausius−Mossotti equation:

Figure 4. Dependence of (a) electronic polarizability (αe/(cm3·mol−1) and (b) molar refraction (Rm/mol−1) as a function of water mole fraction (x1) for binary mixtures of water + 1,n-alkanediol at T = 298.15 K.

i

∂VmE ∂x1

( )

(4)

where ε, N, and α are the dielectric constant permittivity, number of particles per cm, and polarizability, respectively. With the help of the Maxwell relation for the dielectric constant and the refractive index, ε = nD2, the molar refraction can be rewritten as follows:

(2)

Rm =

(3) H

nD 2 − 1 M 4π = NAαe 2 3 nD + 2 ρ

(5) DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

where αe denotes the part of the polarizability related with the electronic polarization. The plots of molar refraction and electronic polarizability for all binaries at T = 298.15 K are presented in Figure 4. Because the molar refraction is directly proportional to molecular polarizability, this quantity has behavior similar to electronic polarizability. Moreover, it can be seen that the molar refraction and the electronic polarizability of diol mixtures decrease linearly with increasing water concentration. In fact, when the structure of a molecule becomes less complicated , its electron cloud becomes less distributed and decentralized, and therefore, the polarizability of the molecule as well as its molar refraction decrease. The innovative approach to determine the excess refractive index of binaries was reported by Reis and co-workers49 as in the following equation: nDE = nD − [φ1(nD*1)2 + φ2(nD*2)2 ]1/2

values for the binary systems of water and alkanediols are positive over the entire concentration range and change to more positive values at higher temperatures. The positive contributions are a consequence of the less light speed for the studied binaries in comparison to their ideal solutions. This phenomenon arises from suitable interstitial accommodation giving more compact structure of mixtures and strong intermolecular interactions, attributed to the charge transfer and hydrogen bonding between unlike molecules. Furthermore, as opposed to the negative sign of excess molar volume, the excess refractive index has a positive sign. When the excess molar volume is negative, there is less available free volume than in the ideal solution and photons will be more likely to interact with the molecules or ions constituting the compound. Therefore, light will travel at a lower velocity in the medium concerned, and its refractive index will be higher than in an ideal solution. 3.3. Viscosities. The measured values of viscosity for all systems at atmospheric pressure and temperatures of 293.15, 298.15, and 303.15 K are reported in Table 6. Also, Figure 6 shows the dependence of viscosity as a function of water mole fraction for the studied systems in this work and in the literature. The viscosities of aqueous 1,2-ethanediol with the measurements of Tsierkezos and Molinou20 and Yang et al.34 and also the viscosities of 1,3-propanediol mixtures with those reported by Maximino50 and George and Sastry19 indicate similar

(6)

Although the refractive index is an optical property, they found a reasonable method between the concept of optical properties and excess thermodynamic quantities using the relationships between refractive index and density as well as Maxwell’s equations of electromagnetism. The calculated values of excess refractive index are tabulated in Tables S1−S4. Figure 5 shows the dependence of excess refractive index as a function of water mole fraction for the studied binaries at different temperatures. Accordingly, it can be observed that the excess refractive index

Figure 5. Dependence of excess refractive index (nED) as a function of water mole fraction (x1) for binary mixtures of water + alkanediol: (a) 1,2-ethanediol, (b) 1,3-propanediol, (c) 1,4-butanediol, and (d) 1,5-pentanediol at different temperatures. I

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 6. Experimental Values of Viscosity (η/(mPa·s) for Water + 1,2-Ethanediol, 1,3-Propanediol, 1,4-Butanediol, and 1,5-Pentanediol Binary Mixtures at Different Temperatures and Pressure of 0.1 MPaa x1 0.0000 0.1105 0.2014 0.3002 0.4011 0.4969 0.6022 0.6995 0.7994 0.8997 1.0000 0.0000 0.0749 0.1625 0.239 0.3361 0.4223 0.5134 0.6309 0.7446 0.8720 1.0000 a

