Vapor–Liquid Equilibrium for Methyl Isobutyl ... - ACS Publications

Jun 7, 2017 - systems were correlated with the Wilson and NRTL activity coefficient models using the gamma−phi approach (γ−Φ) with the Hayden an...
0 downloads 0 Views 524KB Size
Article pubs.acs.org/jced

Vapor−Liquid Equilibrium for Methyl Isobutyl Ketone (MIBK) + (1Propanol or 2‑Propanol) Binary Mixtures Funmilola Avoseh, Samuel A. Iwarere, Caleb Narasigadu, and Deresh Ramjugernath* Thermodynamics Research Unit, School of Engineering, University of KwaZulu-Natal, Durban 4041, South Africa ABSTRACT: Isothermal vapor−liquid equilibrium (VLE) data for 1-propanol + methyl isobutyl ketone at 338.15, 353.15, and 368.15 K and 2-propanol + methyl isobutyl ketone were measured at 323.15, 338.15, and 353.15 K. All the measurements were undertaken on a recirculating VLE glass still operated at isothermal conditions. The experimental VLE data for all the measured binary systems were correlated with the Wilson and NRTL activity coefficient models using the gamma−phi approach (γ−Φ) with the Hayden and O’Connell correlation employed for calculating the second virial coefficients that are used in accounting for nonidealities in the vapor phase. The models provided satisfactory correlation of the measured alcohol−ketone-binary mixtures at all the temperatures reported. The thermodynamic consistency of the data was checked using the point test method of Van Ness et al. (Van Ness, H. C.; Byer, S. M.; Gibbs, R. E. AIChE J. 1973, 19, 238−244) as well as the more rigorous direct test of Van Ness (Van Ness, H. C. Pure Appl. Chem. 1995, 67, 859−872).



INTRODUCTION The extensive use of methyl isobutyl ketone (MIBK) makes it an important chemical in the petrochemical industry. This ketone is generally used as an industrial solvent. Methyl isobutyl ketone is further used as a chemical intermediate in the production of paints, rubber products, chemicals, and drugs. Ketones are constituent components found in the waste stream of the Fischer−Tropsch process, generally together with alcohols.3 Thermodynamic knowledge on the separation and purification of these alcohols + ketones binary mixtures is required as this helps in designing and optimizing the process plants in which they are found. There are isothermal VLE data available in the open literature4 for MIBK + 2-butanol system at 368 K, MIBK + tert-pentanol at 294, 338, and 368 K, and MIBK + 2-ethyl-1hexanol at 388 K. There are also extensive isobaric VLE data by Beatriz et al.5 for MIBK + cyclohexane and cyclohexene at 100 kPa. Ballard and Van Winkle6 reported measured isobaric VLE data for 2-propanol + methyl isobutyl ketone at 760 mmHg, while VLE binary data sets for 2-propanol + methyl isobutyl ketone at 330.0 and 340.0 K were reported by Psutka and Wichterle.7 In this work, isothermal VLE data were measured for 1propanol + MIBK at 338.15, 353.15, and 368.15 K and 2propanol + MIBK at 323.15, 338.15, and 353.15 K. Beyond the 2-propanol + MIBK system mentioned above, the authors are not aware of any other VLE data available in open literature for these systems. The newly measured VLE data were regressed using the combined gamma-phi (γ−Φ) method with the NRTL and Wilson activity coefficient models to account for the liquid phase nonideality and the Hayden and O’Connell8 correlation to calculate the second virial coefficient in the virial equation of © XXXX American Chemical Society

state for the vapor phase nonideality. The measured data were further subjected to thermodynamic consistency testing using the point test1 and direct test.2



EXPERIMENTAL SECTION

Chemicals. Methyl isobutyl ketone (CAS No. 108-10-1) and 2-propanol (CAS No. 67-63-0) were purchased from Merck, while 1-propanol (CAS No. 71-23-8) was purchased from Sigma-Aldrich. Purity checks were performed on each chemical before commencing with experimental measurements. The purity of these chemicals was checked by means of gas chromatographic analysis, as well as refractive index and density measurements. The refractive index of each chemical was measured at 293.15 K using the ATAGO 7000α refractometer with a reported standard uncertainty of 0.0001. The results were compared with refractive index values in the literature. The density (ρ) of each chemical was measured using the Anton Paar DMA 5000 densitometer with a reported standard uncertainty of 0.0005 g/cm3 in density and 0.03 K in temperature. According to the information provided by the manufacturer (Anton Paar),10 the DMA 5000 density meter has an accuracy and repeatability of ±0.000005 and 0.0000001 g/ cm3 for density, respectively, and a temperature accuracy of ±0.01 K and temperature repeatability of 0.001 K. The density values measured were compared with literature at 293.15 K. There was no further purification of the reagents as no significant impurities were found. The refractive index values Received: January 10, 2017 Accepted: May 13, 2017

