Sucrose Solubility in Binary Liquid Mixtures Formed by Water

Jul 29, 2016 - ABSTRACT: This work presents a study of solid−liquid equilibrium of sucrose in three different binary liquid mixtures, methanol−eth...
3 downloads 0 Views 409KB Size
Article pubs.acs.org/jced

Sucrose Solubility in Binary Liquid Mixtures Formed by Water− Methanol, Water−Ethanol, and Methanol−Ethanol at 303 and 313 K Alessandro C. Galvaõ ,* Weber S. Robazza, Geiza N. Sarturi, Fernanda C. Goulart, and Daiane Conte Department of Food and Chemical Engineering − DEAQ, Laboratory ApTher − Applied Thermophysics, Santa Catarina State University − UDESC, SC 160 − km 68, Pinhalzinho, Santa Catarina, Brazil, 89870-000 ABSTRACT: This work presents a study of solid−liquid equilibrium of sucrose in three different binary liquid mixtures, methanol−ethanol, methanol−water, and ethanol−water for the temperatures of 303 and 313 K covering the whole composition range of the binary liquid solution. The data are presented as a function of the dielectric constant of the binary liquid mixture taking into account the self-association of water and alcohol. From a dissolution phenomenon standpoint, it seems that the temperature is more important than the dielectric constant of the binary liquid solution for the solubility mechanism. For each temperature, it was observed that the sucrose solubility is directly proportional to the dielectric constant of the binary liquid mixture. and water at different temperatures. Bouchard et al.8 measured the solubility of sucrose in water−ethanol solutions at 310 K. Gong et al.9 published data of solid−liquid equilibrium involving sucrose and water−ethanol mixtures from 273 to 293 K and water concentration up to 40 wt %. Van Putten et al.10 investigated the solubility of sucrose in methanol−water mixtures containing up to 25 wt % water from 295 to 353 K. The authors observed an increase of solubility with the increase of temperature and a decrease of solubility with the addition of an antisolvent to water. Generally, the solubility phenomenon follows a very simple rule, like dissolves like. The statement indicates that a solvent solubilizes a solute if both components have affinity with each other and this affinity depends mainly on the polarity of the substances. The polarity of a solvent or a mixture is directly proportional to its dielectric constant (ε) and that information is very useful to understand solubility. Usually the solubility of a solid in a mixture has been represented as a function of the solution mole fraction in solid free basis. From a theoretical standpoint, it would be more informative to represent the solubility distribution curve as a function of the dielectric constant of a mixture. This work presents the solubility of sucrose in three different binary mixtures, methanol−ethanol, methanol−water, and ethanol−water, for the temperatures of 303.2 and 313.2 K under atmospheric pressure and over the whole composition range of the binary mixture. The dielectric constant of the performed mixture was calculated taking into account the existence of self-association of water and alcohol and that information was used to analyze the data.

1. INTRODUCTION The development of technologies that yield the production of chemicals and fuel from a renewable source of carbon still depends on many different studies. For the use of biomass to accomplish this task, reliable data is necessary to raise information on how each component present in the biomass interacts with a liquid. Among many compounds present in the biomass, sucrose is a very important substance, widely used by the food industry and as a source of carbon; it may represent a sustainable source for the production of chemical products.1,2 The production of chemicals and energy from biomass is the goal of an industrial complex called biorefinery.3,4 Among many processes that can be present in such chemical plant, the treatment of biomass, conversion of biopolymers, fermentation reaction, and separation techniques require the use of solvents. The development and design of effective processes, such as separation units, still depend on many studies and many require the knowledge about how the compounds interact with pure solvents and especially with a mixture of solvents. The studies of solid−liquid equilibrium raise fundamental information that may be useful for the development and optimization of processes involving sugars and related molecules. Another motivation for the study deals with fundamental research; studies of solid−liquid solubility provide important information for understand the dissolution phenomenon and help the development and test of models and theories. Despite the usefulness of investigations involving solid− liquid equilibrium, there are relatively few works reporting solubility of sucrose in liquid mixtures. Segur and Miner5 studied the solubility of sucrose in aqueous glycerol mixtures. Peres and Macedo6 measured the solubility of sucrose in binary liquid mixtures formed by water, methanol, and ethanol for three different temperatures. Tsavas et al.7 investigated the solubility of sucrose in binary and ternary mixtures of alcohol, acid, ester, © XXXX American Chemical Society