T = 293.15 K

T = 298.15 K

Water + 1,2-Ethanediol 21.13 17.24 17.58 14.33 14.95 12.19 12.40 10.18 10.05 8.310 8.185 6.791 6.297 5.208 4.584 3.884 3.122 2.682 1.918 1.664 0.949 0.843 Water + 1,3-Propanediol 51.55 40.8 46.34 36.93 40.76 32.58 35.86 29.01 29.72 24.27 24.46 20.06 19.25 15.64 13.16 10.59 7.990 6.430 3.720 3.130 0.949 0.843

T = 303.15 K

x1

14.23 11.86 10.14 8.450 6.968 5.653 4.396 3.345 2.328 1.446 0.755

0.0000 0.1257 0.2783 0.3033 0.4009 0.5008 0.6662 0.7003 0.8013 0.8979 1.0000

32.74 29.67 26.38 23.45 19.97 16.66 13.18 8.95 5.460 2.780 0.755

0.0000 0.2586 0.3000 0.4864 0.4969 0.6777 0.7611 0.8474 0.9205 1.0000

T = 293.15 K

T = 298.15 K

Water + 1,4-Butanediol 90.7 69.89 75.28 57.95 58.11 44.52 54.65 42.00 44.04 33.94 33.85 25.34 19.58 13.90 16.598 11.57 9.634 6.814 4.127 3.267 0.949 0.843 Water + 1,5-Pentanediol 118.7 86.79 92.52 70.36 85.43 65.59 51.43 39.51 49.66 38.56 25.01 19.75 16.10 12.71 8.205 6.627 3.733 3.103 0.949 0.843

T = 303.15 K 55.24 45.05 33.59 31.63 25.07 18.65 10.11 8.774 5.035 2.369 0.755 68.3 54.61 51.03 31.00 30.24 15.78 10.16 5.439 2.615 0.755

Standard uncertainties are u(T) = 0.01 K, u(p) = 10 kPa, and u(x1) = 0.0002; relative standard uncertainty is ur(μ) = 0.08.

Figure 6. Dependence of viscosity (η/(mPa·s)) as a function of water mole fraction (x1) for binary mixtures of water + alkandiol: (a) 1,2-ethanediol, (b) 1,3-propanediol, (c) 1,4-butanediol, and (d) 1,5-pentanediol at different temperatures (T).

behavior against concentration and temperature. According to Figure 6, the viscosity of alkanediol mixtures decreases with

a third order polynomial equation with increasing water concentration. Moreover, the magnitude of the viscosity values J

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 7. Adjustable Parameters of Grunberg−Nissan (G12) and McAllister Equations (A12 and A21) along with Their Absolute Average Deviation Percent at Different Temperatures and Pressure of 0.1 MPa Grunberg−Nissan

