Process Safety and Environmental Protection 122 (2019) 1–12
Contents lists available at ScienceDirect
Process Safety and Environmental Protection
journal homepage: www.elsevier.com/locate/psep
Comparison of pressure-swing distillation and heterogeneous
azeotropic distillation for recovering benzene and isopropanol from
wastewater
Yue Cui, Xiaojing Shi, Chao Guang, Zhishan Zhang ∗ , Chao Wang, Chen Wang
College of Chemical and Environmental Engineering, Shandong University of Science and Technology, Qingdao, 266590, China
a r t i c l e
i n f o
Article history:
Received 2 October 2018
Received in revised form
18 November 2018
Accepted 20 November 2018
Available online 22 November 2018
Keywords:
Pressure-swing distillation
Heterogeneous azeotropic distillation
Wastewater
Benzene
Isopropanol
Heat-intergration
a b s t r a c t
Two methods of pressure-swing distillation (PSD) and heterogeneous azeotropic distillation (HAD) are
investigated to recover benzene and isopropanol from wastewater. This ternary mixture can form two
homogenous azeotropes (binary) and two heterogeneous azeotropes (binary and ternary), which all are
sensitive to pressure. The proposed configurations are rigorously simulated and optimized based on
the minimum total annual cost (TAC) via the sequential iterative procedure, and meanwhile the partial
heat integration is taken into account respectively. The results show that both configurations of PSD are
superior to the HAD configuration combining with PSD from the perspectives of economic and energy.
© 2018 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
1. Introduction
In the isopropanol (IPA) related industries, large amounts of
wastewater containing benzene and IPA are produced. Both IPA and
benzene are widely used as good organic solvents in many fields.
Therefore, it is attractive and necessary to separate and reuse
them from the wastewater in order to protect the environment
and conserve resources. However, this mixture can form multiple
complex azeotropes (heterogeneous and homogenous, binary and
ternary), the conventional methods cannot effectively accomplish
the separation of this mixture.
There have been some methods for separating the azeotropic
mixtures, such as pressure-swing distillation (PSD) (Cao et al., 2017;
Fissore et al., 2006; Kiran and Jana, 2015; Knapp and Doherty, 1992;
Lee et al., 2011; Li and Xu, 2017; Liang et al., 2017; Lladosa et al.,
2011; Luyben, 2012, 2018; Modla et al., 2010; Wang et al., 2015;
Zhu et al., 2016), azeotropic distillation (Arifin and Chien, 2007;
Chaniago et al., 2016; Le et al., 2015; Tabari and Ahmad, 2015;
Zhao et al., 2017), extractive distillation (Modla, 2013; Wang et al.,
∗ Corresponding author.
E-mail address: tjzza@163.com (Z. Zhang).
2018a,b; Wang et al., 2018d), and membrane separation (Ahmad
et al., 2018; Mukherjee et al., 2003; Xia et al., 2011).
Pressure-swing distillation is a kind of prevail method taking
advantage of the feature that the azeotrope composition has a great
shift with the change of column pressure. It can avoid the potential problems of adding third-party components. Recently, this
approach is gradually applied to the separation of the mixtures containing multiple azeotropes. Zhu et al. (2016) proposed a process
for separation of acetonitrile/methanol/benzene ternary azeotrope
by the PSD, and compared four different separation configurations.
Luyben (2017) studied the process of separating ternary mixture acetonitrile/methanol/benzene from different pressure-swing
distillation sequences. Aurangzeb and Jana (2018) explored the separation of ternary mixtures of acetonitrile/methanol/benzene via
pressure-swing dividing-wall column. Of course, partial and full
heat-integration technologies are often coupled with PSD processes
due to its economic attractiveness (Chen et al., 2018; Luyben, 2014;
Qasim et al., 2015; Zhang et al., 2017, 2016).