A

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. List of Chemicals Used in This Study and Their Respective Purities (p = 0.1 MPa) density/g·cm−3

refractive index a

b

chemical

supplier

minimum purity claimed (mass %)

GC analysis (peak area %)

exp

1-propanol 2-Propanol MIBK

Sigma-Aldrich Merck Merck

≥99.5 ≥99.5 ≥99.0

99.3 99.4 99.8

1.38512 1.37718 1.39586

lit

d

1.3850 1.3776 1.3962

expc

litd

0.80366 0.78499 0.80030

0.8035 0.7855 0.7978

a As stated by the supplier. bRefractive index at 293.15 K, standard uncertainties u are u(T) = 0.01 K and u(nD) = 0.0004. cDensity at 293.15 K; u(ρ) = 0.0005 g·cm−3 and u(T) = 0.03 K. dLiterature values for the refractive indices and density taken from Lide.9

and purities of the chemicals from the gas chromatograph are presented in Table 1. Apparatus and Procedure. The VLE measurements were undertaken using a low pressure glass recirculating still that is described in previous publications.11,12 The VLE stills have been used extensively for low-pressure phase equilibria measurements in the Thermodynamics Research Unit at the University of KwaZulu-Natal. The schematic diagram, description of the apparatus, as well as the experimental procedure are available in the previous publication.11,12 The equilibrium temperature was measured with a Pt-100 temperature sensor calibrated against a WIKA CTH 6500 standard temperature sensor. The Pt-100 sensor was connected to a Hewlett-Packard 6 1/2 digit multimeter with has a 0.004% claimed accuracy, equivalent to 0.005 ohms in the measurement range, which translates to an uncertainty of 0.013 K. The digital multimeter enables the resistance of the sensor to be displayed and then converted to temperature in Kelvin using a calibration equation. Taking into account uncertainty from the calibration along with other sources of error in measuring the temperature, the combined uncertainty in the temperature measurement is estimated to be 0.18 K. A WIKA model P10 pressure transmitter was used to measure the system pressure. This pressure transmitter has a pressure range of 0 to 1 bar. This is in line with the maximum pressure limit of 101 kPa, for the dynamic recirculating glass still of Joseph et al.11,12 The combined standard uncertainty, U(p), in the pressure measurement was estimated to be 0.1 kPa. A Shimadzu 2014 gas chromatograph (GC-2014) equipped with a thermal conductivity detector was used to undertake the composition analyses with a CRS Porapak Q column. Helium gas was used as the carrier gas. The packed column is 4 m in length with an inner diameter of 2 mm. The maximum allowable operating temperature of the column is 250 °C. The equilibrium samples were withdrawn from the sample traps using a 0.5 μL gastight GC syringe, once the system is deemed to have attained equilibrium. In this study, equilibrium was assumed to have been reached when the pressure remained constant for about 45 min. Samples were taken in triplicate for the determination of the liquid and vapor compositions. The combined standard uncertainty for both the liquid and vapor compositions is 0.002 mole fraction.

Table 2. Physical Properties and Parameters Used for the Data Reduction in This Study pure component property

1-propanol

2-propanol

536.78 508.30 Tc13/K Pc13/kPa 5175 4762 Vc13/cm3·mol−1 219.00 220.00 compressibility factor (Zc)13 0.254 0.248 acentric factor (ω)13 0.629 0.665 dipole moment13/debye 1.7 1.7 radius of gyration14,15/Å 2.736 2.726 Association (ηii) and Solvation Parameters (ηij)15 n-PrOH 1.40 1.55 i-PrOH 1.55 1.32 MIBK 1.0 1.0 A 16.349 16.425 B 3592.820 3462.439 C 209.422 210.989

MIBK 574.60 3270 340.60 0.256 0.351 2.8 3.83 1.0 1.0 0.9 14.819 3519.985 229.739

coefficients for the MIBK + (1-propanol or 2-propanol) systems. Unlike methyl isobutyl ketone, 1-propanol and 2propanol have the tendency for self-association. Thus, the Hayden and O’Connell correlation uses a parameter to describe chemical association between the binary mixtures, which is dependent on the type of functional group the components are categorized. The association parameters used in this study were taken from Fredenslund et al.14 and Prausnitz et al.15 The Hayden and O’Connell correlation approach is reliable at low to moderate pressure for describing the behavior of two polar binary mixtures in the vapor phase because mixing rules are employed for predicting the cross coefficients, and association and solvation effect is accounted for. The liquid phase activity coefficient (γi) was calculated for the ketone− alcohol system using the conventional equilibrium condition for the γ−Φ approach presented in eq 1 yi ΦiP = xiγipisat