T = 293.15 K T = 298.15 K T = 303.15 K

water + 1,4-butanediol

G12

G12

8.33 4.29 8.12 4.42 7.68 3.80 Grunberg−Nissan

AAD%

2.36 2.27 1.89

water + 1,3-propanediol T = 293.15 K T = 298.15 K T = 303.15 K

new Grunberg−Nissan A

water + 1,2-ethanediol

AAD%

AAD%

20.22 16.74 14.32

10.48 10.88 11.22

16.36 11.50 4.53 new Grunberg−Nissan A

water + 1,3-propanediol

water + 1,5-pentanediol AAD%

G12

AAD%

G12

AAD%

AAD%

3.86 3.79 3.73

12.41 11.29 11.43

5.84 5.95 5.35 McAllister

20.87 18.84 19.02

7.30 7.94 7.60

A12

A21

2.19 2.23 1.76

2.40 1.79 2.00

26.76 26.55 24.50 new Grunberg−Nissan B

water + 1,4-butanediol

water + 1,2-ethanediol

water + 1,4-butanediol

A12

A21

ADD%

ADD%

ADD%

2.87 4.62 3.80 4.15 2.99 3.79 McAllister

3.68 3.44 3.19

7.32 5.42 5.42

ADD%

water + 1,3-propanediol T = 293.15 K T = 298.15 K T = 303.15 K

water + 1,4-butanediol

AAD%

water + 1,5-pentanediol

water + 1,2-ethanediol T = 293.15 K T = 298.15 K T = 303.15 K

water + 1,2-ethanediol

15.71 11.67 16.20 6.74 16.22 2.81 new Grunberg−Nissan B

water + 1,5-pentanediol

water + 1,3-propanediol

water + 1,5-pentanediol

A12

A21

ADD%

A12

A21

ADD%

ADD%

ADD%

3.46 3.14 3.09

3.42 3.20 3.11

4.35 3.83 5.28

5.34 4.92 3.79

3.92 4.00 3.19

9.58 8.92 10.92

2.99 2.86 2.53

23.33 23.10 20.90

varies in the order of water + 1,2-ED < water + 1,3-PD < water + 1,4-BD < water + 1,5-PD. The slope of the viscosity plot against concentration strongly depends on the molecular weight of the used components in the mixtures as well as their structures and functional groups. Regarding the influence of temperature on viscosity, a decrement in viscosity is observed when temperature increases, which is the routine behavior in liquids. Temperature increase causes agitations in the fluid, and the bonding among the molecules begins to break; hence, the resistance of fluid to flow decreases. Based on the dependence of viscosity to temperature and composition, numerous models have been proposed to predict the fluid viscosities.51 Herein, Grunberg−Nissan52 and McAllister53 models were utilized to predict the viscosity values of water + 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol solutions. Table 7 represents the obtained values of the adjustable parameters of Grunberg−Nissan and McAllister equations along with the average absolute percentage deviation. According to the results, it can be seen that the McAllister model with more adjustable parameters than Grunberg−Nissan shows a better prediction for viscosity of aqueous diols. For both models, the deviations increase with increasing carbon chain length of alkanediols. In order to substitute the constant parameter in the Grunberg−Nissan model with the available variables in the equation, the different expressions were examined and finally the following equations for aqueous alkanediols systems were obtained with the least deviations:

new proposed Grunberg−Nissan B: ⎛ ⎞ η x2 ln η = x1 ln η1 + x 2 ln η2 + x1x 2⎜⎜ ln 2 + |x 2 − x1|⎟⎟ η1x1 ⎝ ⎠ (8)

where, x1, η1 and x2, η2 are mole fraction and viscosity of components 1 and 2, respectively. These new Grunberg− Nissan models were examined for water + 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol binary mixtures and represent the acceptable results for all mixtures at all temperatures. The AAD % based on these new models are reported in Table 7. Both models show the least average absolute percentage deviation for water + 1,3-propanediol binaries.

4. CONCLUSIONS Densities, viscosities, and refractive indices of water + 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol binary mixtures were measured at atmospheric pressure and different temperatures between T = 288.15 and 318.15 K over the whole composition range. From the experimental data, it was found that by increasing the carbon chain length of alkanediol, the density of pure alkanediol decreases and refractive index and viscosity increase, indicating the dominance of structural hindrance to H-bonding strength between molecules with the increase of carbon number of the diols. Moreover, the negative excess molar volume and positive excess refractive index were obtained which indicated the presence of specific molecular interactions between unlike components. Furthermore, the negative excess molar volumes and positive excess refractive indices shifted to zero when temperature increased. The partial molar volumes of alkanediols decreased

new proposed Grunberg−Nissan A: ⎛ η x2 ln η = x1 ln η1 + x 2 ln η2 + x1x 2⎜⎜ ln 2 η1x1 ⎝