Received: December 28, 2015 Accepted: July 18, 2016

A

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

2. EXPERIMENTAL SECTION The determination of sucrose solubility in the studied mixtures was performed by gravimetric method. In conducting the experiments, water, methanol P.A. and ethanol P.A. were used. None of the liquid reagents used in this work received further purification. Prior the study, the sucrose was dried in an electrical furnace for 24 h under the temperature of 353 K and kept in a desiccator until the beginning of the experiment. The source and purity of the materials used are presented in Table 1.

the vessel was used. The monitoring of the temperature of the experiment was taken by a bulb thermometer (uncertainty ±0.5 K) also adapted on top of the equilibrium cell. After feeding the equilibrium cell with the reagents, connecting the cell with the thermostatic bath, adjusting the stirring speed, and adjusting the thermostatic bath temperature, the circulation system of the heating fluid was activated. During the first hour of the experiment, the temperature of the bath was adjusted to ensure the accuracy of the temperature. With the solution inside the equilibrium cell in the test temperature, the system remained under strong agitation for 4 h and after that the system was kept without stirring for 24 h to ensure the solid− liquid phase separation. Samples of known weight were collected and evaporated under the temperature of 353 K until the entire removal of solvent. After sampling (performed in triplicate), a new amount of solid and liquid was placed in the equilibrium cell to complete the volume, a new temperature was set, and the experiment started again. The solubility data is expressed in mass fraction; the experimental uncertainty was estimated by a propagation method.17 The experimental results were evaluated as a function of the mixture dielectric constant (ε) as represented by eq 1 wherein ε1 and ε2 are the dielectric constant of each pure component of the binary mixture and Φ2 is the volume fraction of the dispersed component.18

Table 1. Source and Purity of the Chemicals Used in This Work component (i)

source

mass fraction

water methanol ethanol sucrose

our lab Biotec - Brazil Biotec - Brazil Biotec - Brazil

0.999 0.995 0.990

analysis method

purification method

refractive index doubly distilled refractive index none refractive index none none drying

In order to check the purity of the liquid reagents, analyses of refractive index were performed at 298.15 K and the results were compared with those available in the literature as shown in Table 2. To perform refractive index measurements a digital Table 2. Comparison of Experimental Refractive Index Lit (nExp D ) and Literature Values (nD ) at 298.15 K and 0.1 MPa component (i)

nExp D

nLit D

water methanol ethanol

1.33251 1.32655 1.35947

1.33255a, 1.33248b 1.32629a, 1.32645c, 1.32661e 1.35922b, 1.3591d, 1.35972f

ε = [(ε21/3 − ε11/3)Φ2 + ε11/3]3

(1)

In order to take into account the self-association of the components used in this work, the volume fraction was calculated following eq 2 as a function of mole fraction xi and characteristic volume of each pure component i.19

a

Lin et al.11 bUrréjola et al.12 cOrge et al.13 dMutalik et al.14 Papanastaslou and Ziogas15 fPapanastaslou et al.16 Standard uncertainties u are u(T) = 0.03 K, u(p) = 0.01 MPa, u(nD) = 2 × 10−5. e

Φ2 =

refractometer (Atago, Model RX-5000i, accuracy ±0.00004) was used. The analyses results showed that the liquid reagents acquired had purity enough for the purpose of the work. The liquid mixtures were prepared by using a balance (uncertainty ±0.001 g) and placed in a volumetric flask to facilitate homogenization and to wait until the start of the experiment. The binary solution and the sucrose (an excess of solid was used) were then transferred to a glass vessel, commonly known as equilibrium cell (working volume of approximately 125 mL), jacketed with input and output for connection to a thermostatic bath (uncertainty ±0.1 K). As a means to ensure the contact of liquid solution and solid reagent, a magnetic stirrer was used and for the removal of samples a syringe connected to a glass pipe fitted to the lid of

x 2V 2* * x1V1 + x 2V 2*

(2)