Azeotropic distillation sometimes is an attractive special separation method as well, especially heterogeneous azeotropic
distillation (HAD). Homogeneous azeotropic distillation needs add
an entrainer that can generate a new minimum-boiling azeotrope
with the original one or more components, which can be separated
as the distillate (Nava and Krishna, 2004). HAD is adapted to the
https://doi.org/10.1016/j.psep.2018.11.017
0957-5820/© 2018 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
2
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Nomenclature
AC
AR
CSCinst
H
HAD
ID
LP
M&S
MP
NF
NR
NT
P
PSD
QC
QR
RR
TAC
TCinst
UC
UR
T
heat transfer area of condenser [m2 ]
heat transfer area of reboiler [m2 ]
installed cost of column shell [$]
column height [m]
heterogeneous azeotropic distillation
inner diameter [m]
low pressure
Marshall & Swift Equipment Cost Index
middle pressure
feed location
recycle location
number of stages
operation pressure
pressure-swing distillation
condenser duty [kW]
reboiler duty [kW]
reflux ratio
total annual cost [$/y]
installed cost of trays [$]
heat transfer coefficient of condenser [kW/(K·m2 )]
heat transfer coefficient of reboiler [kW/(K·m2 )]
temperature difference [K]
separation of the mixtures containing minimum-boiling heterogeneous azeotropes (Wang and Huang, 2012). In general, it is much
easier than the separation of homogeneous azeotropes because the
split of two liquid phases in a decanter can be used to facilitate the
separation.
The purpose of this article is to recover benzene and IPA from
wastewater. Three separation configurations are proposed based
on the non-ideal features of the water/benzene/IPA system, including two PSD sequences and one combined sequence of HAD and
PSD. With the goal of minimum total annual cost (TAC), the optimal
parameters are determined by the global sequential iterative search
method. Furthermore, the partial heat-integration technology is
applied to the optimal configuration for enhancing energy efficiency. Finally, the preferred solution for the wastewater treatment
is confirmed through the comparisons of economic and energysaving.
2. Design basis
Table 1
NRTL model parameters for the water/benzene/IPA system.
Component i
Component j
water
benzene
water
IPA
benzene
IPA
Aij
Aji
Bij
Bji
Cij
140.0874
45.1905
−5954.31
591.3676
0.2
6.8284
−1.3115
−1483.46
426.3978
0.3
−0.7484
−0.1279
713.3123
148.1214
0.3
Fig. 1. The ternary diagram of the water/benzene/IPA system at 1 atm.
The non-ideal features of the water/benzene/IPA system at
atmospheric pressure are illustrated in Fig. 1, which can be applied
to the feasibility analysis for the developed separations solutions.
There are three distillation regions divided by distillation boundaries and one heterogeneous region in the ternary diagram. In each
region, the component removed from the bottom of the column
can be indicated along the residue curves from the unstable node
toward the stable node. Note that these characteristics are also
sensitive to the pressure variation. Therefore, two methods are
alternative for this separation problem, including pressure-swing
distillation with different sequences and heterogeneous azeotropic
distillation.
2.1. Problem statement
2.3. Economic optimization and CO2 emission
An industrial separation process is simulated with following
data: the handled feed is an effluent from a pharmaceutical company consisting of 80 mol% water, 12 mol% benzene and 8 mol% IPA
with the flowrate of 100 kmol/h. The purity of the recovered products is 99.99 mol% benzene and 99.99 mol% IPA, and the purified
water with the composition of 99.999 mol% is discharged. In this
work, the Aspen Plus v10.0 software is applied to the steady date
simulation and the optimization results validation.
2.2. Non-ideality of system
In the simulated study, the thermodynamic property method
is fairly important for the creditability of simulation results. The
non-random two liquid (NRTL) model with built-in binary interaction parameters (Table 1) is used to describe the phase equilibrium
behaviors of the system in this work, because of the agreement
between the predicted values and experiment data from NIST
database.
In order to make the process assessed from aspects of economy
and environment impact, the minimum total annual cost (TAC) is
used as the economic optimization objective as well as CO2 emission is calculated. Table 2 provides the basis of economics, the sizing
relationships and parameters (Zhu et al., 2015). The major pieces of
equipment including column vessels and plates, reboilers and condensers are concerned. Other items with small costs such as tubes,
pumps, and valves are ignored in the preliminary design stage. The
operating costs mainly consist of the consumption of steam and
cooling water. The sequential iterative optimization method in this
study is used to obtain the optimal parameters of three distillation
columns, as shown in Fig. 2. During the optimization, the convergent condition is that the relative error of TAC as the objective
function is within 1%. The constraints are the purity requirements
of three products maintained by adjusting the corresponding bottom flowrates. In addition, the initial reflux ratios are assumed to
be 1 and a step size of ±0.1 is used to update them. When reach-
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
3
ing near the minimum TAC, a smaller step size of ±0.01 is used to
determine an optimum reflux ratio accurately.