⎞ ⎟⎟ ⎠

(7) K

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Coefficient of Semidiluted Aqueous and Nonaqueous Polymer Solutions in Vapor−Liquid Equilibrium. J. Chem. Eng. Data 2015, 60, 2701−2708. (8) Farzi, N.; Hosseini, P. A new equation of state for gaseous, liquid, and supercritical fluids. Fluid Phase Equilib. 2016, 409, 59−71. (9) Checoni, R. F. Experimental study of the excess molar volume of ternary mixtures containing {water + (1,2-propanediol, or 1,3propanediol, or 1,2-butanediol, or 1,3-butanediol, or 1,4-butanediol, or 2,3-butanediol) + electrolytes} at a temperature of 298.15 K and atmospheric pressure. J. Chem. Thermodyn. 2010, 42, 612−620. (10) Li, Q. S.; Tian, Y. M.; Wang, S. J. Densities and Excess Molar Volumes for Binary Mixtures of 1,4-Butanediol + 1,2-Propanediol, + 1,3-Propanediol, and + Ethane-1,2-diol from (293.15 to 328.15) K. J. Chem. Eng. Data 2008, 53, 271−274. (11) Atilhan, M.; Aparicio, S. PqT measurements and derived properties of liquid 1,2-alkanediols. J. Chem. Thermodyn. 2013, 57, 137−144. (12) Zorebski, E.; Dzida, M. The effect of temperature and pressure on acoustic and thermodynamic properties of 1,4-butanediol. The comparison with 1,2-, and 1,3-butanediols. J. Chem. Thermodyn. 2012, 54, 100−107. (13) Góralski, P.; Tkaczyk, M. Heat Capacities of Some Liquid α,ωAlkanediols within the Temperature Range between (293.15 and 353.15) K. J. Chem. Eng. Data 2008, 53, 1932−1934. (14) Zemánková, K.; Troncoso, J.; Romaní, L. Excess volumes and excess heat capacities for alkanediol + watersystems in the temperature interval (283.15−313.15)K. Fluid Phase Equilib. 2013, 356, 1−10. (15) Domańska, U.; Papis, P.; Szydłowski, J.; Królikowska, M.; Królikowski, M. Excess Enthalpies of Mixing, Effect of Temperature and Composition on the Density, and Viscosity and Thermodynamic Properties of Binary Systems of {Ammonium-Based Ionic Liquid + Alkanediol}. J. Phys. Chem. B 2014, 118, 12692−12705. (16) Kestin, J.; Sokolov, M.; Wakeham, W. A. Viscosity of liquid water in the range −8 °C to 150 °C. J. Phys. Chem. Ref. Data 1978, 7 (3). (17) Bajić, D. M.; Ž ivković, E. M.; Šerbanović, S. P.; Kijevčanin, M. L. Experimental measurements and modelling of volumetric properties,refractive index and viscosity of selected binary systems with butyllactate at 288.15−323.15 K and atmospheric pressure.New UNIFAC-VISCO interaction parameters. Thermochim. Acta 2013, 562, 42−55. (18) Cocchi, M.; Marchetti, A.; Sanna, G.; Tassi, L.; Ulrici, A.; Vaccari, G. Kinematic viscosities of ternary mixtures containing ethane-1,2-diol, 2-methoxyethanol and water from −100C to 800C. Fluid Phase Equilib. 1999, 157, 317−342. (19) George, J.; Sastry, N. V. Densities, Dynamic Viscosities, Speeds of Sound, and Relative Permittivities for Water + Alkanediols (Propane-1,2- and −1,3-diol and Butane-1,2-, −1,3-, −1,4-, and −2,3-Diol) at Different Temperatures. J. Chem. Eng. Data 2003, 48, 1529−1539. (20) Tsierkezos, N. G.; Molinou, I. E. Thermodynamic Properties of Water + Ethylene Glycol at 283.15, 293.15, 303.15, and 313.15 K. J. Chem. Eng. Data 1998, 43, 989−993. (21) Nain, A. K. Refractive Indices and Deviations in Refractive Indices for Binary Mixtures of Formamide + 1-Butanol, + 2-Butanol, + 1,3-Butanediol, and + 1,4-Butanediol at Temperatures from (293.15 to 318.15) K. J. Chem. Eng. Data 2008, 53, 1208−1210. (22) Hnědkovský, L.; Cibulka, I. J. Partial Molar Isentropic Compressions and Partial Molar Volumes of Isomeric Butanediols at Infinite Dilution in Water at Temperatures T = (278 to 318) K and Atmospheric Pressure. J. Chem. Eng. Data 2013, 58, 388−397. (23) Schiebener, P.; Straub, J.; Levelt Sengers, J. M. H.; Gallagher, J. S. Refractive Index of Water and Steam as Function of Wavelength, Temperature and Density. J. Phys. Chem. Ref. Data 1990, 19, 677−717. (24) Thormahlen, I.; Straub, J.; Grigull, U. Refractive Index of Water and Its Dependence on Wavelenght, Temperature and Density. J. Phys. Chem. Ref. Data 1985, 14, 933−945. (25) Kijevčanin, M. L.; Ž ivković, E. M.; Djordjević, B. D.; Radović, I. R.; Jovanović, J.; Šerbanović, S. P. Experimental determination and