The characteristic volume (Vi*) is represented by eq 3 wherein vi is the molar volume of the components, αi is the thermal expansion coefficient and α*i represents the correction of the thermal expansion coefficient of each pure component as a function of temperature T. ⎡ ⎤3 * 1 + ( ∝ − ∝ ) T i i ⎥ V i* = vi⎢ ⎢⎣ 1 + 4 ( ∝i − ∝*i )T ⎥⎦ 3

(3)

The correction of the thermal expansion coefficient is necessary due to the effects of self-association of water, methanol and ethanol. The calculus is represented by eq 4 in which Δvi* is the molar volume of association, Δhi* is the molar

Table 3. Thermal Expansion Coefficient (α), Molar Enthalpy of Association (Δh*), Dielectric Constant (ε), Molar Volume of Association (Δv*) and Molar Volume (v) component (i) water methanol ethanol

T/K 303.2 313.2 303.2 313.2 303.2 313.2

αi × 104/K‑1 b

3.03 3.85b 11.95e 12.09e 11.13f 11.39f

Δh*i /kJ·mol‑1

Ki

−24.50 −21.50h −25.10a −25.10a −25.10a −25.10a

i

h

1959.14 1522.08i 834.35i 607.08i 268.25i 195.18i

εi c

76.73 73.12c 30.68c 29.03c 23.55c 22.20c

Δv*i /cm3·mol‑1

vi/cm3·mol‑1

−0.2 −0.2b −5.6a −5.6a −5.6a −5.6a

18.09g 18.16g 41.00d 41.49d 59.03f 59.68f

b

a

Funke et al.21 bScharlin et al.22 cÅkerlöf.23 dBender and Van Hook.24 eJain et al.25 fCorrelated from Ivanov.26 gCalculated from density measurement: Perry and Green.27 hInterpolated and extrapolated from Scharlin et al.22 iCalculated by eq 5. B

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

enthalpy of association, Ki represents the equilibrium constant of association and R is the universal gas constant.

∝*i = Δvi*Δhi*

(4K i + 1)1/2 − 2K i(4K i + 1)−1/2 − 1 2K iV i*RT 2 (4)

The dependence of the equilibrium constant of association with temperature is given by the Vant’Hoff equation in which K(T0) represents equilibrium constant of association at a reference temperature of 298.15 K. It was used equilibrium constant values of 2306,20 986,21 and 31721 for water, methanol, and ethanol, respectively. ⎡ Δh*(T ) ⎛ 1 1 ⎞⎤ K(T ) = K(T0) exp⎢ − ⎜ − ⎟⎥ ⎢⎣ R ⎝T T0 ⎠⎥⎦ Figure 1. Dielectric constant of binary liquid mixtures at 313.2 K: ▲, data17 for methanol (1)/water (2); ■, data28 for ethanol (1)/water (2); - - -, this work for methanol (1)/water (2); , this work for ethanol (1)/water (2).

(5)

Table 3 summarizes the properties of each pure component i used to calculate the dielectric constant of each binary liquid mixture applied in the study.

Table 4. Experimental Solubilities of Sucrose Expressed As Mass Fraction (ws) as a Function of Mole Fraction (x), Volume Fraction (Φ), and Dielectric Constant (ε) of the Binary Liquid Mixtures for Different Temperatures at 0.1 MPaa 303.2 K

