You are on page 1of 14

Journal of Membrane Science 595 (2020) 117470

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Non-ideal modelling of polymeric hollow-fibre membrane systems: Pre- T


combustion CO2 capture case study
Ehsan Soroodan Miandoab, Sandra E. Kentish, Colin A. Scholes∗
Peter Cook Centre for Carbon Capture and Storage Research, Department of Chemical Engineering, The University of Melbourne, Australia

ARTICLE INFO ABSTRACT

Keywords: Some membrane gas separation applications operate at high temperature and pressure. However, the majority of
Membrane membrane gas separation models employ simplifying assumptions which are not realistic under these conditions.
Modelling In this study, a rigorous model is developed for polymeric hollow-fibre membrane modules incorporating non-
Pre-combustion CO2 capture isothermal separation (the Joule-Thomson effect), real gas behavior and concentration polarization. The model
Joule-thomson effect
also accounts for temperature-dependent permeability, friction-based pressure loss on both feed and permeate
Real gas behavior
sides and variable physical and transport properties. The rigorous model is applied for pre-combustion CO2
capture, i.e. CO2/H2 separation, and compared with a simplistic model for various polymeric membranes
through changing temperature-independent activation energy of permeation and pre-exponential factor. Two
types of H2- and CO2-selective membranes are then chosen for further analysis. As feed conditions change, the
deviation between the rigorous and simplistic models ranges approximately from 2 to 12% for stage-cut and
2–20% for the permeate composition. The difference is mostly because of real gas behavior at low stage-cuts,
while the Joule-Thomson effect adds to this behavior at high stage-cuts ( 40%) resulting in the greater devia-
tion. The influence of concentration polarization, however, is found negligible even at high stage-cuts.

1. Introduction the partial pressure of components (i.e. ideal gas behavior) in upstream
and downstream is considered the driving force for permeation [7].
Polymeric membrane gas separation is an attractive technology for Given that chemical potential is a complex function of temperature and
many industrial applications; because the process is conceptually fugacity, a rigorous approach should be employed to better describe
straightforward and energy-efficient, using a semi-permeable polymeric permeation through membranes. For most industrial applications, the
membrane that enables some gases to pass through easily, while other systems under consideration are complex, given the high-pressure feed
gases experience a barrier [1,2]. This provides greater versatility in streams, mixture of gases and vapours, as well as, significant pressure
separation applications compared to other technologies. Membrane gas drop through the membrane. Robust membrane models are therefore
separation is commercialised in natural gas sweetening and air en- needed to characterise membrane processes under high pressure and
richment [3,4], and research efforts are focused on applying mem- temperature, to better assist in the development of membrane gas se-
branes to other gas separation applications. One area in which mem- paration technology and ensure accurate process simulations are un-
brane gas separation has strong potential is CO2 capture; a strategy dertaken. Currently, membrane models are severely limited for in-
proposed to reduce CO2 emissions from large point sources. Pre-com- dustry, as the majority of commercial process simulators do not include
bustion CO2 capture in particular holds promise. In this case, an in- membrane process unit operations in their software suites, or the
tegrated gasification combined cycle (IGCC) process generates syngas membrane simulation is relatively basic. This is also the case in the
(mainly CO and H2) through partial oxidation, which is further con- academic literature, where membrane models are often simplistic
verted through the water-gas shift reaction into CO2 and H2. The CO2 is [8–10]. For these models, the permeance of components is considered
separated from H2 to yield a low-carbon high heating value fuel in constant, whereas it is inherently dependent on temperature, pressure
combined cycle plants for power generation [5]. and concentration of individual components [11]. The conventional
The driving force for membrane gas separation is the chemical po- models often use simple correlations for required physical and transport
tential difference across the polymeric membrane [6]. However, in the properties and neglect the variations of the properties as a result of
majority of the membrane gas separation literature, difference between change in concentration, temperature and pressure [12].


Corresponding author.
E-mail address: cascho@unimelb.edu.au (C.A. Scholes).

https://doi.org/10.1016/j.memsci.2019.117470
Received 3 May 2019; Received in revised form 12 September 2019; Accepted 13 September 2019
Available online 13 September 2019
0376-7388/ © 2019 Elsevier B.V. All rights reserved.
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Pre-combustion CO2 capture represents a good model system to 2.2. Mathematical model
investigate more complex behavior in membrane processes through
simulations, since in this CO2 capture scenario, high temperature and Assumptions governing the proposed model are as follow:
pressure gas is fed to the CO2/H2 separation unit. Xu et al. [13] con-
ducted a parametric analysis of pre-combustion CO2 capture using 1. The membrane module operates at steady-state.
single and double stage membrane separation but the model assumed 2. Laminar plug flow occurs on both the shell and lumen sides of
isothermal operation and ignored both concentration polarization and hollow fibres. Depending upon the feed flow rate being studied, the
real gas behavior. Similar models which neglect some or all of these Reynolds numbers for the current simulation are less than 300 on
effects have been also developed by Choi et al. [14] and Giordano et al. both the retentate and permeate sides. By this simplification, the
[15]. Franz and Scherer developed a more detailed Aspen model to momentum balance is reduced to the Hagen-Poiseuille equation.
evaluate the separation and energy efficiency loss of an IGCC power This equation is then used to evaluate the pressure drop in each cell.
plant but the model still assumed constant permeance of components Consequently, the pressure profile of the retentate and permeate
[16]. sides can be determined along the membrane module.
A number of robust mathematical models capable of simulating 3. For the current simulation, the permeability of components is as-
membrane operations have been presented in the literature. Scholz sumed independent of pressure and concentration. Permeability and
et al. [17] provided a detailed membrane gas separation model that selectivity are then assumed temperature-dependent. However, at a
included effects such as real gas behavior, non-isothermal operation later stage the proposed model can be easily modified to account for
(the Joule-Thomson effect) and concentration polarization. They pressure and concentration dependent permeabilities.
showed these non-ideal effects adversely influence membrane gas se- 4. The concentration of components changes from the bulk flow to the
paration. However, physical and transport properties were taken as membrane surface and inside the porous support due to con-
constant at feed and permeate conditions. In reality, these properties centration polarization.
can change, as temperature, pressure and concentration of the gas vary 5. The thickness of the dense layer, as well as, the inner and outer
along the membrane module. diameter of fibres remain uniform under high-pressure operation.
Here, a rigorous membrane model incorporating non-isothermal
operation, real gas behavior, concentration polarization, pressure loss In this work, the rigorous model is compared with a simplistic
along the module, temperature-dependent permeance and variable model which ignores real gas behavior, concentration polarization and
physical and transport properties is presented and applied to a mem- pressure drop due to friction and assumes constant permeability, iso-
brane process undertaking pre-combustion CO2 capture. The rigorous thermal operation and hence, constant physical and transport proper-
model is validated against experimental data presented in the literature ties along the membrane module.
and compared with a simplistic model to demonstrate the deviation in The degree of freedom (DOF) for the rigorous model in terms of the
performance due to individual effects of real gas behavior, non-iso- number of variables that must be fixed is given by:
thermal operation and concentration polarization.
DOF = 6 + 2 × (number of components ) (1)
Once the model is provided with the DOF number of fixed variables, a
2. Model development simulation can be successfully solved. For example, in case of a binary
separation, it is required to fix 10 stream variables (i.e. flow rate, tem-
2.1. Methodology perature, pressure and composition) in advance to perform the simulation.
Fig. 2 represents a single cell of the membrane module operating at
The mathematical model is based on an asymmetric hollow-fibre the counter-current mode where subscripts “R” and “P” represent in-
membrane module, i.e. a dense polymeric membrane layer on a porous formation on the retentate and permeate sides, respectively. Super-
support. To build the rigorous model, as illustrated in Fig. 1, the scripts “in” and “out” delineate inlet and outlet variables of the single
membrane module is discretized into N cells for each of which the mass, cell, respectively. In the current simulation, variables F, T, P, x and y
energy and momentum balance equations are simultaneously solved stand for flow rate, temperature, pressure, mole fraction on the re-
with the membrane transport equation. tentate side and mole fraction on the permeate side, respectively.
The model can simulate co-current and counter-current operations, The overall mole balance for the single cell is described as:
shell-side or lumen-side feed and constant or variable gas perme-
FRin + FPin = FRout + FPout (2)
abilities, depending upon the nature of membrane and process being
studied. The model also enables a sweep gas stream to be added, if it is Likewise, the mole balance of an individual component i is given by:
assumed impermeable to the membrane. This last assumption can be
FRin xiin + FPin yiin out out out out
, P = FR x i, R + FP yi, P (3)
justified if the feed gas is at high pressure and the sweep gas at ambient ,R