(i = 1, 2)

(1)

where xi and yi are the measured liquid phase and vapor phase mole fractions, respectively, P is the system pressure, pisat is the saturated vapor pressure for the pure component i at the system temperature T, and γi is the activity coefficient. The term (Φi) is given as follows



DATA REDUCTION OF NEWLY MEASURED VLE SYSTEMS The measured (P−x−y) VLE data for the ketone−alcohol binary mixtures were regressed with the combined γ−Φ method using the pure component properties listed in Table 2. The NRTL and Wilson liquid phase activity coefficient equations were used to account for any deviations from ideality in the liquid phase. For describing the vapor phase nonideality, the virial equation of state was used employing the Hayden and O’Connell8 correlation for the calculation of the second virial

⎛ φV ⎞ ⎡ ⎤ (ViL /cm 3·mol−1) (Pisat /kPa − P /kPa)⎥ Φi = ⎜⎜ isat ⎟⎟exp⎢ 3 −1 −1 ⎦ ⎝ φi ⎠ ⎣ (R /kPa·cm ·mol ·K )(T /K)

(2)

ϕvi

ϕsat i

where is the fugacity coefficient for the vapor phase, is the fugacity coefficient for the saturated vapor phase of the pure component, R is the universal gas constant, and VLi is the molar volume of the liquid phase. The truncated two term virial equation of state was employed for the computation of the pure component and mixture fugacity coefficients. The Rackett B

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

correlation16 was employed for the estimation of liquid molar volumes.

Table 3. Experimental Boiling Point Temperatures (Texp) and Literature Temperatures (Tlit) for 1-Propanol, 2Propanol, and Methyl Isobutyl Ketone at Its Saturated Vapor Pressure (P)a



RESULTS AND DISCUSSION The vapor pressures for the three components used in this study were measured and compared with the literature values17−19 as shown in Figure 1. The vapor pressures for 1-

P/kPa

Texp/K

1-Propanol 14.90 326.98 25.89 338.09 29.90 341.12 39.89 347.45 51.21 353.14 59.89 356.86 69.89 360.59 79.88 363.91 93.89 368.02 99.88 369.61 2-Propanol 14.90 314.43 23.43 323.03 29.90 327.97 39.89 333.98 48.34 338.14 59.89 342.90 69.89 346.45 79.88 349.64 92.40 353.14 98.88 354.82 Methyl Isobutyl Ketone 9.22 322.82 9.90 324.41 17.52 337.85 18.41 339.11 31.36 352.88 32.13 353.54 39.89 359.60 53.99 368.46 69.89 376.36

Figure 1. Experimental vapor pressures of 1-propanol (●), 2-propanol (■), and MIBK (▲) compared with literature values of (○),14 (□),15 and (Δ)16 for 1-propanol, 2-propanol, and MIBK, respectively.

propanol, 2-propanol, and methyl isobutyl ketone were measured in the temperature ranges of 327−369.6, 314.4− 354.8, and 322.8−376.4 K, respectively. The measured vapor pressures for 1-propanol, 2-propanol, and methyl isobutyl ketone were found to be in good agreement with the Antoine coefficients given by the authors such as Kemme and Kreps,17 or the Antoine coefficients calculated by NIST from Biddiscombe et al.18 and Fuge et al.’s19 data as reported on the NIST Chemistry WebBook20 with deviations from literature as reported in Table 3. The maximum relative errors are 0.10% for 1-propanol, 0.03% for 2-propanol, and 0.11% for methyl isobutyl ketone. The experimental vapor pressure data were regressed to obtain the Antoine parameter constants (A, B, C) presented in Table 2 using eq 3. The obtained values were used in the VLE data reduction algorithm B ln pisat /kPa = A − (i = 1, 2) (3) T /K + C − 273.15

Tlit./Kb

ΔT/Kc

327.16 338.36 341.43 347.77 353.50 357.21 360.95 364.27 368.37 369.97

−0.18 −0.27 −0.31 −0.32 −0.36 −0.35 −0.36 −0.36 −0.35 −0.36

314.47 323.01 327.87 333.87 338.03 342.83 346.40 349.57 353.10 354.78

−0.04 0.02 0.10 0.11 0.11 0.07 0.05 0.07 0.04 0.04

322.95 324.51 337.83 339.05 352.94 353.61 359.69 368.67 376.78

−0.13 −0.10 0.02 0.06 −0.06 −0.07 −0.09 −0.21 −0.42

a

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2). bNIST Chemistry WebBook.20 cΔT/K = (Texp − Tlit).