a

x1

Φ1

ε

1.0000 0.9000 0.8000 0.6999 0.6000 0.5000 0.4000 0.3000 0.2000 0.1001 0.0000

1.0000 0.9457 0.8856 0.8186 0.7437 0.6593 0.5633 0.4533 0.3260 0.1771 0.0000

30.68 32.50 34.60 37.04 39.91 43.31 47.40 52.41 58.62 66.49 76.73

1.0000 0.9000 0.7992 0.6998 0.6000 0.5026 0.4019 0.3000 0.2000 0.1000 0.0000

1.0000 0.9622 0.9185 0.8684 0.8094 0.7409 0.6554 0.5481 0.4144 0.2392 0.0000

23.55 24.86 26.44 28.33 30.67 33.53 37.36 42.56 49.69 60.17 76.73

1.0000 0.9006 0.7999 0.6997 0.5998 0.5002 0.3999 0.3001 0.1999 0.1000 0.0000

1.0000 0.8610 0.7321 0.6143 0.5061 0.4062 0.3130 0.2267 0.1459 0.0706 0.0000

30.68 29.61 28.65 27.78 27.00 26.30 25.65 25.06 24.51 24.01 23.55

313.2 K ws

x1

Methanol (1) + Water (2) 0.008134 1.0000 0.014626 0.8999 0.032468 0.8001 0.074915 0.7000 0.159546 0.6000 0.272230 0.5000 0.414636 0.4002 0.518577 0.3000 0.617732 0.2000 0.664085 0.1000 0.698458 0.0000 Ethanol (1) + Water (2) 0.000941 1.0000 0.002238 0.8987 0.006264 0.8000 0.022237 0.7008 0.058875 0.5994 0.145163 0.5001 0.269579 0.3999 0.412350 0.2998 0.529703 0.2000 0.612895 0.1000 0.698458 0.0000 Methanol (1) + Ethanol (2) 0.008134 1.0000 0.006246 0.9004 0.004660 0.8000 0.003652 0.6998 0.002851 0.6003 0.002090 0.5000 0.001675 0.4001 0.001327 0.3002 0.001199 0.2000 0.000985 0.1001 0.000941 0.0000

Φ1

ε

ws

1.0000 0.9469 0.8881 0.8222 0.7483 0.6647 0.5695 0.4593 0.3314 0.1805 0.0000

29.03 30.73 32.69 34.98 37.67 40.89 44.76 49.54 55.50 63.12 73.12

0.010324 0.020116 0.044455 0.099994 0.201882 0.343116 0.468493 0.556299 0.622578 0.673142 0.707301

1.0000 0.9625 0.9204 0.8714 0.8123 0.7432 0.6584 0.5533 0.4197 0.2432 0.0000

22.20 23.44 24.89 26.65 28.87 31.63 35.25 40.11 46.90 56.98 73.12

0.001097 0.003512 0.009899 0.028370 0.076490 0.161615 0.316026 0.433982 0.572700 0.650959 0.707301

1.0000 0.8610 0.7327 0.6150 0.5072 0.4067 0.3137 0.2272 0.1463 0.0708 0.0000

29.03 28.01 27.08 26.26 25.51 24.83 24.21 23.65 23.12 22.64 22.20

0.010324 0.007703 0.005847 0.004612 0.003765 0.002984 0.002274 0.001967 0.001501 0.001387 0.001273

Standard uncertainties u are u(T) = 0.3 K, u(p) = 0.01 MPa, ur(x1) = 0.0005, ur(ws) = 0.02. C

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

3. RESULTS AND DISCUSSION In the interest of illustrating the approach applied to calculate the dielectric constant of the binary liquid mixtures, Figure 1 shows part of the results plotted against literature experimental data as a function of mass fraction of the binary liquid mixture. It was observed that the modeling used to calculate the dielectric constant of the binary mixture is able to reproduce the main features of the physical quantity. On the other hand, it seems that the model tends to underestimate the data. For example, the mixture of methanol/water and ethanol/water at 313.2 K presents a deviation between literature data and theoretical data of about 7.5% and 8.3%, respectively, in the equimolar mixture. The existing deviation is probably due to the theoretical approach, which does not take into account the chemical interaction related to possible cross-association of different molecules. The experimental solubilities expressed as mass fraction of sucrose for the three mixtures and at two different temperatures are presented in Table 4 along with mole fraction, volume fraction, and dielectric constant of the binary liquid mixtures (free of sucrose). It was estimated an average of the experimental solubility uncertainty as ±1.4 × 10−5. In order to illustrate the quality of the data presented in this work, Figure 2 shows part of the experimental values combined