CO2 emission can be calculated for a given amount of fuel burnt
through the expression given in the literature (Gadalla et al., 2006;
Wang et al., 2018c). A detailed calculation procedure is shown as
follows:
[CO2] emission =
Amt fuel
NHV
C
×
100
˛
(1)
Where ␣ is the molar mass ratio of CO2 and C, 3.67; NHV is the net
heating value; and C% is the carbon content of a specific fuel. Coal
was taken as the fuel in this study, for which values for NHV and
C% are 22,000 kJ/kg and 0.865 kg/kg. Amtfuel is the amount of fuel
burnt (kW) and depends on the heat duty (Qseq) according to Eq.
(2),
Amt fuel =
Qseq × hseq − 419 ×
seq
T − T F
0
TF − TS
(2)
Where ␭seq (kJ/kg) is the latent heat of utility steam; hseq (kJ/kg) is
the enthalpy of utility steam; TF (K) is the flame temperature of the
boiler flue gases; TS (K) is the stack temperature, and T0 (K) is the
ambient temperature.
3. Process design and optimization
3.1. PSD sequence (water-benzene-IPA)
Fig. 3(a) gives the flowsheet of the triple-column PSD separation
sequence with the bottom product order of water-benzene-IPA. The
distillates of these columns are close to the composition of ternary
azeotropes at the individual column pressure. Fig. 3(b) illustrates
the change of distillation regions and boundaries with the variation
of pressure and overall mass-balance lines for the whole separation
process. The straight line in the region II, B1-F-D3-D1 denotes the
split of the column C-1 at 0.7 atm. The straight line in region III, B2D1-D2 denotes the split of the column C-2 at 6.5 atm. The straight
line in region I, B3-D2-D3 denotes the split of the column C-3 at
5 atm.
In order to achieve the minimum TAC of this process, twelve
deign variables need to be optimized including the number of
stages (NT1 , NT2 and NT3 ), reflux ratios (RR1, RR2 and RR3), feed
locations (NR , NF1 , NF2 , and NF3 ) of three columns, and operating
pressures (P2 and P3 ) of the columns C-2 and C-3. The bottom
flowrates (B1, B2 and B3) of three columns are adjusted to meet
the purity of the bottom products, respectively. The operating pressure of the column C-1 (P1 ) is fixed at 0.7 atm rather than higher
pressure because it can reduce the energy consumption and also
the cooling water still can be used for the high enough reflux temperature (330.70 K). Fig. 4 shows the relationships of TAC and each
optimization variable listed.
As shown in Fig. 5, the optimal partial heat-integrated PSD
sequence (water-benzene-IPA) can reduce great energy consumption (1108.15 kW) by installing a condenser/reboiler with
the transfer heat temperature difference (32.68 K) between the
columns C-1 and C-2. Most of the overhead vapor of the column
C-2 still needs the cooling water (303.15 K) to condense.
3.2. PSD sequence (water-IPA-benzene)
Fig. 2. The sequential iterative optimization procedure.
Fig. 6 (a) gives the flowsheet of the triple-column PSD separation
sequence with the bottom product order of water-IPA-benzene.
The distillates of these columns are close to the composition of
ternary azeotropes at the individual column pressure. Fig. 6(b) illustrates the change of distillation regions and boundaries with the
variation of pressure and overall mass-balance lines for the whole
4
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Table 2
Basis of economic evaluation.
Parameters
Formulas or data
Units
Condenser:
Heat transfer coefficient (UC )
Heat transfer area (AC )
Capital cost
0.852
QC /(UC ×T)
(M&S/280) ×3489.996×AC 0.65
kW/(K·m2 )
m2
$
Reboiler:
Heat transfer coefficient (UR )
Heat transfer area (AR )
Capital cost
0.568
QR /(UR ×T)
(M&S/280)×3489.996×AR 0.65
kW/(K·m2 )
m2
$
Column vessel:
H
TCinst
CSCinst
Capital cost
Energy cost:
1.2 × 0.6096×(NT -2)
(M&S/280)×262.556×ID1.55 ×H
(M&S/280)×5485.171×ID1.066 ×H0.802
TCinst +CSCins
m
$
$
$
LP steam (433 K)
MP steam (457 K)
Cooling water
M&S
Payback period
TAC
7.72
8.22
0.354
1431.7
3
(Total capital cost/payback period) +Annual energy cost
$/GJ
$/GJ
$/GJ
—
y
$/y
Fig. 3. The optimal flowsheet with detailed information (a) and overall material balance lines (b) for PSD sequence (water-benzene-IPA).