Table 4. Vapor−Liquid Equilibrium Dataa for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 1Propanol (1) + Methyl Isobutyl Ketone (2) at T = 338.15 K

VLE were measured at isothermal conditions for the 1propanol + MIBK and 2-propanol + MIBK systems with correlation of the data with two liquid-phase activity coefficient models, namely Wilson and NRTL equations. The Marquardt method21 with an objective function that minimizes the square root of the sum of the squares of the deviations in pressure and vapor phase composition was used for the fitting of the Wilson and NRTL parameters. The P−x−y data along with the experimental activity coefficients are listed in Tables 4−9 for the measured binary systems. The P−x−y plots for the 1propanol + MIBK binary mixture at 338.15, 353.15, and 368.15 K are shown in Figure 2. For the 2-propanol + MIBK system, the P−x−y plots at 323.15, 338.15, and 353.15 K are shown in Figure 3. In modeling the 1-propanol + MIBK and 2-propanol + MIBK binary systems, the vapor phase nonideality was taken into account using the second virial coefficients obtained from the Hayden and O’Connell correlation equation. The measured

P/kPa

x1

y1

17.57 21.32 23.11 24.53 25.16 25.81 26.09 26.61 26.95 26.98 26.94 26.84 26.46 25.92

0.000 0.114 0.202 0.304 0.375 0.445 0.499 0.635 0.707 0.792 0.839 0.888 0.953 1.000

0.000 0.229 0.359 0.444 0.499 0.556 0.587 0.677 0.727 0.788 0.827 0.872 0.942 1.000

γ1 1.651 1.581 1.380 1.295 1.243 1.186 1.100 1.067 1.033 1.020 1.010 1.003 1.000

γ2 1.000 1.038 1.037 1.096 1.133 1.157 1.208 1.327 1.406 1.540 1.620 1.713 1.827

a

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2). C

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 5. Vapor−Liquid Equilibrium Dataa for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 1Propanol (1) + Methyl Isobutyl Ketone (2) at T = 353.15 K P/kPa

x1

y1

31.41 37.91 41.61 44.45 46.03 47.29 48.18 49.76 50.66 51.26 51.66 51.64 51.54 51.00

0.000 0.110 0.209 0.300 0.377 0.442 0.495 0.645 0.729 0.796 0.843 0.886 0.936 1.000

0.000 0.230 0.362 0.455 0.523 0.568 0.607 0.696 0.747 0.808 0.845 0.885 0.931 1.000

γ1 1.559 1.420 1.320 1.254 1.192 1.158 1.057 1.018 1.016 1.006 1.004 0.999 1.000

Table 7. Vapor−Liquid Equilibrium Dataa for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 2Propanol (1) + Methyl Isobutyl Ketone (2) at T = 323.15 K

γ2

P/kPa

x1

y1

1.000 1.029 1.055 1.081 1.105 1.147 1.174 1.341 1.484 1.508 1.588 1.625 1.737

9.22 12.76 15.07 16.76 17.95 19.52 20.99 21.67 22.30 22.91 23.14 23.39 23.48

0.000 0.097 0.194 0.268 0.342 0.459 0.601 0.677 0.760 0.843 0.899 0.953 1.000

0.000 0.323 0.490 0.579 0.625 0.681 0.752 0.798 0.838 0.887 0.920 0.958 1.000

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2).

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2).

Table 8. Vapor−Liquid Equilibrium Dataa for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 2Propanol (1) + Methyl Isobutyl Ketone (2) at T = 338.15 K

a

Table 6. Vapor−Liquid Equilibrium Data for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 1Propanol (1) + Methyl Isobutyl Ketone (2) at T = 368.15 K x1

y1

γ1

54.04 63.68 70.25 76.25 79.52 82.17 84.15 89.97 91.80 92.57 93.43 93.90 93.92

0.000 0.111 0.206 0.300 0.375 0.442 0.497 0.739 0.811 0.850 0.892 0.937 1.000

0.000 0.220 0.365 0.464 0.527 0.577 0.617 0.769 0.821 0.858 0.897 0.939 1.000

1.364 1.342 1.255 1.190 1.141 1.110 1.001 0.988 0.992 0.995 0.997 1.000

1.809 1.640 1.541 1.400 1.235 1.120 1.087 1.046 1.022 1.007 0.999 1.000

γ2 1.000 1.013 1.022 1.020 1.085 1.221 1.383 1.435 1.594 1.741 1.942 2.214

a

a

P/kPa

γ1

γ2 1.000 1.042 1.046 1.072 1.107 1.145 1.176 1.473 1.601 1.612 1.635 1.671

P/kPa

x1

y1

17.57 28.92 32.18 38.07 41.39 43.03 44.68 46.13 46.91 47.71 48.45

0.000 0.190 0.275 0.460 0.594 0.675 0.756 0.841 0.895 0.950 1.000

0.000 0.486 0.561 0.688 0.771 0.814 0.854 0.898 0.927 0.963 1.000

γ1 1.538 1.375 1.186 1.114 1.076 1.043 1.018 1.005 1.000 1.000

γ2 1.000 1.020 1.090 1.224 1.294 1.365 1.477 1.636 1.806 1.955

a

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2).

Table 9. Vapor−Liquid Equilibrium Dataa for Pressure (P), Liquid-Phase Mole Fraction (x1), Vapor Phase Mole Fraction (y1), and Activity Coefficients (γ1, γ2) for 2Propanol (1) + Methyl Isobutyl Ketone (2) at T = 353.15 K

a

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2).

VLE binary systems show positive deviations from Raoult’s law with the 1-propanol + MIBK system exhibiting a maximum pressure azeotrope for the 338.15 and 353.15 K isotherms. The experimental azeotropic composition, x1AZ,exp and pressure PAZ,exp for the 1-propanol + MIBK system in this study were obtained by the numerical smoothing of the (x1−y1) versus x1 and P−x1 curves, which according to Pavlι ́ček and Wichterle22 is better than using correlative procedures. The azeotropic compositions for each isotherm are shown in Table 10. Representative plots of the activity coefficients (γ1, γ2) for 2propanol (1) + MIBK (2) and 1-propanol (1) + MIBK (2) at 323.15 and 338. 15 K are shown in Figures 4 and 5, respectively. The activity coefficients in Figures 4 and 5 are plotted alongside those calculated using the NRTL equation and found to be well-represented by the NRTL Gibbs excess model. The Wilson equation gave similar results to the NRTL equation for the calculation of the activity coefficients and hence only one of the models was used for the representative

P/kPa

x1

y1

31.41 42.51 61.56 65.66 69.08 76.20 79.66 83.35 86.70 88.57 90.61 92.48

0.000 0.099 0.336 0.399 0.455 0.590 0.670 0.756 0.839 0.895 0.948 1.000

0.000 0.301 0.618 0.671 0.710 0.789 0.827 0.866 0.907 0.936 0.969 1.000

γ1 1.425 1.246 1.211 1.178 1.108 1.070 1.037 1.016 1.006 1.003 1.000

γ2 1.000 1.030 1.101 1.114 1.135 1.205 1.284 1.405 1.534 1.657 1.655

a

U(T) = 0.15 K (k = 2); U(P) = 0.1 kPa (k = 2); U(x, y) = 0.002 (k = 2).

D

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 4. Comparison of experimental (■, □) liquid phase activity coefficients (γ1, γ2) and those calculated using NRTL model () for the 2-propanol + MIBK (2) system at 323.15 K.

Figure 2. P−x1−y1 VLE data for 1-propanol (1) + MIBK (2) at (from bottom to top) T = 338.15, 353.15, and 368.15 K: (■, □), experimental values; (−), calculated from regressed NRTL parameters.

Figure 5. Comparison of experimental (□, ■) liquid phase activity coefficients (γ1, γ2) and those calculated using NRTL model () for the 1-propanol (1) + MIBK (2) system at 338.15 K.

Figure 3. P−x1−y1 VLE data for 2-propanol (1) + MIBK (2) at (from bottom to top) T = 323.15, 338.15, 353.15 K: (■, □), experimental values; (−), calculated from regressed NRTL parameters.

representative plot of the point test for the 2-propanol + MIBK system at 323.15 K is shown in Figure 6. The use of the more rigorous Van Ness2 direct test for thermodynamic consistency check for all the measured VLE data fall within the consistency index of 2 for the RMS δ ln(γ1/ γ2), with the exception of the 1-propanol + MIBK binary mixture at 368.15 K which falls under the consistency index of 3 as reported in Table 11. The NRTL and Wilson models fitted the data sufficiently well and only marginal differences could be observed from the results of both models, as can be seen in Table 10. An example of the Van Ness2 consistency direct test is shown in Figure 7 for the 1-propanol + MIBK system at 338.15 K. Comparisons and Predictions. The only data sets available for comparison with the experiments in this study are those of Psutka and Wichterle7 for 2-propanol + MIBK. However, no direct comparison is possible because their data were at 330.0 and 340.0 K. Therefore, in order to demonstrate the reliability of the experimental data and modeling procedure for the new binary data sets, the data from Psutka and Wichterle7 at both temperatures of 330.0 and 340.0 K were regressed using the NRTL two parameter Gibbs excess model with α fixed at 0.47 as in this study. The regressed parameters were then used in predicting the P−x−y diagram at temperatures of 323.15 and 338.15 K, which are close to the temperatures of 330.0 and 340.0 K reported for 2-propanol + MIBK by Psutka and Wichterle.7 The results from the

Table 10. Experimental Azeotropic Composition x1 and pressure PAZ,exp for 1-Propanol (1) + Methyl Isobutyl Ketone (2) System AZ,exp

T/K

x1AZ,exp

PAZ,exp (kPa)

338.15 353.15 368.15

0.780 0.872 0.995

26.98 51.65 93.92

plots shown in Figures 4 and 5. The regressed NRTL and Wilson parameters for all the measured VLE data are reported in Table 11. Although, the thermodynamic consistency test does not confirm the “correctness” of the measured data, it is a test that is deemed necessary in validating the quality and the degree of reliability of the measured data. According to Danner and Gess,23 VLE data is deemed to be thermodynamically consistent if the absolute deviation between the experimental vapor phase composition and its calculated values obtained through Van Ness et al.1 point test is ≤0.01 mole fraction. Thus, the average absolute values for the vapor phase composition (Δy1) reported in Table 11 was found to satisfy the consistency criterion recommended by Danner and Gess23 for all the measured binary systems, except the 1-propanol + MIBK binary mixture at 368.15 K that has a Δy1 of 0.012. A E

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 11. Model Parameters (Δg12, Δg21, and α12 for NRTL Model); (Δλ12 and Δλ21 for Wilson Model) and Average Absolute Deviations for the Correlation of the {1-Propanol (1) + Methyl Isobutyl Ketone (2)} and {2-Propanol (1) + Methyl Isobutyl Ketone (2)} VLE Data to NRTL and Wilson Modelsa model NRTL

newly measured binary mixtures T/K = 338.15

T/K = 353.15

1-Propanol (1) + Methyl Isobutyl Ketone (2) Δg12/(J mol−1) 764.43 534.80 Δg21/(J mol−1) 1497.51 1462.16 α12 0.47 0.47 Δy1 0.004 0.006 ΔP /kPa 0.074 0.140 RMS δ ln(γ1/γ2) 0.036 0.044 Wilson Δλ12/(J mol−1) 2685.46 2687.11 Δλ21/(J mol−1) −405.83 −676.01 Δy1 0.004 0.006 ΔP /kPa 0.073 0.138 RMS δ ln(γ1/γ2) 0.040 0.050 NRTL T/K = 323.15 T/K = 338.15

T/K = 368.15 427.93 1285.44 0.47 0.012 0.288 0.094 2532.72 −811.50 0.012 0.286 0.102 T/K = 353.15

2-Propanol (1) + Methyl Isobutyl Ketone (2) Δg12/(J mol−1) 1466.82 1343.70 Δg21/(J mol−1) 834.25 709.50 α12 0.47 0.47 Δy1 0.005 0.005 ΔP /kPa 0.043 0.076 RMS δ ln(γ1/γ2) 0.040 0.045 Wilson Δλ12/(J mol−1) 1937.07 1848.41 Δλ21/(J mol−1) 380.84 217.23 Δy1 0.005 0.005 ΔP/kPa 0.044 0.075 RMS δ ln(γ1/γ2) 0.040 0.045 a

Figure 7. Direct test for thermodynamic consistency for 1-propanol (1) + MIBK (2) system at 338.15 K.

1664.00 147.33 0.47 0.006 0.107 0.043 1321.74 500.16 0.006 0.104 0.044

Figure 8. Comparison and prediction of P−x−y diagram for the 2propanol (1) + MIBK (2) system. Experimental: □, ■, this work at 323.15 and 338.15 K; ○, ●, 330.0 K and Δ, ▲, 340.0 K, Psutka and Wichterle;7 ―, NRTL model for this work; ----, NRTL model for Psutka and Wichterle.7

n exp ΔP/kPa = 1/n∑ni |Pexp − Pcalc − ycalc i i |; Δy = 1/n∑i |yi i |.

Figure 6. Point test for thermodynamic consistency for 2-propanol (1) + MIBK (2) system at 323.15 K.

ThermoData Engine (TDE) software implemented in ASPEN24 and the prediction using parameters obtained by regressing the VLE data of Psutka and Wichterle.7 According to the prediction using the regressed binary parameters from literature, the relative deviations of the pure component vapor pressure were 0.19% and −0.28% for the MIBK at 323.15 and 338.15 K, respectively. The relative deviations of the pure component vapor pressure for 2-propanol was −1.42% and −1.39% at 323.15 and 338.15 K, respectively. When the pure component vapor pressures for the 2-propanol and MIBK in this study were revaluated at 323.15 and 338.15 K using the NIST TDE software, the relative deviations obtained were −0.21% and −0.42% for the MIBK at 323.15 and 338.15 K, respectively, and the values for 2-propanol were −0.49% and −0.52% at 323.15 and 338.15 K, respectively.

predictions using the NRTL two parameter Gibbs excess models from the data of Psutka and Wichterle7 to this study presented in Figure 8 shows that the model predictions for 2propanol + MIBK based on this study compare reasonably well. Comparisons were made between the 2-propanol and MIBK pure component vapor pressures in this study and those obtained by re-evaluating vapor pressures using the NIST

CONCLUSIONS Three isothermal VLE data sets were each measured for the 1propanol + MIBK and 2-propanol + MIBK systems, respectively. All six isotherms were successfully correlated using a combination of the NRTL and Wilson liquid phase Gibbs excess models with the Hayden and O’Connell correlation for calculating the second virial coefficients. The 1-propanol + MIBK system exhibited an azeotrope at 338.15



F

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(3) De Klerk, A. Fischer−Tropsch refining: technology selection to match molecules. Green Chem. 2008, 10, 1237−1344. (4) Laavi, H.; Zaitseva, A.; Pokki, J.-P.; Uusi-kyyny, P.; Kim, Y.; Alopaeus, V. Vapor−Liquid Equilibrium, Excess Molar Enthalpies, and Excess Molar Volumes of Binary Mixtures Containing Methyl Isobutyl Ketone (MIBK) and 2-Butanol, tert-Pentanol, or 2-Ethyl-1-hexanol. J. Chem. Eng. Data 2012, 57, 3092−3101. (5) Marrufo, B.; Loras, S.; Sanchotello, M. Isobaric Vapor - Liquid Equilibria for Binary and Ternary Mixtures with Cyclohexane, Cyclohexene, and Methyl Isobutyl Ketone at 100 kPa. J. Chem. Eng. Data 2010, 55, 5812−5817. (6) Psutka, S.; Wichterle, I. Isothermal vapour liquid equilibria in the binary and ternary systems composed of 2-propanol, diisopropyl ether and 4-methyl-2-pentanone. Fluid Phase Equilib. 2005, 235, 58−63. (7) Ballard, L. H.; Van Winkle, M. Vapor-liquid equilibria at 760 mm pressure 2-propanol + methyl propyl ketone and 2-propanol + methyl isobutyl ketone systems. Ind. Eng. Chem. 1953, 45, 1803−1804. (8) Hayden, J. G.; O’Connell, J. P. Generalized Method for Predicting Second Virial Coefficients. Ind. Eng. Chem. Process Des. Dev. 1975, 14, 209−216. (9) Lide, D. R. Physical Constants of Organic Compounds. In CRC Handbook of Chemistry and Physics; 76th ed.; CRC Press: Boca Raton, FL, 1995−1996. (10) Anton Paar Density Meters: DMA Generation M. http://www. anton-paar.com/corp-en/products/details/density-meters-dmageneration-m/ (accessed May 1, 2017). (11) Joseph, M. A.; Raal, J. D.; Ramjugernath, D. Phase equilibrium properties of binary systems with diacetyl from a computer controlled vapour-liquid equilibrium still. Fluid Phase Equilib. 2001, 182, 157− 176. (12) Joseph, M. A.; Ramjugernath, D.; Raal, J. D. Computer−aided Measurement of Vapor-Liquid Equilibria in a Dynamic Still at SubAtmospheric Pressures. Dev. Chem. Eng. Miner. Process. 2002, 10, 615− 637. (13) Poling, B. E.; Prausnitz, J. M.; O’Connell, J. P. The Properties of Gases and Liquids, 5th ed.; McGraw-Hill: New York, 2001. (14) Fredenslund, A.; Gmehling, J.; Rasmussen, P. Vapor-Liquid Equilibria Using Unifac: A Group-Contribution Method; Elsevier: Amsterdam, 1977. (15) Prausnitz, J. M.; Anderson, T. F.; Grens, E. A.; Eckert, C. A.; Hsieh, R.; O’Connell, J. P. Computer Calculations for Multicomponent Vapor-Liquid and Liquid-Liquid Equilibria; Prentice Hall: Englewood Cliffs, NJ, 1980. (16) Rackett, H. G. Equation of State for Saturated Liquids. J. Chem. Eng. Data 1970, 15, 514−517. (17) Kemme, H. R.; Kreps, S. I. Vapour Pressure of Primary n-Alkyl Chlorides and Alcohols. J. Chem. Eng. Data 1969, 14, 98−102. (18) Biddiscombe, D. P.; Collerson, R. R.; Handley, R.; Herington, E. F. G.; Martin, J. F.; Sprake, C. H. S. Thermodynamic Properties of Organic Oxygen Compounds. Part 8. Purification and Vapor Pressures of the Propyl and Butyl Alcohols. J. Chem. Soc. 1963, 1954−1957. (19) Fuge, E. T. J.; Bowden, S. T.; Jones, W. J. Some Physical Properties of Diacetone Alcohol, Mesityl Oxide and Methyl Isobutyl Ketone. J. Phys. Chem. 1952, 56, 1013−1016. (20) NIST Chemistry WebBook. http://webbook.nist.gov/ chemistry/ (accessed 5 July 2015). (21) Marquardt, D. W. An Algorithm for Least-Squares Estimation of Non-Linear Parameters. J. Soc. Ind. Appl. Math. 1963, 11, 431−441. (22) Pavlíček, J.; Wichterle, I. Isothermal (vapour + liquid) equilibria in the binary and ternary systems composed of 2-propanol, 2,2,4trimethylpentane, and 2,4-dimethyl-3-pentanone. J. Chem. Thermodyn. 2012, 45, 83−89. (23) Danner, R. P.; Gess, M. A. A Data Base Standard for the Evaluation of Vapor-Liquid Equilibrium Models. Fluid Phase Equilib. 1990, 56, 285−301. (24) AspenTech; Aspen Technology, Inc.: Massachusetts, 2016, http://aspentech.com.

and 353.15 K. Experimental values of the activity coefficients for the 1-propanol + MIBK and 2-propanol + MIBK systems obtained using the traditional approach compared well with those obtained from the NRTL and Wilson equation. The measured VLE data were found to be thermodynamically consistent based on the point test of Van Ness et al.1 In addition, the use of the rigorous direct test of Van Ness2 gave good RMS values for the residuals δ ln(γ1/γ2) for both the 1propanol + MIBK and 2-propanol + MIBK systems.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel: +27 (0)31 2603128. ORCID

Samuel A. Iwarere: 0000-0001-8566-6773 Deresh Ramjugernath: 0000-0003-3447-7846 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors would like to thank the Department of Science and Technology and National Research Foundation for the financial support to carry out this study under the South African Research Chair for Fluorine Process Engineering and Separation Technology.



NOMENCLATURE A Antoine vapor pressure equation constant B Antoine vapor pressure equation constant C Antoine vapor pressure equation constant (gij − gii) energy parameters in the NRTL equation (J/mol) P system pressure (kPa) Pc critical pressure (kPa) pisat saturated vapor pressures for the ith component with (i = 1, 2) (kPa) R universal gas constant (J/mol K) T temperature (K) Tc critical temperature (K) Vc critical volume (cm3·mol−1) Zc critical compressibility factor x1, x2 measured liquid-phase mole fractions of components 1 and 2 y1, y2 measured vapor-phase mole fractions of components 1 and 2 Greek Letters

αij

nonrandomness parameter in the NRTL equation (dimensionless) (λij − λii) energy parameters in the Wilson equation (J/mol) Δ residual (e.g., ΔP, Δy) γi liquid-phase activity coefficients for the ith component with (i = 1, 2) Superscripts

AZ azeotrope exp experimental lit literature



REFERENCES

(1) Van Ness, H. C.; Byer, S. M.; Gibbs, R. E. Vapour-Liquid Equilibrium; Part1. An Appraisal of Data Reduction methods. AIChE J. 1973, 19, 238−244. (2) Van Ness, H. C. Thermodynamics in the treatment of vapour/ liquid equilibrium (VLE) data. Pure Appl. Chem. 1995, 67, 859−872. G

DOI: 10.1021/acs.jced.7b00026 J. Chem. Eng. Data XXXX, XXX, XXX−XXX