Perez and Macedo (1997) studied the same binary system (methanol/water, ethanol/water, and methanol/ethanol) at 313.2 K and the comparison with the data of this work is reasonable. The calculus of average relative deviation between experimental data and those from the literature at the same molar composition of the binary liquid mixture give estimated values of about 5.5% for methanol/water, 7.6% for ethanol/water, and 18.1% for methanol/ethanol. Analyzing the data, it was observed that for all mixtures under investigation the solubility of sucrose has a higher magnitude for the higher temperature. Two steps may accomplish the dissolution of a solid in a liquid, first the solid melts and second the two liquids are mixed. Because the fusion stage is accompanied by an intake of heat,29 an increase of temperature leads to a larger amount of solid to dissolve. Sucrose is highly soluble in water followed, in a lower order of magnitude, by methanol and ethanol. The expressive difference in solubility is mainly due to the structure of the components. The molecule of sucrose has eight hydroxyl groups and three hydrophilic oxygen atoms and thus the molecule of water interacts easily by hydrogen bonds and for that reason the amount of sucrose that dissolves is high. On the other hand, methanol and ethanol have just one hydroxyl group, and the interaction among molecules of sucrose and alcohol is weak leading to a lower amount of solid to dissolve. Moreover, the molecule of ethanol has a chain length longer than the molecule of methanol and consequently the interaction of ethanol and sucrose is even weaker. Mixtures of water−methanol and water−ethanol present the capability to dissolve sucrose as a function of the amount of alcohol in the solution. For both mixtures, the addition of alcohol leads to a decrease of sucrose solubility. The same effect happens for the mixture of methanol and ethanol; the addition of ethanol leads to a decrease of sucrose solubility. The reduction of solubility due to the addition of an antisolvent is related to the decrease of the dielectric constant of the binary mixture. The solubilization of a solid depends directly on the dielectric constant of the mixture. Sucrose is a polar molecule that is highly soluble in water, which has a dielectric constant larger than methanol and ethanol, and methanol presents a dielectric constant larger than ethanol. A mixture of two of these three liquids leads to the formation of a solution with intermediate dielectric constant reflecting therefore in intermediate values of solubility as is observed in Table 4. Mixtures of water and methanol or water and ethanol have different structures for each alcohol concentration. The interaction among molecules cause a contraction of volume (negative excess molar volume)30,31 that may influence the attraction force and therefore there is a decrease of sucrose solubility. Figure 3 presents the behavior of sucrose solubility as a function of the dielectric constant of the binary liquid mixture for part of the experimental data obtained. The analyses of the experimental data related to the dielectric constant of the binary liquid mixture show that the solubility values may be satisfactory correlated. There is an indication that the temperature plays a more important role than the dielectric constant in the dissolution process because the solubility increases with the increase of temperature and the dielectric constant decreases with the increase of temperature. Figure 4 presents part of the experimental data as a function of the dielectric constant for mixtures of methanol and ethanol

Figure 2. Solubility of sucrose in pure water (a), methanol (b), ethanol (c) and in mixtures of methanol/water (d), ethanol/water (e), and methanol/ethanol (f) at 313.2 K: ×, this work; ●, Peres and Macedo;6 ▲, Bouchard et al.;8 ⧫, Gong et al.;9 ■, Van Putten et al.10

with values published. For the pure solvents, it is observed that the solubility of sucrose presented in this work compared with data available in the literature show significant difference indicating the need of further investigations. For mixtures, D

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

and development of mathematical models and theories of solution. The dependence of solubility with temperature and dielectric constant was observed. The temperature is more important than the dielectric constant of the binary liquid mixture for the dissolution process. There is the possibility of establishing correlations to describe solubility as a function of dielectric constant of the binary liquid mixture.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: +55 49 2049-9597. Figure 3. Solubility of sucrose as a function of the dielectric constant for mixtures formed by ethanol and water (a) and by methanol and ethanol (b): ■, 303.2 K; ▲, 313.2 K.

Notes

The authors declare no competing financial interest. Funding

The authors wish to thank FAPESC (Fundaçaõ de Amparo à Pesquisa e Inovaçaõ do Estado de Santa Catarina) project 2014TR3085 and CNPq (Conselho Nacional de Desenvolví mento Cientifico e Tecnológico) project 484037/2013-7 for financial support.