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Fig. 4. The relationships of TAC and each optimization variable for PSD sequence (water-benzene-IPA): (a) RR1, (b) RR2, (c) RR3, (d) NT1 , (e) NT2 , (f) NT3 , (g) P2 and (h) P3 .
5
6
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Fig. 5. The optimal partial heat-integrated PSD sequence (water-benzene-IPA).
Fig. 6. The optimal flowsheet with detailed information (a) and overall material balance lines (b) for PSD sequence (water-IPA-benzene).
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
7
Fig. 7. The relationships of TAC and each optimization variable for PSD sequence (water-IPA-benzene): (a) RR1, (b) RR2, (c) RR3, (d) NT1 , (e) NT2 , (f) NT3 , (g) P2 and (h) P3 .
process. The material line of the column C-1 still is the straight line
in the region II, B1-F-D3-D1. The straight line in region I, B2-D1-D2
denotes the split of the column C-2 at 3 atm. The straight line in
region III, B3-D2-D3 denotes the split of the column C-3 at 7 atm.
The optimization variables for this sequence (water-IPAbenzene) are identical with that of the sequence (water-benzeneIPA). Here no more detailed description. Fig. 7 shows the
relationships of TAC and each optimization variable.
8
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Fig. 8. The optimal partial heat-integrated PSD sequence (water-IPA-benzene).
As shown in Fig. 8, the optimal partial heat-integrated PSD
sequence (water-IPA-benzene) can reduce great energy consumption (1040.86 kW) by installing a condenser/reboiler with
the transfer heat temperature difference (36.20 K) between the
columns C-1 and C-3. Most of the overhead vapor of the column
C-3 still needs the cooling water (303.15 K) to condense.
and organic phase. The combination of the column C-3 and the
decanter is the typical HAD arrangement, which is used to remove
the benzene product. The atmospheric column C-1 and the high
pressure column C-2 constitute the conventional PSD configuration, which is used to achieve the separation of the water/IPA
azeotrope. Fig. 9(b) illustrates the distillation regions and boundaries at the different pressures as well as heterogeneous region and
overall mass-balance lines for the whole process. The straight line
in the region II, B1-F1-D1 denotes the split of the column C-1 at
1 atm. The straight line in region I, B2-D1-D2 denotes the split of
the column C-2 at 5 atm. The straight line in region III, B3-F3-D3
denotes the split of the column C-3 at 1 atm. The straight line, D3-
3.3. HAD and PSD combined sequence
Fig. 9 (a) gives the flowsheet of the HAD and PSD combined
sequence with three columns and a decanter. The whole separation process begins with the decanter to obtain the aqueous phase
Table 3
Summary of detailed optimization parameters of three separation configurations proposed.
Parameters
Pressure (atm)
NT
NR
NF
ID
RR
Condenser duty(MW)
Reboiler duty(MW)
Capital cost (105 $)
Energy cost (105 $/y)
TAC(105 $/y)
Total TAC(105 $/y)
[CO2 ]Emiss (103 kg/h)
Total [CO2 ]Emiss (103 kg/h)
PSD (water-benzene-IPA)
PSD(water-IPA-benzene)
HAD and PSD
C-1
C-2
C-3
C-1
C-2
C-3
C-1
C-2
C-3
0.7
12
2
3
0.76
0.04
1.25
1.11
6.75
2.59
4.84
15.60
0.89
3.86
6.5
28
–
19
0.90
0.4
1.25
1.66
8.55
4.05
6.90
5
34
–
8
0.56
0.01
0.85
0.80
6.01
1.86
3.86
3
31
–
6
0.78
0.01
1.06
1.32
7.60
3.22
5.76
7
15
–
7
0.79
0.3
1.13
1.32
7.05
3.06
5.41
1
17
–
10
0.73
0.3
0.59
0.68
4.89
1.58
3.21
1.64
1.06
1.06
1
35
–
31
1.28
1.2
3.51
3.90
15.77
9.04
14.29
32.19
3.14
6.72
5
49
–
16
1.18
1.6
3.42
3.77
16.25
9.27
14.69
1.33
0.7
8
2
4
0.69
0.01
1.30
1.04
6.49
2.44
4.61
15.78
0.84
2.96
3.03
0.55
Table 4
The head-to-head comparison of three partial heat-integrated separation configurations.