pressure, as the partial pressure of the sweep gas that will permeate is a
trivial fraction of the feed.

Fig. 1. Membrane module consisted of N cells (counter-current operation is


shown here). Fig. 2. A single cell of the membrane module.

2
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

for which: where variables and parameters with superscript “S” are evaluated in
C the porous support of the hollow-fibre membranes, Ji and Jt are trans-
xiout
,R = 1
membrane flux of component i and total transmembrane flux of all
i=1 (4) components, respectively, ρt is total molar density of components, k is
C
local mass transfer coefficient, D is diffusivity and ε is the porosity of
yiout the porous support. To obtain the mole fraction of components at the
,P = 1
i=1 (5) membrane surface and in the porous support, mass transfer coefficients
should be calculated by applying appropriate empirical correlations for
where i = 1, 2, 3, …, C represents the index of components and C is the
Sherwood numbers. For laminar flow regime with the Reynolds number
number of streams components. To accomplish a mole balance, an
(Re) less than 2100, the Sherwood numbers (Sh) of the retentate and
appropriate equation is needed to describe the transmembrane flow
permeate streams are evaluated by [21]:
rate. With a conservative approach, the mass transport across the
polymeric membranes is obtained by:
1
ki, R dhyd dhyd 3
ShR = = 1.62 ReR ScR
i Di, R l (15)
FPout yiout FPin yiin
,P = A ( M x M PRout M M out
i, P yi PP )
,P M c i, R i (6)
1
ki, P di d 3
where: ShP = = 1.62 ReP ScP i
Di, P l (16)
d o Nf l
Ac = where di is fibre internal diameter. The hydraulic diameter of the
N (7)
membrane module (dhyd) is given by [19,22,23]:
Here, parameters and variables taking superscript “M” refer to the
dense layer, is permeability, δ is thickness, Ac is the active membrane 1
dhyd = do
(17)
area in a single cell, do is fibre external diameter, Nf is the number of
hollow-fibre membranes, l is active fibre length, N is the number of cells for which the packing fraction of the module, Ψ, is determined by
and, is the fugacity coefficient (the fugacity coefficients and mole [19,22]:
fractions on the right-hand side of Equation (6) are evaluated at the 2
surfaces of the dense layer facing the retentate and permeate sides). For do
= Nf
high-pressure applications, such as pre-combustion CO2 capture, de- dm (18)
viation from the ideal gas behavior may lead to inaccuracy in the si-
where dm is the module diameter. To be consistent with the variables
mulation outcomes. Equation (6) considers the fugacity difference of
used for the present simulation, the Reynolds numbers of the retentate
components in the retentate and permeate sides as the driving force of
and permeate streams are estimated by [23]:
permeation [18]. The fugacity coefficients in Equation (6) were calcu-
lated using the composition of components at the membrane surface, as 4FRout MWR
ReR =
well as, temperature and pressure of the retentate and permeate flow in Nf do µR (19)
each cell. The Soave-Redlich-Kwong (SRK) equation of state (EOS) was
used to calculate the fugacity coefficients [19]. 4FPout MWP
ReP =
Because of concentration polarization, the permeation of the more Nf di µP (20)
permeable component is retarded, while it is accelerated for the less
in which MW and μ delineate total molecular weight of components and
permeable component [20]. If the feed stream enters the shell side and
viscosity, respectively. Additionally, Schmidt numbers (Sc) of the re-
the dense membrane layer is located at the external surfaces of hollow-
tentate and permeate sides are calculated by [24]:
fibres, the governing equations considering concentration polarization
are provided by [21]: µR
ScR =
t , R Di, R (21)
x iout
,R xi Jt
= exp µP
xiM xi t , R k i, R (8) ScP =
t , P Di, P (22)
yiM xi Jt S
= exp Friction forces inside fibres and on shell, create pressure drop (ΔP)
yiS xi t
S
DiS (9) leading to the decrease in the transmembrane driving force of per-
meation along the membrane module. Assuming laminar plug flow for
yiS xi Jt both sides, the pressure drop of the retentate and permeate sides is
= exp
yiout
,P xi t , P k i, P (10) given by [8]:
128µR l
where: PR = FRout
2 2
t , R dhyd (dm Nf do2) N (23)
i M M out M M out
J M ( i, R x i PR i, P yi PP )
xi = i = C 128µP l
Jt i
( M x M PRout
i = 1 M i, R i
M M out
i, P yi PP ) (11) PP = 4
FPout
t , P di Nf N (24)
C
For the polymeric membranes, if permeation is temperature-de-
xiM = 1
i (12) pendent, the permeability of components is usually described by an
Arrhenius type equation [25]:
C
yiM = 1 Ep, i
(13) i = i,0 exp
i RTRout (25)
C
where i,0 and Ep,i are temperature-independent pre-exponential factor
xiS = 1
i
and activation energy of permeation for component i, respectively, and
(14) R is the ideal gas law constant (8.314 J/mol.K). To provide an insight

3
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

into the composition of the retentate and permeate streams, membrane where h and λ are local convective heat transfer coefficient and thermal
selectivity should be taken into consideration [2]. For a binary mixture conductivity, respectively. Here, feed is assumed on the shell side. Assuming
of components i and j, temperature-dependent selectivity of i with re- laminar flow within the shell and in the lumen, convective heat transfer
spect to j is defined as [6,26]: coefficients of the retentate and permeate sides are determined by [19]:

= i
=
i,0 exp ( Ep, i
RTRout )= 0
exp
E
hP = 3.66 P
di (34)
exp ( )
i/j i/j
j
Ep, j RTRout
j,0 RTRout (26) M)
1
R ReR PrR (d o + 2 3
hR = 3.66 + 1.077
in which: do + 2 M 2dhyd
(35)
0
0
i/j = i
0 where the Prandtl number (Pr) of the retentate side is defined as:
j (27)
cR µR
PrR =
E = Ep, i Ep, j (28) R (36)

Selectivity is used as a criterion for determining the purity of where c delineates specific heat capacity. The effective conductivity of the
components in the permeate and retentate streams. Applying Equation porous support (λS) is estimated by the conductivities of the polymeric
(6) for i and j, and dividing the transmembrane flow rate of i by j results material forming the support (λSP) and permeate gas molecules inside the
in: pores of the support (λP) [19]:

FPout yiout
,P FPin yiin
,P
S = SP + (1 ) P (37)
FPout (1 yiout
,P ) FPin (1 yiin
,P ) The thermal conductivity of many polymers are reported in the
M M out M M out literature [28–30]. In the current simulation, the polymer thermal
0 E i, R x i PR i, P yi PP
= i/ j exp R M conductivity was fixed at 0.2 W/mK (λM,λSP) [19].
RTout j, R (1 x iM ) PRout M
j, P (1 yiM ) PPout (29) The rigorous model is coded in Aspen Custom Modeller (V9) which
In the model, appropriate values of ΔE and α are required as input. 0 provides access to the component database and thermodynamic
If the information of the previous cell (indicated by superscript “in”) packages of Aspen Properties. The present model can easily be exported
and the mole fraction of components at the membrane surface are to Aspen Plus to enable simulations of more complex processes.
known, Equation (29) is solved to give the mole fractions of compo- Fig. 3 demonstrates the methodology of modelling. Gas components and
nents in the permeate stream. Once the composition of the permeate the thermodynamic package (Equation of State, EOS) are defined in Aspen
stream is obtained, the mole fraction of components in the retentate Properties and imported to the membrane model. The geometric char-
stream can be calculated by solving the mass balance equations in each acteristics of the hollow-fibre membrane module (number of fibres, inner
cell. and outer diameter etc.) and number of cells (which was fixed at 100 in this
Enthalpy exchange along the membrane module and between the work) must also be input into the model. The membrane model is made up
retentate and permeate streams accounts for temperature variations. of two sub-models namely “cell” and “membrane transport” providing it
This is likely to happen when membrane gas separation occurs at high with a solution to the mass, momentum and energy transport equations and
pressure [27]. When this is the case, the permeability of components the permeability of components, respectively. When the simulation has
may be changing as temperature varies within the membrane module. converged, flow rate, composition, temperature and pressure, as well as,
Hence, to obtain the accurate results of simulation, an energy balance is enthalpy and density of the retentate and permeate streams will be de-
considered. The overall energy balance equation for the single cell is termined through solving transport equations, connecting the cells together
expressed as: and linking them to the feed, retentate and permeate streams.
Aspen Custom Modeller utilises an equation-oriented programming
FRin HRin + FPin HPin = FRout HRout + FPout HPout (30) language (i.e. declarative) in which mathematical equations are decom-
where H indicates enthalpy per unit mole. As the high-pressure gas posed to break down the sets of equations into groups. The groups are then
passes across the membrane into a lower pressure environment, isen- solved sequentially by choosing an appropriate solver [31]. For the present
thalpic expansion happens similar to that which occurs across a throt- simulation, Newton's method was employed. The maximum number of
tling valve. This causes a significant change of gas temperature. Such iterations was fixed at 100 with convergence achieved once the following
temperature variations are often referred to as the Joule-Thomson ef- condition is met:
fect [19,27]. As permeation continues, the permeation enthalpy is
< 10 10
transported from the retentate to the permeate side. Due to gas ex-
(38)
pansion from high to low pressure, a temperature difference between
the retentate and permeate streams may be induced, resulting in a net where represents any unknown variables used in the membrane model
rate of heat transfer (q) across the membrane material. These heat ef- and indicates the difference between the values of the unknown vari-
fects are summarized as: able in two successive iterations. Newton's method relies on an accurate
C
initial guess for unknown variables and hence, a detailed initialization
FRin HRin = FRout HRout + HRin i
A ( M x M PRout M M out procedure was applied to ensure the convergence of the simulation.
M c i, R i i, P yi PP ) +q
i=1 (31) The model also allows for easy calculation of physical and transport
properties including density, viscosity, diffusivity, thermal con-
q = UO Ac (TRout TPout ) (32) ductivity, specific heat capacity and molecular weight of the retentate
for which UO is overall heat transfer coefficient. Coker et al. [19] and permeate sides in each cell. The properties were calculated in each
conducted a theoretical analysis and obtained the overall heat transfer cell by calling built-in procedures of Aspen Custom Modeller. These
coefficient by: procedures need temperature, pressure and mole fraction as input
1
parameters. In this simulation, the temperature, pressure and mole
M M M M
UO =
1 do + 2
+
d o do + 2
ln
do d +
+ o M ln
do + 2
+
1 fraction of the retentate and permeate flow exiting the cells (i.e.
hP di 2 S do di do hR TRout , TPout , PRout , PPout , xiout , yiout ) were considered as input to these built-in
(33) procedures to calculate physical and transport properties. Hence, the

4
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 3. Methodology of modelling.

properties varied within cells. The enthalpies of the retentate and bara. Table 1 represents the geometrical characterises of the hollow-
permeate flow in each cell were also calculated in the same manner. fibre membrane module used for this case study.

3. Results and discussion 3.2.1. Comparison of the rigorous and simplistic models: effect of membrane
selectivity
3.1. Model validation The parameters influencing selectivity are the temperature-in-
dependent activation energy of permeation and pre-exponential factors.
The proposed membrane model is validated with experimental data The case study varied the selectivity parameters to compare the rigorous
published by Pan [32], Sidhoum [33] and Feng et al. [34]. The per- and simplistic models at 50% stage-cut in terms of the compositions of the
formance of the membrane modules is represented by the dimensionless retentate and permeate streams (Fig. 4). The parameters of the tempera-
parameter stage-cut which is the ratio of feed flow rate to permeate ture-dependent selectivity are provided for selected polymers in Table 2.
flow rate. The activation energies of permeation have not been reported While constant H2/CO2 selectivity at the feed temperature (100 °C) is
for the systems described by Pan [32] and Sidhoum [33], and hence, used for the simplistic model, this changes along the membrane module in
constant permeabilities have been used in these cases. Temperature- the rigorous model due to temperature variations calculated from the
dependent permeabilities are used for the air separation system re- energy balance. Compared to the simplistic model, the rigorous model
ported by Feng et al. [34]. predicts a retentate stream that is 2–5% lower in CO2 concentration and a
The details of the membrane gas separation systems and the full model permeate stream which has 0.5–8% more CO2 at 50% stage-cut (Fig. 4).
fitted to the experimental data are provided in Table S1 and Figs. S1–S4 of Conversely, at 10% stage cut, the difference between the two models is
the Supplementary Information. In the current simulation, the number of less than 1% and 5–12% for the composition of the retentate and permeate
cells (N) was fixed at 100, as selecting values greater than 100 did not streams respectively (see Fig. S5 of the Supplementary Information).
affect the simulation results. This is shown in Fig. S2 of the Supplementary As ΔE increases, the H2/CO2 selectivity decreases (Equation (26))
Information along with the model fitted to the experimental data of Sid- and the retentate and permeate mole fractions of CO2 obtained by both
houm. As N is doubled in value, there is no significant difference between models tend to converge. This behavior is independent of the values of
the model predictions with N = 100 and N = 200. The residual error α0 implying that the separation performance of polymers having large
between the model calculations and the experimental data of Sidhoum and ΔE, i.e. low H2/CO2 selectivity, can reasonably be simulated by the
Feng et al. was less than 1%, while it was less than 3% for the work of Pan. simplistic model. Conversely, the use of the simplistic model for the
This relatively larger error is within the experimental error (3.5%) re-
ported by Pan. The membrane model successfully simulated the different Table 1
Characteristics of the hollow-fibre membrane module [19].
sets of experimental data highlighting the rigorous approach of the model
to be applied for multi-component systems at both high or low feed Feed Shell-side
pressures, as well as, temperature-dependent permeation case.
Fibre internal diameter 150 μm
Fibre external diameter 300 μm
3.2. Pre-combustion CO2 capture simulations Number of fibres 500000
Fibre length 0.8
Thickness of separative layer 0.5 μm
Having validated the model, simulations of CO2/H2 separation were
Porosity of the porous support 0.5
conducted. These simulations assumed a base case of a feed gas of 60% Module diameter 0.3048 m
CO2 and 40% H2 at 100 °C and 60 bara, with a permeate pressure of 1.5

5
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 4. Mole fraction of CO2 in the retentate and permeate streams obtained by the rigorous (solid line) and simplistic (dash line) models as a function of the
difference in CO2 and H2 activation energies of permeation and pre-exponential factor. Feed is 60% CO2 at 60 bara and 100 °C. Stage-cut is fixed at 50%.

membranes with high H2/CO2 selectivity may make errors in simula- Increased feed temperature results in an increased permeate flow
tion. Polymers having a small gap between the activation energies of rate for both membranes, and this leads to a greater stage-cut at a fixed
the components can be good candidates for CO2/H2 separation. Refer- feed flow rate. Similarly, the membrane selectivity decreases with
ring to Table 2, this suggests that PBI has potential for CO2/H2 se- temperature for both membranes, reflecting the positive values of ΔE
paration, as the difference in the activation energies of CO2 and H2 is used in Equation (29). This results in greater permeate purities at lower
least among similar polymers. More CO2 is recovered from the retentate temperature in both cases (Fig. 6). However, achieving higher permeate
side and less CO2 is withdrawn from the permeate side as the pre-ex- purities comes at the expense of lower purity of the less permeable
ponential factor increases from 5 to 45 (Fig. 4). A high value is thus component in the retentate stream. The deviation between the sim-
desirable, although this parameter has less impact across the range of plistic and rigorous models developed here, is also shown in Figs. 5 and
values presented in Table 2 than the activation energy. 6. The deviation ranges from being negligible to 20% depending upon
the process parameters. The differences are generally greater at lower
feed temperatures, as the Joule-Thomson effect escalates when tem-
3.2.2. Comparison of the rigorous and simplistic models: effect of feed perature is reduced (see Section 3.3).
temperature, pressure and composition The rigorous model predicts permeate flow rates slightly larger than
H2-selective ( H0 2 / CO2 = 25, ΔE = 2000 J/mol) and CO2-selective what is given by the simplistic model for the H2-selective membrane,
( CO2/ H2 = 0.2, ΔE = 12000 J/mol) membranes were next compared
0
while for the CO2-selective membrane the permeate flow rates are
( E = Ep, H2 Ep, CO2 ). The deviation between the rigorous and sim- lower. Similarly, the mole fraction of H2 is somewhat higher for the
plistic models was highlighted for these two membranes and their rigorous model than the simplistic model (Fig. 6) for the H2-selective
ability to separate CO2/H2 for various feed conditions (composition, membrane, while the reverse is true for most feed flowrates for the CO2-
temperature and pressure) was investigated. selective membrane. These changes reflect the fact that H2 has a fu-
The exiting permeate flow rate and stage-cut observed within the gacity coefficient greater than unity, while that for CO2 is less than
membrane module as a function of feed temperature and flowrate are unity (see Section 3.3 below).
provided in Fig. 5. As feed flow rate increases, the proportion per- Figs. 7 and 8 display the difference between the rigorous and sim-
meating (the stage-cut) decreases. In turn, this means that the retentate plistic models as a function of feed pressure and composition at 50%
composition remains closer to that of the original feed. This increases and 10% stage-cuts, respectively. In pre-combustion CO2 capture, the
the average driving force for permeation (the difference in composition feed concentration of CO2 can vary from 15 to 60% (dry basis) with
between feed and permeate) leading both to a greater permeate flow- total feed pressure ranging from 20 to 70 bar [40]. Here, the results are
rate (Fig. 5) and a greater permeate purity (Fig. 6).

Table 2
Pre-exponential factor ( H2 / CO2 )
0
and activation energy (ΔE) for the H2/CO2 selectivity of selected polymers.

Polymer Ep (kJ/mol) ΔE (kJ/mol) 0


H2
Reference
0 =
H2 / CO2 0
CO 2
H2 CO2 Ep, H2 Ep,CO2

Polydimethylsiloxane (PDMS) 13.4 2.2 11.2 26 [26]


Poly (ether-b-amide) (PEBAX) 12.3 2.7 9.6 26 [35]
Polyethylene oxide (PEO) 76 70 6 60 [36]
Polybenzimidazole (PBI) 10 9.2 0.8 24 [37]
24 21.9 2.1 28 [38]
Polyurethane (PU) 41 33.8 7.2 33 [39]

6
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 5. Permeate flow rate and stage-cut obtained by the rigorous (solid line) and simplistic (dash line) models as a function of feed temperature and flow rate. (a) and
(c): H2-selective ( H0 2 / CO2 =25, ΔE = 2000 J/mol), (b) and (d): CO2- selective ( CO
0
2 / H2 = 0.2, ΔE = 12000 J/mol). Feed is 60% CO2 and at 60 bara.

demonstrated for 20, 40 and 60% CO2 (H2 balance) in the feed at this difference is only 0.3% on average for the CO2-selective membrane
pressures ranging from 30 to 60 bara. The deviation between the two (Fig. 7b). As stage-cut is reduced to 10%, an average deviation of 0.8%
models increases at elevated feed pressures. This is due to the stronger is calculated for the H2-selective membrane, while an average differ-
impact of real gas behaviour and Joule-Thomson effects at higher ence of 1.2% is obtained for the CO2-selective membrane (Fig. 8b). This
pressures (see Section 3.3). As CO2 concentration increases, the devia- reflects the greater magnitude of the Joule-Thomson effect on CO2 re-
tion enlarges for the H2-selective membrane, so that it ranges from lative to H2 at high stage-cuts and the larger departure from ideal gas
0.2% (20% CO2) to 2% (60% CO2) at 50% stage-cut (Fig. 7a). However, behaviour at lower stage-cuts (see Section 3.3 below).

Fig. 6. Mole fraction of components in the permeate stream obtained by the rigorous (solid line) and simplistic models (dash line for 100 °C, dot line for 200 °C) as a
function of feed temperature and flow rate. (a): H2-selective ( H0 2 / CO2 =25, ΔE = 2000 J/mol), (b): CO2- selective ( CO
0
2 / H2 = 0.2, ΔE = 12000 J/mol). Feed is 60%
CO2 and at 60 bara.

7
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 7. Mole fraction of components in the permeate stream obtained by the rigorous (solid line) and simplistic (dash line) as a function of feed pressure and
composition. (a): H2-selective ( H0 2 / CO2 =25, ΔE = 2000 J/mol), (b): CO2- selective ( CO
0
2 / H2 = 0.2, ΔE = 12000 J/mol). Feed is at 100 °C. Stage-cut is fixed at 50%.

3.3. Individual influence of real gas behavior, concentration polarization simplistic model. Consequently, the rigorous model calculates a greater
and non-isothermal operation permeate side mole fraction of H2 than the simplistic model (Fig. 6). In
the Supplementary Information, Fig. S6 provides the H2-selective
Fig. 9 indicates the influence of real gas behavior on the H2-selective membrane with the driving force of permeation and fugacity coeffi-
membrane at 50% stage-cut, by plotting the driving force of permeation cients at 10% stage-cut where trends similar Fig. 9 are observed. For
and fugacity coefficient along the membrane module. The fugacity both stage-cuts, the fugacity coefficients of CO2 and H2 are 0.9 and 1.09
coefficient of CO2 is approximately 0.9 due to high pressure operation on average. This results in an approximately constant difference be-
(60 bara) and mildly decreases along the module on the retentate side. tween the driving force of permeation (3.4 bara for CO2 and 2.1 bara for
This, in turn, means that the driving force for permeation decreases H2). Therefore, real gas behavior exists over a wide range of stage-cuts.
compared to the ideal behavior. On the other hand, the fugacity coef- For the CO2-selective membrane, the driving force of permeation
ficient of H2 is 1.1 on the retentate side and increases along the module and fugacity coefficients of components are presented in Fig. 10 and
resulting in a greater driving force of permeation in comparison to the Fig. S7 of the Supplementary Information at 50% and 10% stage-cuts,

Fig. 8. Mole fraction of components in the permeate stream obtained by the rigorous (solid line) and simplistic (dash line) as a function of feed pressure and
composition. (a): H2-selective ( H0 2 / CO2 =25, ΔE = 2000 J/mol), (b): CO2- selective ( CO
0
2 / H2 =0.2, ΔE = 12000 J/mol). Feed is at 100 °C. Stage-cut is fixed at 10%.

8
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 9. Driving force of permeation for the H2-selective membrane ( H0 2 / CO2 =25, ΔE = 2000 J/mol). Fugacity coefficients of components in the retentate and
permeate streams along the membrane module. Feed is 60% CO2 at 60 bara and 100 °C. Stage-cut is fixed at 50%.

respectively. With 50% stage-cut, the fugacity coefficient of CO2 becomes less than unity with increasing stage-cut, as concentration
slightly increases from 0.9 to 0.92 on the retentate side (Fig. 10). This polarization increases due to the greater transmembrane flux [41].
reduces permeation flux compared to the simplistic model. Conversely, Although the trends observed in Fig. 11 confirms the existence of
the driving force of H2 increases with respect to the simplistic model. concentration polarization, the magnitude of is still close to unity for
This is because H2 shows a small positive departure from the ideal gas even the extreme cases. For example, a stage-cut of 0.9 corresponds to θ
behavior with the fugacity coefficient above unity. The results clearly of 0.9985 on average for the membrane systems studied. This confirms
demonstrate the existence of real gas behavior at different stage-cuts. that the influence of concentration polarization on the separation per-
The ratio of the mole fraction on the retentate side (x iout
, R ) to the mole formance of the membrane module will be negligible.
fraction at the membrane surface ( x iM ) is defined as parameter and The impact of the non-isothermal operation on the H2-selective
plotted versus the module length (Fig. 11). At the low stage-cut (10%), membrane is illustrated in Fig. 12. The temperature change along the
the parameter is essentially equal to unity meaning that the gas module is directly proportional to the feed pressure and stage-cut and
composition does not change from the bulk retentate flow to the increases at higher stage-cuts and elevated feed pressures. A similar
membrane surface. This is expected, as concentration polarization is trend was observed in the literature [19,42,43]. The temperature
reduced at lower stage-cuts [41]. The parameter θ decreases and change from feed to retentate is positive, meaning that the retentate

Fig. 10. Driving force of permeation for the CO2-selective membrane CO2- selective ( CO
0
2 / H2 = 0.2, ΔE = 12000 J/mol). Fugacity coefficients of components in the
retentate and permeate streams along the membrane module. Feed is 60% CO2 at 60 bara and 100 °C. Stage-cut is fixed at 50%.

9
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 11. The ratio of the mole fraction on the retentate side ( x iout
, R ) to the mole fraction at the membrane surface ( x i ) for H2 and CO2. (a) the H2-selective membrane
M

( H0 2 / CO2 =25, ΔE = 2000 J/mol) and (b) the CO2-selective ( CO 0


/
2 2 H = 0.2, ΔE = 12000 J/mol). Feed is 60% CO 2 at 60 bara and 100 °C.

temperature is higher than the feed temperature. Such a positive tem- are in a good agreement with the literature [45,46].
perature change from feed to retentate was predicted by Gorissen [27], For the H2-selective membrane, H2 is the dominant permeating
for synthesis gas (carbon monoxide (CO) and H2) separation at 100 bara component. As the Joule-Thomson coefficient of H2 is negative within
and 52% stage-cut. A qualitative explanation for the heating of the the temperature range of the current study, the overall temperature
retentate stream may be described by using the isenthalpic Joule- deviation is positive, leading to the heating of the retentate stream.
Thomson coefficient given by [44]: The temperature change along the module for the CO2-selective
membrane is provided in Fig. 13. In contrast to the H2-selective mem-
T 1 V brane, expansion driven cooling is seen for this CO2-selective mem-
µJT = = T V
P H cp T (39) brane. This is because CO2 is the dominant permeating component and
has positive Joule-Thomson coefficients within the temperature range
in which V and cp represent the molar volume and specific heat capa-
of the current study (Table 3). Compared to the H2-selective membrane,
city, respectively. The Joule-Thomson coefficients of CO2 and H2 are
the change in temperature is greater for the CO2-selective membrane, as
calculated at different temperatures (Table 3) using molar volume and
the Joule-Thomson coefficient of CO2 is much larger than that for H2.
specific heat capacity data provided by Aspen Properties. These values

Fig. 12. Temperature change from feed to retentate for the H2-selective membrane ( H2 / CO2 =25,
0
ΔE = 2000 J/mol) as a function of feed pressure. Feed is 60% CO2.

10
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Table 3 the variation in the selectivity is much less significant at 200 °C


The Joule-Thomson coefficients of CO2 and H2 at 60 bara. (Fig. 14). Non-isothermal operation clearly has more impact on the
Temperature (°C) CO2 H2 CO2-selective membrane at lower temperatures.
For the H2-selective membrane, concentration polarization does not
25 1.24 −0.031 affect the simulation results (Fig. 15). This conclusion is in agreement
50 0.98 −0.034
with the study by Mourgues and Sanchez in which they conducted a
100 0.57 −0.038
150 0.45 −0.04
theoretical analysis of concentration polarization in membrane gas se-
200 0.36 −0.042 paration and came to the conclusion that the phenomenon becomes
noticeable when the permeances of components are greater than 1000
GPU and the selectivity of more permeable to less permeable compo-
For the feed pressure and temperature of 80 bar and 100 °C, respec- nent is higher than 100 [41]. For the current study, the selectivity of
tively, there is a substantial temperature change of 39 °C for a 50% both H2- and CO2-selective membranes is below 100 (Fig. 14) and
stage-cut (feed flow rate of 3.2 kmol/h) and 58 °C for a 60% stage-cut hence, concentration polarization is negligible. This is also confirmed
(feed flow rate of 2.3 kmol/h). Such a large temperature drop was also by the discussion presented for Fig. 11. As feed flow rate increases, the
predicted by Coker et al. for CO2 separation from methane (CH4), with a rigorous model converges on the model taking only real gas behavior
temperature drop of 40 °C. This temperature drop was calculated at a into account. This means the rigorous model is essentially described by
53% stage-cut and corresponded to the feed flow rate of 12.8 kmol/h real gas behavior at lower stage-cuts. For the feed flow rates less than
[19]. The Coker study neglected real gas behavior and assumed con- 20 kmol/h, i.e. at stage-cuts higher than 45%, the rigorous model
stant physical and transport properties despite operating at elevated slightly diverges from the curve of real gas behavior and approaches to
feed pressure and high concentration of CO2 (which would normally the model for which only non-isothermal operation, i.e. Joule-Thomson
cause non-ideality in the system). effect, is considered. This is verified by Fig. 12 where the Joule-
The temperature change should affect the selectivity by changing Thomson effect is more pronounced at higher stage-cuts.
permeability (see Equation (25)) and hence, the rigorous model should Fig. 16 summarises how the individual effects of the rigorous model
diverge from the simplistic model due to these effects. Fig. 14 illustrates influence the CO2-selective membrane. The simplistic model converges
the selectivity of the H2- and CO2-selective membranes along the on a model that solely takes concentration polarization into con-
membrane module as a result of these temperature variations. How- sideration. Just like the H2-selective membrane, this implies that con-
ever, for the H2-selective membrane, there is a very small change in H2/ centration polarization has almost no impact on the separation per-
CO2 selectivity, although this variation is slightly greater at 100 °C formance of the CO2-selective membrane (see Fig. 11). The rigorous
compared to 200 °C. For the CO2-selective membrane, the CO2/H2 se- model lies between the curves of real gas behavior and non-isothermal
lectivity increases by around 10% and 40% at the feed temperatures of operation within the flow rates (i.e. stage-cuts) studied in this simula-
200 °C and 100 °C, respectively. The ratio of the Joule-Thomson coef- tion. Similar to the H2-selective membrane, the rigorous model tends
ficient of CO2 to H2 drops from 15 to 8.5 as temperature rises from towards the real gas behavior at higher feed flow rates, i.e. lower stage-
100 °C to 200 °C. This decline results in the smaller temperature change cuts. However, compared to the H2-selective membrane (Fig. 15), these
at 200 °C, as confirmed by Figs. 12 and 13. This can also be inferred curves are not completely congruent. Given that there is no significant
from Fig. 6, where the mole fraction of CO2 is larger at the lower Joule-Thomson effect at lower stage cuts (Fig. 13), this divergence
temperature (100 °C). The average Joule-Thomson coefficient of the implies that CO2 causes greater deviation from ideal gas behavior
mixed H2/CO2 gas stream will decrease at 200 °C due to the lower CO2 compared to H2. By reducing the feed flow rate, i.e. increasing stage-
content, leading to a smaller change in temperature at 200 °C compared cut, the influence of the Joule-Thomson effect is more induced (Fig. 13),
to 100 °C. Consequently, this smaller temperature change means that as the rigorous model diverges from real gas behavior and rather

Fig. 13. Temperature change between feed and retentate for the CO2-selective membrane 0
CO2 / H2 = 0.2, ΔE = 12000 J/mol) as a function of feed pressure. Feed is
60% CO2.

11
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 14. Selectivity of (a): H2-selective ( H0 2 / CO2 =25, ΔE = 2000 J/mol), (b): CO2- selective ( 0
CO2 / H2 = 0.2, ΔE = 12000 J/mol). membranes along the membrane
module. Feed is 60% CO2 at 60 bara. Stage-cut is fixed at 50%.

Fig. 15. Influence of individual non-ideal effects on the mole fraction of H2 in the permeate stream as a function of the feed flow rate for the H2-selective membrane
( H0 2 / CO2 =25, ΔE = 2000 J/mol). The dashed curve represents both the simplistic model and one including concentration polarization, as these curves are congruent.
Feed is 60% CO2 at 60 bara and 100 °C.

approaches to the curve of non-isothermal operation. This confirms that been coded in Aspen Custom Modeller enabling end-users to have ac-
it is the combination of these two effects determining the specifications cess to the massive component database and thermodynamic properties
of the retentate and permeate streams. of Aspen Properties. It is also capable of being exported to Aspen Plus,
as a process simulator. The model has been validated by several ex-
perimental data. A case study of pre-combustion CO2 capture with H2-
4. Conclusion
selective and CO2-selective membranes was chosen to demonstrate the
difference in performance between this model and more simplistic ones.
A model of single-stage membrane gas separation has been devel-
This comparison showed that concentration polarization had the least
oped for hollow-fibre modules. The model utilises non-isothermal op-
impact on the separation performance of both CO2-selective and H2-
eration with temperature-dependent permeance, concentration polar-
selective membranes for this application and at these process flowrates.
ization, real gas behavior, frictional pressure loss and variable physical
Real gas behavior was the most important effect for an H2-selective
and transport properties along the membrane module. The model has

12
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

Fig. 16. Influence of individual non-ideal effects on the mole fraction of CO2 in the permeate stream as a function of the feed flow rate for the CO2-selective
membrane ( CO 0
2 / H2 = 0.2, ΔE = 12000 J/mol). The dashed curve represents both the simplistic model and one including concentration polarization, as these curves
are congruent. Feed is 60% CO2 at 60 bara and 100 °C.

membrane, whereas a combination of the Joule-Thomson effect and [8] D.T. Coker, B.D. Freeman, G.K. Fleming, Modeling multicomponent gas separation
real gas behavior affected the separation performance of the CO2-se- using hollow-fiber membrane contactors, AIChE J. 44 (1998) 1289–1302, https://
doi.org/10.1002/aic.690440607.
lective membrane. Expansion driven heating and cooling of the re- [9] P. Cruz, J.C. Santos, F.D. Magalhães, A. Mendes, Simulation of separation processes
tentate stream were observed due to the negative and positive Joule- using finite volume method, Comput. Chem. Eng. 30 (2005) 83–98, https://doi.org/
Thomson coefficients of H2 and CO2, respectively. Overall, the non- 10.1016/j.compchemeng.2005.08.004.
[10] R. Khalilpour, A. Abbas, Z. Lai, I. Pinnau, Modeling and parametric analysis of
ideal behavior of membrane-based pre-combustion carbon capture was hollow fiber membrane system for carbon capture from multicomponent flue gas,
primarily because of the non-ideal behavior of both CO2 and H2 and the AIChE J. 58 (2012) 1550–1561, https://doi.org/10.1002/aic.12699.
large Joule-Thomson coefficient of CO2. [11] M.J. Thundyil, W.J. Koros, Mathematical modeling of gas separation permeators —
for radial crossflow, countercurrent, and cocurrent hollow fiber membrane mod-
ules, J. Membr. Sci. 125 (1997) 275–291, https://doi.org/10.1016/S0376-7388(96)
Acknowledgements 00218-9.
[12] P.K. Kundu, A. Chakma, X. Feng, Simulation of binary gas separation with asym-
metric hollow fibre membranes and case studies of air separation, Can. J. Chem.
Ehsan Soroodan Miandoab acknowledges The University of
Eng. 90 (2012) 1253–1268, https://doi.org/10.1002/cjce.20631.
Melbourne for the Melbourne Research Scholarship. [13] J. Xu, Z. Wang, C. Zhang, S. Zhao, Z. Qiao, P. Li, J. Wang, S. Wang, Parametric
analysis and potential prediction of membrane processes for hydrogen production
Appendix A. Supplementary data and pre-combustion CO2 capture, Chem. Eng. Sci. 135 (2015) 202–216, https://doi.
org/10.1016/j.ces.2015.04.033.
[14] J.H. Choi, M.-J. Park, J. Kim, Y. Ko, S.-H. Lee, I. Baek, Modelling and analysis of
Supplementary data to this article can be found online at https:// pre-combustion CO2 capture with membranes, Korean J. Chem. Eng. 30 (2013)
doi.org/10.1016/j.memsci.2019.117470. 1187–1194, https://doi.org/10.1007/s11814-013-0042-7.
[15] L. Giordano, J. Gubis, G. Bierman, F. Kapteijn, Conceptual design of membrane-
based pre-combustion CO2 capture process: role of permeance and selectivity on
References performance and costs, J. Membr. Sci. 575 (2019) 229–241, https://doi.org/10.
1016/j.memsci.2018.12.063.
[16] J. Franz, V. Scherer, An evaluation of CO2 and H2 selective polymeric membranes
[1] R.W. Baker, Gas separation, Membrane Technology and Applications, John Wiley &
for CO2 separation in IGCC processes, J. Membr. Sci. 359 (2010) 173–183, https://
Sons, Ltd, 2004, pp. 301–353, , https://doi.org/10.1002/0470020393.ch8.
doi.org/10.1016/j.memsci.2010.01.047.
[2] D.F. Sanders, Z.P. Smith, R. Guo, L.M. Robeson, J.E. McGrath, D.R. Paul,
[17] M. Scholz, T. Harlacher, T. Melin, M. Wessling, Modeling gas permeation by linking
B.D. Freeman, Energy-efficient polymeric gas separation membranes for a sustain-
nonideal effects, Ind. Eng. Chem. Res. 52 (2013) 1079–1088, https://doi.org/10.
able future: a review, Polymer 54 (2013) 4729–4761, https://doi.org/10.1016/j.
1021/ie202689m.
polymer.2013.05.075.
[18] R. Wang, S.L. Liu, T.T. Lin, T.S. Chung, Characterization of hollow fiber membranes
[3] R.W. Baker, Future directions of membrane gas separation technology, Ind. Eng.
in a permeator using binary gas mixtures, Chem. Eng. Sci. 57 (2002) 967–976,
Chem. Res. 41 (2002) 1393–1411, https://doi.org/10.1021/ie0108088.
https://doi.org/10.1016/S0009-2509(01)00435-3.
[4] C.A. Scholes, G.W. Stevens, S.E. Kentish, Membrane gas separation applications in
[19] D.T. Coker, T. Allen, B.D. Freeman, G.K. Fleming, Nonisothermal model for gas
natural gas processing, Fuel 96 (2012) 15–28, https://doi.org/10.1016/j.fuel.2011.
separation hollow-fiber membranes, AIChE J. 45 (1999) 1451–1468, https://doi.
12.074.
org/10.1002/aic.690450709.
[5] M. Kanniche, R. Gros-Bonnivard, P. Jaud, J. Valle-Marcos, J.-M. Amann,
[20] G. He, Y. Mi, P. Lock Yue, G. Chen, Theoretical study on concentration polarization
C. Bouallou, Pre-combustion, post-combustion and oxy-combustion in thermal
in gas separation membrane processes, J. Membr. Sci. 153 (1999) 243–258, https://
power plant for CO2 capture, Appl. Therm. Eng. 30 (2010) 53–62, https://doi.org/
doi.org/10.1016/S0376-7388(98)00257-9.
10.1016/j.applthermaleng.2009.05.005.
[21] T. Melin, R. Rautenbach, Membranverfahren: Grundlagen der Modul- und
[6] D.R. Paul, Y.P. Yumpolski, Polymeric gas separation membranes, taylor & francis,
Anlagenauslegung, Springer-Verlag, 2007.
n.d. http://www.taylorfrancis.com/books/e/9781351084338 , Accessed date: 19
[22] J.-M. Zheng, Z.-W. Dai, F.-S. Wong, Z.-K. Xu, Shell side mass transfer in a transverse
January 2019.
flow hollow fiber membrane contactor, J. Membr. Sci. 261 (2005) 114–120,
[7] K. Ghosal, B.D. Freeman, Gas separation using polymer membranes: an overview,
https://doi.org/10.1016/j.memsci.2005.02.035.
Polym. Adv. Technol. 5 (1994) 673–697, https://doi.org/10.1002/pat.1994.
[23] S. Shen, S.E. Kentish, G.W. Stevens, Shell-side mass-transfer performance in hollow-
220051102.
fiber membrane contactors, Solvent Extr. Ion Exch. 28 (2010) 817–844, https://doi.

13
E. Soroodan Miandoab, et al. Journal of Membrane Science 595 (2020) 117470

org/10.1080/07366299.2010.515176. oxide), J. Membr. Sci. 239 (2004) 105–117, https://doi.org/10.1016/j.memsci.


[24] R.E. Treybal, Mass Transfer Operations, third ed., McGraw-Hill College, n.d. 2003.08.031.
[25] R. Waack, N.H. Alex, H.L. Frisch, V. Stannett, M. Szwarc, Permeability of polymer [37] L. Zhu, M.T. Swihart, H. Lin, Tightening polybenzimidazole (PBI) nanostructure via
films to gases and vapors, Ind. Eng. Chem. 47 (1955) 2524–2527, https://doi.org/ chemical cross-linking for membrane H 2/CO 2 separation, J. Mater. Chem. 5
10.1021/ie50552a045. (2017) 19914–19923, https://doi.org/10.1039/C7TA03874G.
[26] T.C. Merkel, R.P. Gupta, B.S. Turk, B.D. Freeman, Mixed-gas permeation of syngas [38] D.R. Pesiri, B. Jorgensen, R.C. Dye, Thermal optimization of polybenzimidazole
components in poly(dimethylsiloxane) and poly(1-trimethylsilyl-1-propyne) at meniscus membranes for the separation of hydrogen, methane, and carbon dioxide,
elevated temperatures, J. Membr. Sci. 191 (2001) 85–94, https://doi.org/10.1016/ J. Membr. Sci. 218 (2003) 11–18, https://doi.org/10.1016/S0376-7388(03)
S0376-7388(01)00452-5. 00129-7.
[27] H. Gorissen, Temperature changes involved in membrane gas separations, Chem. [39] M. Azizi, S.A. Mousavi, CO2/H2 separation using a highly permeable polyurethane
Eng. Process: Process Intensification 22 (1987) 63–67, https://doi.org/10.1016/ membrane: molecular dynamics simulation, J. Mol. Struct. 1100 (2015) 401–414,
0255-2701(87)80032-6. https://doi.org/10.1016/j.molstruc.2015.07.029.
[28] Z. Han, A. Fina, Thermal conductivity of carbon nanotubes and their polymer na- [40] D. Jansen, M. Gazzani, G. Manzolini, E. van Dijk, M. Carbo, Pre-combustion CO2
nocomposites: a review, Prog. Polym. Sci. 36 (2011) 914–944, https://doi.org/10. capture, Int. J. Greenhouse Gas Control 40 (2015) 167–187, https://doi.org/10.
1016/j.progpolymsci.2010.11.004. 1016/j.ijggc.2015.05.028.
[29] C.L. Choy, Thermal conductivity of polymers, Polymer 18 (1977) 984–1004, [41] A. Mourgues, J. Sanchez, Theoretical analysis of concentration polarization in
https://doi.org/10.1016/0032-3861(77)90002-7. membrane modules for gas separation with feed inside the hollow-fibers, J. Membr.
[30] D.R. Anderson, Thermal Conductivity of Polymers, (n.d.) 14. Sci. 252 (2005) 133–144, https://doi.org/10.1016/j.memsci.2004.11.024.
[31] Aspen Custom Modeler, Aspen Modeler Reference Guide, (2004) https://esupport. [42] R. Rautenbach, W. Dahm, Gas permeation — module design and arrangement,
aspentech.com/S_Article?id=000001852. Chem. Eng. Process: Process Intensification 21 (1987) 141–150, https://doi.org/10.
[32] C.Y. Pan, Gas separation by high-flux, asymmetric hollow-fiber membrane, AIChE J. 1016/0255-2701(87)87003-4.
32 (1986) 2020–2027, https://doi.org/10.1002/aic.690321212. [43] F. Ahmad, K.K. Lau, A.M. Shariff, Y. Fong Yeong, Temperature and pressure de-
[33] M. SIDHOUM, Experimental behavior of asymmetric cellulose acetate membranes pendence of membrane permeance and its effect on process economics of hollow
and its use in novel separation schemes, AICHE Symp. Ser. 261 (1988) 102–112. fiber gas separation system, J. Membr. Sci. 430 (2013) 44–55, https://doi.org/10.
[34] X. Feng, J. Ivory, V.S.V. Rajan, Air separation by integrally asymmetric hollow-fiber 1016/j.memsci.2012.11.070.
membranes, AIChE J. 45 (1999) 2142–2152, https://doi.org/10.1002/aic. [44] B.-Z. Maytal, J.M. Pfotenhauer, Miniature Joule-Thomson Cryocooling Principles
690451013. and Practice, Springer, 2013.
[35] C.A. Scholes, J. Bacus, G.Q. Chen, W.X. Tao, G. Li, A. Qader, G.W. Stevens, [45] J.R. Roebuck, T.A. Murrell, E.E. Miller, The joule-thomson effect in carbon dioxide,
S.E. Kentish, Pilot plant performance of rubbery polymeric membranes for carbon J. Am. Chem. Soc. 64 (1942) 400–411, https://doi.org/10.1021/ja01254a048.
dioxide separation from syngas, J. Membr. Sci. 389 (2012) 470–477, https://doi. [46] P. Linstrom, NIST chemistry WebBook, NIST Standard Ref. Database 69 (1997),
org/10.1016/j.memsci.2011.11.011. https://doi.org/10.18434/T4D303.
[36] H. Lin, B.D. Freeman, Gas solubility, diffusivity and permeability in poly(ethylene

14

You might also like