(1) Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107, 2411−2502. (2) van Haveren, J.; Scott, E. L.; Sanders, J. Bulk chemicals from biomass. Biofuels, Bioprod. Biorefin. 2008, 2, 41−57. (3) Huang, H.-J.; Ramaswamy, S.; Tschirner, U. W.; Ramarao, B. V. A review of separation technologies in current and future biorefineries. Sep. Purif. Technol. 2008, 62, 1−21. (4) Demirbas, M. F. Biorefineries for biofuel upgrading: a critical review. Appl. Energy 2009, 86, S151−S161. (5) Segur, J. B.; Miner, C. S. Sugar solubility, sucrose and dextrose in aqueous glycerol. J. Agric. Food Chem. 1953, 1, 567−569. (6) Peres, A. M.; Macedo, E. A. Phase equilibria of D-glucose and sucrose in mixed solvent mixtures: Comparison of UNIQUAC based models. Carbohydr. Res. 1997, 303, 135−151. (7) Tsavas, P.; Polydorou, S.; Voutsas, E. C.; Magoulas, K. G.; Naraghi, K.; Halling, P. J. Sucrose solubility in mixtures of water, alcohol, ester and acid. J. Chem. Eng. Data 2002, 47, 513−517. (8) Bouchard, A.; Hofland, G. W.; Witkamp, G.-J. Properties of sugar, polyol, and polysaccharide water-ethanol solutions. J. Chem. Eng. Data 2007, 52, 1838−1842. (9) Gong, X.; Wang, S.; Qu, H. Solid-liquid equilibria of D-glucose, D-fructose and sucrose in the mixture of ethanol and water from 273.2 to 293.2 K. Chin. J. Chem. Eng. 2011, 19, 217−222. (10) Van Putten, R.-J.; Winkelman, J. G. M.; Keihan, F.; Van der Waal, J. C.; De Jong, E.; Heeres, H. J. Experimental and modelling studies on the solubility of D-arabinose, D-fructose, D-glucose, Dmannose, sucrose and D-xylose in methanol and methanol-water mixtures. Ind. Eng. Chem. Res. 2014, 53, 8285−8290. (11) Lin, C.-P.; Lai, G.-H.; Tu, C.-H. Liquid-liquid equilibria, density, refractive index, and solubility for mixtures of water + methanol + heptane + methylbenzene or + dimethyl carbonate at T = 298.15 K. J. Chem. Eng. Data 2013, 58, 3265−3274. (12) Urréjola, S.; Sánchez, A.; Hervello, M. F. Refractive indices of lithium, magnesium, and copper(II) sulfates in ethanol-water solutions. J. Chem. Eng. Data 2010, 55, 482−487. (13) Orge, B.; Rodríguez, A.; Canosa, J. M.; Marino, G.; Iglesias, M.; Tojo, J. Variation of densities, refractive indices, and speeds of sound with temperature of methanol or ethanol with hexane, heptane, and octane. J. Chem. Eng. Data 1999, 44, 1041−1047. (14) Mutalik, V.; Manjeshwar, L. S.; Sairam, M.; Aminabhavi, T. M. Excess molar volumes, deviations in viscosity and refractive index of the binary mixtures of mesitylene with ethanol, propan-1-ol, propan-2ol, butan-1-ol, pentan-1-ol, and 3-methylbutan-1-ol at 298.15, 303.15, and 308.15 K. J. Mol. Liq. 2006, 129, 147−154.

Figure 4. Solubility of sucrose as a function of the dielectric constant of the binary liquid mixture at 303.2 K (a) and at 313.2 K (b): ×, methanol/ethanol, ●, methanol/water.

and for mixtures of methanol and water for both temperatures investigated in this work. A very interesting behavior for the mixtures of ethanol/water and methanol/ethanol is observed for both temperatures studied; the experimental data may be described by means of the same pattern. The empirical model suggested by eq 6 was applied to correlate the data of both mixtures covering the range of dielectric constant (ε) between 23.55 and 76.73 for 303.2 K and between 22.20 and 73.12 for 313.2 K. The nonlinear fitting was performed by a Levenberg−Marquardt method. The adjustable parameters of the model along with the fitting coefficient are presented in Table 5. The results indicate Table 5. Adjustable Parameters and Fitting Coefficient for the Empirical Model T/K

a

b

c

d

R2

303.2 313.2

0.69786 0.69476

1.13628 0.90955

43.73622 38.57324

0.14220 0.14676

0.99907 0.99893

that the model is able to correlate the data of both mixtures by using the same set of adjustable parameters. ws = a[1 + (b − 1)e−d(ε − c)]1/1 − b