Parameters
PSD (water-benzene-IPA)
PSD(water-IPA-benzene)
HAD and PSD
Condenser duty(MW)
Reboiler duty(MW)
Auxiliary condenser(MW)
Heat integration duty(MW)
Capital cost (105 $)
Energy cost (105 $/y)
Total TAC(105 $/y)
Energy cost saving (%)
Capital cost saving (%)
TAC saving (%)
[CO2 ]Emiss (103 kg/h)
3.35
3.57
0.14
1.11
18.52
5.93
12.10
30.24
13.09
22.44
1.97
3.49
3.68
0.09
1.04
18.37
6.31
12.44
27.64
13.10
21.17
2.12
7.52
8.35
2.73
0.68
35.90
18.30
30.27
7.99
2.74
5.96
6.17
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
9
Fig. 9. The optimal flowsheet with detailed information (a) and overall material balance lines (b) for HAD and PSD combined sequence.
FM-D2 denotes the mixing process of the distillates D3 and D2. The
tie line in the heterogeneous region, F1-F0-F3 denotes the split of
the decanter at 308.15 K.
As for the optimization of this process, there are ten optimization variables including the number of stages (NT1 , NT2 and NT3 ),
reflux ratios (RR1, RR2 and RR3), feed locations (NF1 , NF2 and NF3 )
of three columns, and the operating pressure (P3 ) of the columns
C-3. The bottom flowrates (B1, B2 and B3) of three columns are
adjusted to meet the purity of the bottom product, respectively.
The operating pressures of the columns C-1 and C-3 (P1 and P3 ) are
both fixed at 1 atm. Fig. 10 shows the relationships of TAC and each
optimization variable.
Fig. 11 shows the HAD and PSD combined sequence with
the partial heat-integration between the columns C-2 and C-3.
The reduction of energy consumption equals to the heat duty
(682.51 kW) of the condenser/reboiler with the transfer heat temperature difference (38.38 K). Most of the overhead vapor of the
column C-2 still needs the cooling water (303.15 K) to condense.
10
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Fig. 10. The effect of optimization variables on TAC for HAD and PSD combined sequence: (a) RR1, (b) RR2, (c) RR3, (d) NT1 , (e) NT2 , (f) NT3 and (g) P2 .
The detailed optimization parameters of the above-proposed
three configurations are summarized in Table 3. Table 4 shows a
head-to-head comparison of three partial heat-integrated separation configurations.
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
11
Fig. 11. Partial heat-integration process for HAD and PSD combined sequence.
Both PSD configurations are superior to the HAD and PSD combined sequence in terms of TAC and energy consumptions, and are
acceptable alternatives because of approximately the same TAC.
With regards to heat integration, the PSD (water-benzene-IPA)
sequence with partial heat integration can achieve more energy
saving and CO2 emission than other two schemes. Note that only
the partial heat integration with the heat transfer temperature difference greater than 25 K has been accomplished in this work.
4. Conclusions
This article addressed the recovery of organic solvents benzene and IPA from wastewater via three separation configurations
without adding any extra solvent, including the PSD (waterbenzene-IPA), the PSD (water- IPA-benzene) and the HAD and
PSD combined sequence. The optimal operating conditions and the
minimum TAC were determined by using the sequential iterative
procedure. Furthermore, the partial heat-integration was carried
out for improving energy efficiency and reducing CO2 emission.
These studies revealed that the PSD process with the product
order water-benzene-IPA can high effectively solve the treatment
of wastewater containing benzene and IPA organic solvents.
References
Ahmad, T., Guria, C., Mandal, A., 2018. Synthesis, characterization and performance
studies of mixed-matrixpoly (vinyl chloride)-bentonite ultrafiltration membrane for thetreatment of saline oily wastewater. Process Saf. Environ. Prot. 116,
703–717.