with the increase of water mole fraction and represented the minimum (anomaly behavior) at the water-rich region. Using experimental measurements of refractive index, the molar refraction and electronic polarizability of aqueous diols systems were calculated which decrease linearly with increasing water concentration. Finally, the viscosity of aqueous 1,n-alkanediol systems decreased with increasing water concentration and temperature.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.6b00526. Excess molar volume, partial molar volume and excess refractive index for the binary mixtures of water + 1,2-ethanediol, 1,3-propanediol, 1,4-butanediol, and 1,5-pentanediol at different temperatures, gas chromatography and FTIR spectra of pure alkanediols, and dependence of excess molar volume, partial molar volume, refractive index as a function of water mole fraction for the binary mixtures of water + 1,n-alkanediols (n= 2−5) (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel.: (+98) 1135302384; Fax: (+98) 1135302302. E-mail: [email protected]. ORCID

Abbas Ali Rostami: 0000-0002-3180-658X Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS The instrumental support from Shomal University is gratefully acknowledged. REFERENCES

(1) Llovell, F.; Marcos, R. M.; Vega, L. F. Free-Volume Theory Coupled with Soft-SAFT for Viscosity Calculations: Comparison with Molecular Simulation and Experimental Data. J. Phys. Chem. B 2013, 117, 8159−8171. (2) Su, Y.-C.; Liu, Y. A.; Diaz Tovar, C. A.; Gani, R. Selection of Prediction Methods for Thermophysical Properties for Process Modeling and Product Design of Biodiesel Manufacturing. Ind. Eng. Chem. Res. 2011, 50, 6809−6836. (3) Hendriks, E.; Kontogeorgis, G. M.; Dohrn, R.; De Hemptinne, J. C.; Economou, I. G.; Ẑ ilnik, L. F.; Vesovic, V. Industrial Requirements for Thermodynamics and Transport Properties. Ind. Eng. Chem. Res. 2010, 49, 11131−11141. (4) Anitescu, G.; Bruno, T. J.; Tavlarides, L. L. Dieseline for Supercritical Injection and Combustion in Compression-Ignition Engines: Volatility, Phase Transitions, Spray/Jet Structure, and Thermal Stability. Energy Fuels 2012, 26, 6247−6258. (5) Burgess, W. A.; Tapriyal, D.; Gamwo, I. K.; Morreale, B. D.; McHugh, M. A.; Enick, R. M. Viscosity Models Based on the Free Volume and Frictional Theories for Systems at Pressures to 276 MPa and Temperatures to 533 K. Ind. Eng. Chem. Res. 2012, 51, 16721− 16733. (6) Mazloumi, S. H. Representation of activity and osmotic coefficients of electrolyte solutions using non-electrolyte WilsonNRF model with ion-specific parameters. Fluid Phase Equilib. 2015, 388, 31−36. (7) Sadeghi, R.; Shahebrahimi, Y.; Zonnouri, A. The Chemical Thermodynamic Models for Calculating the Solvent Activity L

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(44) Nagamachi, M. Y.; Francesconi, A. Z. Measurement and correlation of excess molar enthalpy HmE for (1,2-propanediol, or 1,3propanediol, or 1,4-butanediol + water) at the temperatures (298.15, 323.15, and 343.15) K. J. Chem. Thermodyn. 2006, 38, 461−466. (45) Kracht, C.; Ulbig, P.; Schulz, S. Measurement and correlation of excess molar enthalpies for (ethanediol, or 1,2-propanediol, or1,2butanediol + water) at the temperatures (285.65, 298.15, 308.15, 323.15, and 338.15) K. J. Chem. Thermodyn. 1999, 31, 1113−1127. (46) Redlich, O.; Kister, T. ALGEBRAIC REPRESENTATION OF THERMODYNAMIC PROPERTIES AND THE CLASSIFICATION OF SOLUTIONS. Ind. Eng. Chem. 1948, 40, 345−348. (47) Subramanian, D.; Boughter, C. T.; Klauda, J. B.; Hammouda, B.; Anisimov, M. a. Mesoscale inhomogeneities in aqueous solutions of small amphiphilic molecules. Faraday Discuss. 2014, 167, 217−238. (48) Subramanian, D.; Anisimov, M. A. Phase behavior and mesoscale solubilization in aqueous solutions of hydrotropes. Fluid Phase Equilib. 2014, 362, 170−176. (49) Reis, J. C. R.; Lampreia, I. M. S.; Santos, Â . F. S.; Moita, M. L. C. J.; Douhéret, G. Refractive Index of Liquid Mixtures: Theory and Experiment. ChemPhysChem 2010, 11, 3722−3733. (50) Belda Maximino, R. Viscosity and density of binary mixtures of alcohols and polyols with three carbon atoms and water: equation for the correlation of viscosities of binary mixtures. Phys. Chem. Liq. 2009, 47, 515−529. (51) Moosavi, M.; Omrani, A.; Ali Rostami, A.; Motahari, A. Isobaric, isothermal theoretical investigation and examination of different prediction equations on some physicochemical properties in PEG liquid polymer system. J. Chem. Thermodyn. 2014, 68, 205−215. (52) Grunberg, L.; Nissan, A. H. Mixture Low for Viscosity. Nature 1949, 164 (4175), 799−800. (53) McAllister, R. A. The Viscosity of Liq- uid Mixtures. AIChE J. 1960, 6, 427−431.

modeling of excess molar volumes, viscosities and refractive indices of the binary systems (pyridine + 1-propanol, + 1,2-propanediol, + 1,3propanediol, and + glycerol). New UNIFAC-VISCO parameters determination. J. Chem. Thermodyn. 2013, 56, 49−56. (26) Nain, A. K. Densities and Volumetric Properties of Binary Mixtures of Formamide with 1-Butanol, 2-Butanol, 1,3-Butanediol and 1,4-Butanediol at Temperatures between 293.15 and 318.15 K. Int. J. Thermophys. 2007, 28, 1228−1244. (27) Czechowski, G.; Zywucki, B.; Jadzyn, J. Excess Molar Volumes of Binary Mixtures of Diols and Water. J. Chem. Eng. Data 1988, 33, 55−57. (28) Zorębski, E.; Lubowiecka-Kostka, B. Thermodynamic and transport properties of (1,2-ethanediol + 1-nonanol) at temperatures from (298.15 to 313.15) K. J. Chem. Thermodyn. 2009, 41, 197−204. (29) Egorov, G. I.; Makarov, D. M.; Kolker, A. M. Liquid phase PVTx properties of binary mixtures of (water + ethylene glycol) in the range from 278.15 to 323.15 K and from 0.1 to 100 MPa. I. Experimental results, partial and excess thermodynamics properties. Fluid Phase Equilib. 2013, 344, 125−138. (30) Lee, J. W.; Park, S. B.; Lee, H. Densities, Surface Tensions, and Refractive Indices of the Water + 1,3-Propanediol System. J. Chem. Eng. Data 2000, 45, 166−168. (31) Piekarski, H.; Pietrzak, A. Molecular interactions of alkanediols in methanol and in water: density and heat capacity measurements. J. Mol. Liq. 2005, 121, 46−52. (32) Hawrylak, B.; Gracie, K.; Palepu, R. Thermodynamic Properties of Binary Mixtures of Butanediols with Water. J. Solution Chem. 1998, 27, 17−31. (33) Mehta, S. K.; Ram, G.; Kumar, V.; Bhasin, K. K. Structural and interactional studies of homologous series of a,w-alkanediols in N,Ndimethylformamide. J. Chem. Thermodyn. 2007, 39, 781−790. (34) Yang, C.; Ma, P.; Jing, F.; Tang, D. Excess Molar Volumes, Viscosities, and Heat Capacities for the Mixtures of Ethylene Glycol + Water from 273.15 to 353.15 K. J. Chem. Eng. Data 2003, 48, 836− 840. (35) Yang, C.; Ma, P.; Zhou, Q. Excess Molar Volume, Viscosity, and Heat Capacity for the Mixtures of 1,4-Butanediol + Water at Different Temperatures. J. Chem. Eng. Data 2004, 49, 582−587. (36) Nain, A. K. Molecular interactions in binary mixtures of formamide with 1-butanol, 2-butanol, 1,3-butanediol and 1,4butanediol at different temperatures: An ultrasonic and viscometric study. Fluid Phase Equilib. 2008, 265, 46−56. (37) Moosavi, M.; Motahari, A.; Omrani, A.; Rostami, A. A. Investigation on some thermophysical properties of poly(ethylene glycol) binary mixtures at different temperatures. J. Chem. Thermodyn. 2013, 58, 340−350. (38) Fontao, M. J.; Iglesias, M. Effect of Temperature on the Refractive Index of Aliphatic Hydroxilic Mixtures (C2−C3). Int. J. Thermophys. 2002, 23, 513−527. (39) Jiménez, E.; Cabanas, M.; Segade, L.; García-Garabal, S.; Casas, H. Excess volume, changes of refractive index and surface tension of binary 1,2-ethanediol + 1-propanol or 1-butanol mixtures at several temperatures. Fluid Phase Equilib. 2001, 180, 151−164. (40) Aminabhavi, T. M.; Banerjee, K. Density, Viscosity, Refractive Index, and Speed of Sound in Binary Mixtures of Methyl Acetate + Ethylene Glycol or + Poly(ethylene glycol) in the Temperature Interval (298.15−308.15) K. J. Chem. Eng. Data 1998, 43, 852−855. (41) Ciocirlan, O.; Croitoru, O.; Iulian, O. Density and Refractive Index of Binary Mixtures of Two 1-Alkyl-3-methylimidazolium Ionic Liquids with 1,4-Dioxane and Ethylene Glycol. J. Chem. Eng. Data 2014, 59, 1165−1174. (42) Ghanadzadeh Gilani, A.; Ghanadzadeh Gilani, H.; Ansari, M. Ansari, A thermodynamic study of solute−solvent interactions through dielectric properties of the mixtures consisting of 1,4-butanediol, 1octanol, and 1,4-dioxane at different temperatures. J. Chem. Thermodyn. 2012, 55, 203−212. (43) Matsumoto, Y.; Touhara, H.; Nakanishi, K.; Watanabe, N. Molar excess enthalpies for water + ethanediol, + 1,2-propanediol, and + 1,3-propanediol at 298.15 K. J. Chem. Thermodyn. 1977, 9, 801−805. M

DOI: 10.1021/acs.jced.6b00526 J. Chem. Eng. Data XXXX, XXX, XXX−XXX