REFERENCES

(6)

4. CONCLUSION Data of sucrose solubility in mixtures formed by water− methanol, water−ethanol, and methanol−ethanol for 303.2 and 313.2 K are presented as a function of the dielectric constant of the binary liquid mixture. The experimental data are important for fields that deal with solid−liquid equilibrium, being useful as a starting point for other studies as well as for the application E

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(15) Papanastasiou, G. E.; Ziogas, I. I. Physical behavior of some reaction media. 3. Density, viscosity, dielectric constant, and refractive index changes of methanol + dioxane mixtures at several temperatures. J. Chem. Eng. Data 1992, 37, 167−172. (16) Papanastasiou, G. E.; Papoutsis, A. D.; Kokkinidis, G. I. Physical behavior of some reaction media. Density, viscosity, dielectric constant, and refractive index changes of ethanol-dioxane mixtures at several temperatures. J. Chem. Eng. Data 1987, 32, 377−381. (17) Pugh, E. W.; Winslow, G. H. The Analysis of Physical Measurements; Addison-Wesley: Reading, MA, 1966. (18) Looyenga, H. Dielectric constants of heterogeneous mixtures. Physica 1965, 31, 401−406. (19) Kretschmer, C. B.; Wiebe, R. Thermodynamics of alcoholhydrocarbon mixtures. J. Chem. Phys. 1954, 22, 1697−1701. (20) Morrone, S. R.; Francesconi, A. Z. A model for excess volumes of salty water−acetonitrile mixtures at 298.15 K. Fluid Phase Equilib. 2012, 313, 52−59. (21) Funke, H.; Wetzel, M.; Heintz, A. New applications of the ERAS model. Thermodynamics of amine+alkane and alcohol+amine mixtures. Pure Appl. Chem. 1989, 61, 1429−1439. (22) Scharlin, P.; Steinby, K.; Domańska, U. Volumetric properties of binary mixtures of N,N-dimethylformamide with water or water-d2 at temperatures from 277.13 to 318.15 K. J. Chem. Thermodyn. 2002, 34, 927−957. (23) Ǻ kerlöf, G. Dielectric constants of some organic solvent-water mixtures at various temperatures. J. Am. Chem. Soc. 1932, 54, 4125− 4139. (24) Bender, T. M.; Van Hook, W. A. Molar volumes of CH3OH, CH3OD, and CD3OD, and apparent or excess molar volumes for (methanol-OH+water), (methanol-OD+deuterium oxide), and (methanol-OH+methanol-OD), at 288.15, 298.15, and 313.15 K. J. Chem. Thermodyn. 1988, 20, 1109−1114. (25) Jain, D. M.; Shah, V.; Rabadiya, S.; Oswal, S. Viscosity and excess molar volume of binary mixtures of methanol with n-butylamine and di-n-butylamine at 303.15, 313.15 and 323.15 K. Characterization in terms of ERAS model. J. Mol. Liq. 2009, 144, 65−70. (26) Ivanov, E. V. Volumetric properties of dilute solutions of water in ethanol and water-d2 in ethanol-d1 between T = (278.15 and 318.15) K. J. Chem. Thermodyn. 2012, 47, 162−170. (27) Perry, R. H.; Green, D. W. Perry’s Chemical Engineers’ Handbook, 7th ed.; McGraw-Hill: New York, 1997. (28) Wyman, J. The dielectric constant of mixtures of ethyl alcohol and water from −5 to 40°. J. Am. Chem. Soc. 1931, 53, 3292−3301. (29) Hurtta, M.; Pitkänen, I.; Knuutinen, J. Melting behaviour of D-sucrose, D-glucose and D-fructose. Carbohydr. Res. 2004, 339, 2267− 2273. (30) Herráez, J. V.; Belda, R. Refractive Indices, densities and excess molar volumes of monoalcohols + water. J. Solution Chem. 2006, 35, 1315−1328. (31) Nand, D.; Yadav, V. S. Topological investigation of some alkanol + water binary mixtures: molar excess volume at 298.15 and 308.15 K. Int. J. Chem. Res. 2012, 3, 34−39.

F

DOI: 10.1021/acs.jced.5b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX