Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Selective extraction of benzene from

benzene–cyclohexane mixture using 1-


ethyl-3-methylimidazolium tetrafluoroborate
ionic liquid
Cite as: AIP Conference Proceedings 2124, 020028 (2019); https://doi.org/10.1063/1.5117088
Published Online: 24 July 2019

M. A. Hashim, M. Zulhaziman, M. Salleh, et al.

ARTICLES YOU MAY BE INTERESTED IN

Benzene and cyclohexane separation using 1-butyl-3-methylimidazolium thiocyanate


AIP Conference Proceedings 1877, 020005 (2017); https://doi.org/10.1063/1.4999855

Exposure of particulate matter 2.5 (PM2.5) on lung function performance of construction


workers
AIP Conference Proceedings 2124, 020030 (2019); https://doi.org/10.1063/1.5117090

Application of response surface methodology to investigate the effect of different variables


on fusion slagging index
AIP Conference Proceedings 2124, 020033 (2019); https://doi.org/10.1063/1.5117093

AIP Conference Proceedings 2124, 020028 (2019); https://doi.org/10.1063/1.5117088 2124, 020028

© 2019 Author(s).
Selective Extraction of Benzene from Benzene–Cyclohexane
Mixture using 1-Ethyl-3-Methylimidazolium
Tetrafluoroborate Ionic Liquid
M. A. Hashim*1a, M. Zulhaziman M. Salleh1, Emad Ali2, Mohamed K. Hadj-Kali2
1
University of Malaya Centre for Ionic Liquids (UMCiL), Department of Chemical Engineering, University of
Malaya, Kuala Lumpur, Malaysia.
2
Chemical Engineering Department, King Saud University, P.O Box 800, Riyadh 11421, Saudi Arabia
*a
Corresponding author: alihashim@um.edu.my

Abstract. Separation of benzene and cyclohexane remains a challenging process in petrochemical industry due to their
almost similar boiling temperatures. The present work validated the performance of a reportedly benzene-selective ionic
liquid (IL), i.e. 1-ethyl-3-methylimidazolium tetrafluoroborate ([C2Mim][BF4]) to extract benzene from its mixture with
cyclohexane via liquid–liquid extraction. The ternary liquid–liquid equilibria of benzene + cyclohexane + [C2Mim][BF4]
was investigated at 25 oC and under 1 atm with feed concentration of benzene ranging from 10 to 60 wt %. Good agreement
was achieved between COSMO-RS prediction and the experimental data for the tie-lines obtained, whereby the root mean
square deviation (RMSD) was 2.5%. The non-random two-liquid (NRTL) model was also successfully employed to
correlate the experimental tie-lines, whereby the RMSD was 0.5%. The extractive performance of [C2Mim][BF4] was
compared with other organic solvents in the literature and a few common ILs employed in our previous work. At
equilibrium, IL was not present in the cyclohexane-rich layer, and the molar concentration of cyclohexane in the IL-rich
layer was less than 0.01. The results indicated an efficient extraction process as the cross-contamination between the phases
was minimal, and relatively less energy was required to recover the IL.

Keywords: Ionic liquid, benzene, cyclohexane, liquid–liquid extraction, COSMO-RS

INTRODUCTION

Benzene and cyclohexane are two valuable products being widely processed in petrochemical industry. Benzene
is an aromatic hydrocarbon which is commonly used as a raw material to synthesize compounds such as styrene,
phenol, cyclohexane, anilines, alkylbenzenes and chlorobenzenes. Cyclohexane, on the other hand, when converted
into the intermediate cyclohexanone, is a useful feedstock for nylon precursors. At present, virtually all cyclohexane
is produced by catalytic hydrogenation of benzene. In this process, high purity of cyclohexane is only achievable
through a complex process control which involves heat integration and economic analysis.1-2 Hence, the product of
this process is usually a benzene-cyclohexane mixture. In fact, the separation of benzene and cyclohexane has been
regarded as the most difficult process in petrochemical industry. It is challenging to separate them by conventional
distillation because they have nearly equal boiling points, with the difference of only 0.64 oC. This condition requires
an advanced distillation process such as the azeotropic or extractive distillations. However, the industrial applications
of both technologies are limited by disadvantages such as process complexity, high capital and operating costs, and
high energy consumption. Liquid–liquid extraction (LLE) appears to be an attractive solution to the problem due to
its simplicity and feasibility of operation under mild conditions. LLE is also an appropriate method to separate low
concentrations of aromatic compound (20–65%) in aromatic-aliphatic mixtures.3
The essential step in employing LLE is to find a suitable extracting solvent. An ideal solvent should have the
advantages of high selectivity and solvent capacity, low volatility, easy regeneration, low viscosity, and high thermal

6th International Conference on Environment (ICENV2018)


AIP Conf. Proc. 2124, 020028-1–020028-7; https://doi.org/10.1063/1.5117088
Published by AIP Publishing. 978-0-7354-1864-6/$30.00

020028-1
stability.4-6 In general, dimethylsulfoxide, sulfolane, N-methylpyrrolidone, N-formylmorpholine and methylcarbonate
are the most widely used solvents as extractants for aromatic compounds.7 For benzene-cyclohexane system, several
studies have demonstrated good performance of industrial organic solvents such as ethylene glycol, tetra-ethylene
glycol, sulfolane, and N-methylpyrrolidone as extractants.8-10 However, the application of these organic solvents is
associated with critical challenges because they are usually volatile, toxic and flammable. In contrast to conventional
solvents, ionic liquids (ILs) have been found to be promising solvents because of their high thermal stability, negligible
vapor pressure, nonvolatile nature, and high solution capacity.6, 11-14 Several ILs have shown good performance in
extraction of benzene from cyclohexane including [C4mim][BF4],11 [C4py][BF4],11 [C4mim][SCN],11
[C2mim][DMP],15 [Mim][BF4],15 [Mim][ClO4],15 [C4mim][AlCl4],16 [C2Mim][EtSO4],17 and [CnMim][PF6] (n = 4, 5,
6).18 Recently, our group has enumerated the list of potential ILs to be used as extracting solvents for the separation
of benzene and cyclohexane.19 Based on this article, 1-ethyl-3-methylimidazolium tetrafluoroborate ([C 2mim][BF4])
was deemed to be an interesting IL due its high value of selectivity at infinite dilution (S ∞). In the present study, the
actual performance of this IL was investigated through experimental works, computational simulation and
thermodynamic modelling.

METHODOLOGY

Chemicals and materials

The following chemicals were obtained from local suppliers and used without further purification: benzene (99.7
%) and cyclohexane (99.5%) from Merck and [C2mim][BF4] (≥98.0 %) from Sigma Aldrich. For the determination of
ternary composition in extract and raffinate phases at equilibrium, the NMR analysis was performed using deuterated
chloroform with purity ≥99.8% (stabilized with silver) and deuterated methanol.

Liquid–liquid extraction experiment


The feed mixture containing 10 wt% benzene in cyclohexane was firstly prepared inside a 20-mL screw-capped
scintillation vial using an analytical balance with accuracy of ±0.0001 g. This procedure was repeated for other
concentrations of benzene (i.e., 20, 30, 40, 50, and 60 wt%) in the feed. Based on the binary composition, the total
weight of the feed mixture was fixed at 2 g. The IL, [C2mim][BF4] was then mixed into every feed mixture in a weight
ratio of 1:1. Parafilm tape was used to seal each vial so as to avoid component loss due to evaporation. The vials were
then placed in an incubation shaker at 25 oC and under 1 atm, where spring clamps were used to tighten them upon
shaking. The shaking was subsequently operated at 200 rpm for 6 h. Then, the mixture was left for approximately 12
h to achieve equilibrium. This period is more than sufficient to ensure the equilibrium state was fully reached.

Determination of ternary molar compositions

A drop of sample was taken out from the extract and raffinate layers using a micropipette. For the extract layer, a
few bubbles were purged from the micropipette tip to avoid cyclohexane contamination from the raffinate. This drop
of sample was then dissolved in 0.7 mL of deuterated solvent placed inside an NMR tube. The solvents used to dissolve
top and bottom layers were deuterated chloroform and deuterated methanol, respectively. The sample and solvent in
the tube were carefully shaken to form a homogenous mixture. Each tube was tightly sealed with parafilm to avoid
chemical loss. The 1H NMR spectrometer of Bruker 400 MHz was used to measure the peak of hydrogen molecules
in each component and to calculate the molar fraction of each component in both layers.

Experimental selectivity and distribution ratio

The distribution ratio of benzene (DB) and the solvent selectivity (S) of [C2Mim][BF4] were used to evaluate the
extraction performance by employing Equations 1 and 2.
= 1, (1)
1,

020028-2
1,
1,
= ⁄ 2, (2)
2,

In these equations, X1 and X2 are the concentrations of benzene and cyclohexane, respectively. The numbers 1 and 2
represent the bottom and top layer, respectively. Experimental tie-lines were compared with those obtained from the
NRTL regression and the COSMO-RS prediction by calculating the Root Mean Square Deviation (RMSD) described
in our earlier work.20

RESULTS AND DISCUSSION

Ternary LLE data


The LLE data for the ternary system of [C 2Mim][BF4] + benzene + cyclohexane is tabulated in Table 1. The
experimental values of selectivity for [C2Mim][BF4] were in the range of 24.20–36.68. This range demonstrated that
[C2Mim][BF4] is a highly selective IL for the separation of benzene from benzene–cyclohexane mixture. Furthermore,
the selectivity observed in this work is much higher than the reported values using organic solvents such as sulfolane,21
ethylene glycol,22 ethylene carbonate,23 N-dimethylformamide22 and dimethylsulfoxide.24 In comparison to our
previous work,19 the selectivity range in this work is nearly similar to [C2Mim][SCN], but superior to [C2Mim][Ac],
[C2Mim][N(CN)2] and [C2Mim][Tf2N]. As the values of selectivity were far greater than unity, the liquid–liquid
extraction using the present IL is indeed achievable. However, the high range of selectivity was significantly traded
off with low range of benzene distribution ratio, where the values were less than 0.40. This trend could be expected
in any LLE process as solvent selectivity is usually inversely proportional to solute distribution ratio. The contrary
relation between selectivity and distribution ratio in this work is also in agreement with other works that used organic
solvents,22 ILs19, 25 and deep eutectic solvents (DESs)26 to separate benzene and cyclohexane. In addition, at higher
concentration of benzene in the feed, both distribution ratio and selectivity slightly decreased. This indicates a steady
performance of the IL to extract benzene from low to medium concentration in feed. As tabulated in Table 1, the
concentration of [C2mim][BF4] in the top layer was zero (X3,T=0) for all tie-lines, and the concentration of cyclohexane
in the bottom layer was very low (X 2,B<0.01). The absence of the IL in the top layer indicates a favorable extraction
process because the solvent cross-contamination could be avoided. In addition, the low concentration of cyclohexane
in the bottom layer indicated an easier regeneration of the IL.

TABLE 1. Molar composition of the tie-lines with the distribution ratio and selectivity data for benzene (1) + cyclohexane (2) +
[C2mim][BF4] (3) at 25 oC and under 1 atm

Top layer Bottom layer


D S
x1 x2 x3 x1 x2 x3
0.091 0.909 0.00 0.033 0.009 0.955 0.36 36.68
0.186 0.814 0.00 0.062 0.008 0.930 0.33 35.23
0.278 0.722 0.00 0.088 0.007 0.905 0.32 33.37
0.379 0.621 0.00 0.115 0.006 0.879 0.30 30.41
0.477 0.523 0.00 0.143 0.006 0.851 0.30 26.11
0.574 0.426 0.00 0.166 0.005 0.829 0.29 24.20

The ternary liquid–liquid equilibria for the six tie-lines investigated in this work is depicted in Fig. 1. As observed,
all tie-lines showed negative slope. This indicates that the concentration of benzene in the top layer is higher than in
the bottom (X1,T > X1,B). The negative slope also indicates higher affinity of benzene toward cyclohexane than it is
toward [C2mim][BF4], which supported the difficulty of separating benzene–cyclohexane mixture. This also suggested
the necessity to perform multistage extraction to extract large amount of benzene, especially at higher feed

020028-3
composition. The immiscibility region of this ternary LLE was slightly less than that using another highly selective
IL, i.e. [C2Mim][SCN] in our previous work.

0.0
1.0

0.2
0.8

0.4
0.6

0.6
0.4

0.8
0.2

1.0
0.0
0.0 0.2 0.4 0.6 0.8 1.0
C2mimBF4

FIGURE 1. Ternary LLE diagram for [C2mim][BF4] + benzene + cyclohexane at 25 oC and under 1 atm: ̶▲̶ , experimental; --
∆--, COSMO-RS; ··×··, NRTL.

Comparison between experimental and COSMO-RS data


As seen in Fig. 1, the COSMO-RS and experimental tie lines achieved very good agreement, especially in the
raffinate phase. It is worth noting that the accuracy of COSMO-RS prediction decreased at higher concentration of
benzene in feed (X1,F). The quantitative comparison between the experimental and COSMO-RS results can be
evaluated from their value of RMSD. The most accurate prediction was observed at the lowest tie-line (X1,F=0.1) with
RMSD of 2.9 %, and the least accurate prediction was shown at the highest tie line (X 1,F=0.6) with RMSD of 8.4%.
Nonetheless, the average RMSD for all tie-lines were only 6.4%. Considering COSMO-RS a fast and an a priori
method, this value supports the advantage of COSMO-RS as an assisting tool for thermodynamic properties
calculations. It can be observed that the discrepancy between COSMO-RS and experimental data was mostly in the
bottom layer, where COSMO-RS slightly overestimated the solubility of benzene toward cyclohexane and
significantly underestimated its solubility toward the IL. The extractive performance of IL can also be evaluated from
the values of selectivity or benzene distribution ratio with respect to molar concentration of benzene in the top layer
(X1,T), as shown in Fig. 2. As observed, the trend of extractive performance from COSMO-RS was the same as from
experimental, where selectivity and benzene distribution ratio decreased at higher concentration of benzene in the top
layer. By evaluating Equation 2, the underestimation of COSMO-RS for the solubility of benzene toward IL was again
demonstrated when X1,T>0.2, as displayed in Fig 2(b).

020028-4
50

45

40

35

30

25

20

15

10

0
0.0 0.2 0.4 0.6 0.8 1.0
Molar concentration of benzene in the top layer
(a) (b)
FIGURE 2. The distribution ratio (a) and selectivity (b) with respect to concentrations of benzene in the top layer. The full
triangle represents the experimental data, while the empty triangle indicates the COSMO-RS predictions

NRTL regression

The binary interaction parameters τij and τji of the Non-Random Two Liquid (NRTL) model27 were estimated by
minimizing the root mean square deviation (RMSD) between the calculated and the experimental solubilities of each
constituent in each phase using the Simulis® software package.28 The non-randomness parameter αij in the NRTL
model measures the non-randomness in the mixture; i.e. the mixture is said to be completely random when αij is zero.
In this work, αij was fixed equal to 0.20 for all binary combinations. The choice of αij = 0.2 is made following successful
fitting of NRTL model for ternary LLE data for systems containing ILs and deep eutectic solvents as reported in our
previous works.19, 26, 29 The RMSD value between experimental and NRTL calculation was only 0.5 %, which indicates
that NRTL correlation represents the experimental data very well. This excellent fitting is also observable in Fig. 1
(dashed lines). The values of NRTL binary interaction parameters regressed for this ternary system are shown in Table
2. In order to conserve coherence between this work and the previous one applied to deep eutectic solvents for the
same binary mixture, the binary interaction parameters between benzene and cyclohexane, despite the different IL
used, were taken from our previous works 19, 26 without any adjustment.

TABLE 2. NRTL parameters for the ternary system [C2mim][BF4] + benzene + cyclohexane
i–j
Benzene – Cyclohexane -189.57 453.65
Benzene – [C2mim][BF4] 3909.83 193.54
Cyclohexane – [C2mim][BF4] 1726.47 1418.19

CONCLUSION
In the present work, the extractive performance of [C 2Mim][BF4] was investigated for the separation of benzene
from its mixture with cyclohexane. There was no IL present in the cyclohexane layer and the concentration of
cyclohexane in IL layer was very low. A good agreement was observed for the tie lines between COSMO-RS,
experimental and NRTL. This study indicates that apart from being available commercially at affordable price,
[C2Mim][BF4] also showed high selectivity in the extraction of benzene from benzene–cyclohexane mixture. The
selectivity of [C2Mim][BF4] was also comparable to [C2Mim][SCN] in our previous work.

020028-5
ACKNOWLEDGEMENT
The authors extend their appreciation to the International Scientific Partnership Program (ISPP #130) at King Saud
University for funding this research work. The authors also express their gratitude to UMCiL at the University of
Malaya for the facilities provided.

REFERENCES
1. Kassel, L. S., Hydrogenation of benzene to cyclohexane. Patent No. US 2755317 A: 1956.
2. Robert, O.; Dang, V. Q., Manufacture of naphthenic hydrocarbons by hydrogenation of the corresponding
aromatic hydrocarbons. Patent No. US 3597489 A: 1971.
3. Brandrup, J.; Immergut, E. H.; Grulke, E. A.; Abe, A.; Bloch, D. R., Polymer handbook. Wiley New York:
1999; Vol. 89.
4. Lu, Y.; Yang, X.; Luo, G., Liquid− liquid equilibria for benzene+ cyclohexane+ 1-butyl-3-
methylimidazolium hexafluorophosphate. Journal of Chemical & Engineering Data 2009, 55 (1), 510-512.
5. Dong, H.; Yang, X.; Zhang, J., Liquid− liquid equilibria for benzene+ cyclohexane+ N, N-
dimethylformamide+ potassium thiocyanate. Journal of Chemical & Engineering Data 2010, 55 (9), 3972-3975.
6. Wang, R.; Li, C.; Meng, H.; Wang, J.; Wang, Z., Ternary liquid− liquid equilibria measurement for benzene+
cyclohexane+ N-methylimidazole, or N-ethylimidazole, or N-methylimidazolium dibutylphosphate at 298.2 K and
atmospheric pressure. J. Chem. Eng. Data 2008, 53 (9), 2170-2174.
7. Rydberg, J., Solvent Extraction Principles and Practice, Revised and Expanded. Taylor & Francis: 2004.
8. Ali, S. H.; Lababidi, H. M. S.; Merchant, S. Q.; Fahim, M. A., Extraction of aromatics from naphtha reformate
using propylene carbonate. Fluid Phase Equilib. 2003, 214 (1), 25-38.
9. Al-Sahhaf, T. A.; Kapetanovic, E., Measurement and prediction of phase equilibria in the extraction of
aromatics from naphtha reformate by tetraethylene glycol. Fluid Phase Equilib. 1996, 118 (2), 271-285.
10. Wang, W.; Gou, Z.; Zhu, S., Liquid−Liquid Equilibria for Aromatics Extraction Systems with Tetraethylene
Glycol. J. Chem. Eng. Data 1998, 43 (1), 81-83.
11. Zhou, T.; Wang, Z.; Ye, Y.; Chen, L.; Xu, J.; Qi, Z., Deep Separation of Benzene from Cyclohexane by
Liquid Extraction Using Ionic Liquids as the Solvent. Ind. Eng. Chem. Res. 2012, 51 (15), 5559-5564.
12. García, J.; García, S.; Torrecilla, J. S.; Oliet, M.; Rodríguez, F., Separation of toluene and heptane by liquid–
liquid extraction using z-methyl-N-butylpyridinium tetrafluoroborate isomers (z= 2, 3, or 4) at T= 313.2 K. The
Journal of Chemical Thermodynamics 2010, 42 (8), 1004-1008.
13. Meindersma, G. W.; Hansmeier, A. R.; de Haan, A. B., Ionic liquids for aromatics extraction. Present status
and future outlook. Industrial & Engineering Chemistry Research 2010, 49 (16), 7530-7540.
14. Garc a, J. n.; Garc a, S.; Torrecilla, J. S.; Oliet, M.; Rodr guez, F., Liquid− liquid equilibria for the ternary
systems {heptane+ toluene+ N-butylpyridinium tetrafluoroborate or N-hexylpyridinium tetrafluoroborate} at T=
313.2 K. Journal of Chemical & Engineering Data 2010, 55 (8), 2862-2865.
15. Salem, S.; Shen, C.; Li, C.-x., (Liquid + liquid) equilibria of {benzene + cyclohexane + two ionic liquids} at
different temperature and atmospheric pressure. J. Chem. Thermodyn. 2012, 49, 81-86.
16. Lyu, Z.; Zhou, T.; Chen, L.; Ye, Y.; Sundmacher, K.; Qi, Z., Simulation based ionic liquid screening for
benzene–cyclohexane extractive separation. Chem. Eng. Sci. 2014, 113 (0), 45-53.
17. González, E. J.; Calvar, N.; González, B.; Domínguez, Á., Liquid Extraction of Benzene from Its Mixtures
Using 1-Ethyl-3-methylimidazolium Ethylsulfate as a Solvent. J. Chem. Eng. Data 2010, 55 (11), 4931-4936.
18. Zhou, T.; Wang, Z.; Chen, L.; Ye, Y.; Qi, Z.; Freund, H.; Sundmacher, K., Evaluation of the ionic liquids 1-
alkyl-3-methylimidazolium hexafluorophosphate as a solvent for the extraction of benzene from
cyclohexane:(Liquid+ liquid) equilibria. The Journal of Chemical Thermodynamics 2012, 48, 145-149.
19. Salleh, Z.; Hadj-Kali, M.; Hashim, M. A.; Mulyono, S., Ionic liquids for the separation of benzene and
cyclohexane – COSMO-RS screening and experimental validation. J. Mol. Liq. 2018, 266, 51-61.
20. Salleh, M. Z. M.; Hadj-Kali, M. K.; Hizaddin, H. F.; Ali Hashim, M., Extraction of nitrogen compounds from
model fuel using 1-ethyl-3-methylimidazolium methanesulfonate. Sep. Purif. Technol. 2018, 196, 61-70.
21. De Fré, R. M.; Verhoeye, L. A., Phase equilibria in systems composed of an aliphatic and an aromatic
hydrocarbon and sulpholane. Journal of Applied Chemistry and Biotechnology 1976, 26 (1), 469-487.
22. Aspi; Surana, N. M.; Ethirajulu, K.; Vennila, V., Liquid−Liquid Equilibria for the Cyclohexane + Benzene
+ Dimethylformamide + Ethylene Glycol System. J. Chem. Eng. Data 1998, 43 (6), 925-927.

020028-6
23. Mohsen-Nia, M.; Doulabi, F. S. M., Liquid–liquid equilibria for mixtures of (ethylene carbonate + aromatic
hydrocarbon + cyclohexane). Thermochim. Acta 2006, 445 (1), 82-85.
24. Yang, X.; Zhang, X.; Dong, H.; Yue, G.; Liu, J., (Liquid + liquid) equilibria for (benzene + cyclohexane +
dimethyl sulfoxide) system at T = (298.15 or 303.15) K: Experimental data and correlation. J. Chem. Thermodyn.
2015, 84, 14-17.
25. Gómez, E.; Domínguez, I.; Calvar, N.; Palomar, J.; Domínguez, Á., Experimental data, correlation and
prediction of the extraction of benzene from cyclic hydrocarbons using [Epy][ESO4] ionic liquid. Fluid Phase Equilib.
2014, 361, 83-92.
26. Salleh, Z.; Wazeer, I.; Mulyono, S.; El-blidi, L.; Hashim, M. A.; Hadj-Kali, M., Efficient removal of benzene
from cyclohexane-benzene mixtures using deep eutectic solvents – COSMO-RS screening and experimental
validation. J. Chem. Thermodyn. 2017, 104, 33-44.
27. Renon, H.; Prausnitz, J. M., Local compositions in thermodynamic excess functions for liquid mixtures.
AlChE J. 1968, 14 (1), 135-144.
28. Simulis® thermodynamics http://www.prosim.net/.
29. Hizaddin, H. F.; Hadj-Kali, M. K.; Ramalingam, A.; Ali Hashim, M., Extractive denitrogenation of diesel
fuel using ammonium- and phosphonium-based deep eutectic solvents. J. Chem. Thermodyn. 2016, 95, 164-173.

020028-7

You might also like