Arifin, S., Chien, I.-L., 2007. Combined preconcentrator/recovery column design for
isopropyl alcohol dehydration process. Ind. Eng. Chem. Res. 46, 2535–2543.
Aurangzeb, M., Jana, A.K., 2018. Pressure-swing dividing wall column with multiple
binary azeotropes: improving energy efficiency and cost savings through vapor
recompression. Ind. Eng. Chem. Res. 57, 4019–4032.
Cao, Y., Hu, J., Jia, H., Bu, G., Zhu, Z., Wang, Y., 2017. Comparison of pressure-swing distillation and extractive distillation with varied-diameter column in economics
and dynamic control. J. Process Contr. 49, 9–25.
Chaniago, Y.D., Harvianto, G.R., Bahadori, A., Lee, M., 2016. Enhanced recovery
of PGME and PGMEA from waste photoresistor thinners by heterogeneous
azeotropic dividing-wall column. Process Saf. Environ. Prot. 103, 413–423.
Chen, Y., Liu, C., Geng, Z., 2018. Design and control of fully heat-integrated pressure
swing distillation with a side withdrawal for separating the methanol/methyl
acetate/acetaldehyde ternary mixture. Chem. Eng. Processs. 123, 233–248.
Fissore, D., Pin, M., Barresi, A.A., 2006. On the use of detailed models in the MPC
algorithm: The pressure-swing distillation case. AIChE J. 52, 3491–3500.
Gadalla, M., Olujić, Ž., de Rijke, A., Jansensb, P.J., 2006. Reducing CO2 emissions of
internally heat-integrated distillation columns for separation of close boiling
mixtures. Energy 31, 2409–2417.
Kiran, B., Jana, A.K., 2015. A hybrid heat integration scheme for bioethanol separation
through pressure-swing distillation route. Sep. Purif. Technol. 142, 307–315.
Knapp, J.P., Doherty, M.F., 1992. A new pressure-swing-distillation process for separating homogeneous azeotropic mixtures. Ind. Eng. Chem. Res. 31, 346–357.
Le, Q.-K., Halvorsen, I.J., Pajalic, O., Skogestad, S., 2015. Dividing wall columns for
heterogeneous azeotropic distillation. Chem. Eng. Res. Des. 99, 111–119.
Lee, J., Cho, J., Kim, D.M., Park, S., 2011. Separation of tetrahydrofuran and water
using pressure swing distillation: modeling and optimization. Korean J. Chem.
Eng. 28, 591–596.
Li, Y., Xu, C., 2017. Pressure-swing distillation for separating pressure-insensitive
minimum boiling azeotrope methanol/toluene via introducing a light entrainer:
design and control. Ind. Eng. Chem. Res. 56, 4017–4037.
Liang, S., Cao, Y., Liu, X., Li, X., Zhao, Y., Wang, Y., Wang, Y., 2017. Insight into pressureswing distillation from azeotropic phenomenon to dynamic control. Chem. Eng.
Res. Des. 117, 318–335.
Lladosa, E., Montón, J.B., Burguet, M., 2011. Separation of di-n-propyl ether and
n-propyl alcohol by extractive distillation and pressure-swing distillation:
computer simulation and economic optimization. Chem. Eng. Processs. 50,
1266–1274.
Luyben, W.L., 2012. Pressure-swing distillation for minimum- and maximumboiling homogeneous azeotropes. Ind. Eng. Chem. Res. 51, 10881–10886.
Luyben, W.L., 2014. Control of a heat-integrated pressure-swing distillation process
for the separation of a maximum-boiling azeotrope. Ind. Eng. Chem. Res. 53,
18042–18053.
12
Y. Cui et al. / Process Safety and Environmental Protection 122 (2019) 1–12
Luyben, W.L., 2017. Control of a triple-column pressure-swing distillation process.
Sep. Purif. Technol. 174, 232–244.
Luyben, W.L., 2018. Design and control of a pressure-swing distillation process with
vapor recompression. Chem. Eng. Processs. 123, 174–184.
Modla, G., Lang, P., Denes, F., 2010. Feasibility of separation of ternary mixtures by
pressure swing batch distillation. Chem. Eng. Sci. 65, 870–881.
Modla, G., 2013. Energy saving methods for the separation of a minimum boiling
point azeotrope using an intermediate entrainer. Energy 50, 103–109.
Mukherjee, D., Bhattacharya, P., Jana, A., Bhattacharya, S., Sarkar, S., Ghosh, S.,
Majumdar, S., Nava, J.A.O., Krishna, R., 2003. Influence of unequal component
efficiencies on trajectories during distillation of a homogeneous azeotropic mixture. Chem. Eng. Processs. 43, 305–316.
Nava, J.A.O., Krishna, R., 2004. Influence of unequal component efficiencies on trajectories during distillation of a homogeneous azeotropic mixture. Chem. Eng.
Processs. 43, 305–316.
Qasim, F., Shin, J.S., Cho, S.J., Park, S.J., 2015. Optimizations and heat integrations on
the separation of toluene and 1-butanol azeotropic mixture by pressure swing
distillation. Sep. Sci. Technol. 51, 316–326.
Tabari, A., Ahmad, A., 2015. A semicontinuous approach for heterogeneous
azeotropic distillation processes. Comput. Chem. Eng. 73, 183–190.
Wang, C., Guang, C., Cui, Y., Wang, C., Zhang, Z., 2018a. Compared novel thermally
coupled extractive distillation sequences for separating multi-azeotropic mixture of acetonitrile/benzene/methanol. Chem. Eng. Res. Des. 136, 513–528.
Wang, C., Wang, C., Cui, Y., Guang, C., Zhang, Z., 2018b. Economics and controllability
of conventional and intensified extractive distillation configurations for Acetonitrile/Methanol/Benzene mixtures. Ind. Eng. Chem. Res. 57, 10551–10563.
Wang, C., Wang, C., Guang, C., Gao, J., Zhang, Z., 2018c. Hybrid reactive distillation
using polyoctylmethylsiloxane membrane for isopentyl acetate production from
mixed PVA by products. J. Chem. Technol. Biotechnol., http://dx.doi.org/10.1002/
jctb.5799.
Wang, C., Wang, C., Guang, C., Zhang, Z., 2018d. Comparison of extractive distillation separation sequences for acetonitrile/methanol/benzene multi-azeotropic
mixtures. J. Chem. Technol. Biotechnol., http://dx.doi.org/10.1002/jctb.5693.
Wang, S.J., Huang, K., 2012. Design and control of acetic acid dehydration system
via heterogeneous azeotropic distillation using p-xylene as an entrainer. Chem.
Eng. Processs. 60, 65–76.
Wang, Y., Zhang, Z., Zhang, H., Zhang, Q., 2015. Control of heat integrated pressureswing-distillation process for separating azeotropic mixture of tetrahydrofuran
and methanol. Ind. Eng. Chem. Res. 54, 1646–1655.
Xia, S., Wei, W., Liu, G., Dong, X., Jin, W., 2011. Pervaporation properties
of polyvinyl alcohol/ceramic composite membrane for separation of ethyl
acetate/ethanol/water ternary mixtures. Korean J. Chem. Eng. 29, 228–234.
Zhang, Q., Liu, M., Li, C., Zeng, A., 2017. Heat-integrated pressure-swing distillation process for separating the minimum-boiling azeotrope ethyl-acetate and
ethanol. Sep. Purif. Technol. 189, 310–334.
Zhang, Z., Zhang, Q., Li, G., Liu, M., Gao, J., 2016. Design and control of methyl acetatemethanol separation via heat-integrated pressure-swing distillation. Chin. J.
Chem. Eng. 24, 1584–1599.
Zhao, L., Lyu, X., Wang, W., Shan, J., Qiu, T., 2017. Comparison of heterogeneous
azeotropic distillation and extractive distillation methods for ternary azeotrope
ethanol/toluene/water separation. Comput. Chem. Eng. 100, 27–37.
Zhu, Z., Wang, L., Ma, Y., Wang, W., Wang, Y., 2015. Separating an azeotropic mixture
of toluene and ethanol via heat integration pressure swing distillation. Comput.
Chem. Eng. 76, 137–149.
Zhu, Z., Xu, D., Liu, X., Zhang, Z., Wang, Y., 2016. Separation of acetonitrile/methanol/benzene ternary azeotrope via triple column pressure-swing
distillation. Sep. Purif. Technol. 169, 66–77.