Vous êtes sur la page 1sur 10

Chemical Engineering Journal 195196 (2012) 188197

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Membrane gassolvent contactor trials of CO2 absorption from syngas


Colin A. Scholes, Michael Simioni, Abdul Qader, Geoff W. Stevens, Sandra E. Kentish
Cooperative Research Centre for Greenhouse Gas Technologies (CO2CRC), Department of Chemical and Biomolecular Engineering, The University of Melbourne, VIC 3010, Australia

h i g h l i g h t s
" Membrane gas absorption tested with PP and PTFE membranes using carbonate and amine solvents.
" Solvent in shell side led to higher overall mass transfer coefcients than solvent in lumen.
" Pilot scale trials with syngas showed reduced performance due to membrane wetting.

a r t i c l e

i n f o

Article history:
Received 9 January 2012
Received in revised form 10 April 2012
Accepted 10 April 2012
Available online 27 April 2012
Keywords:
Membrane Contactor
Polypropylene
Potassium carbonate
Monoethanolamine
Polytetrauoroethylene
Syngas

a b s t r a c t
Membrane gassolvent contactors incorporate the advantages of both solvent absorption and membrane
gas separation technologies. Here, gassolvent contactors are applied to the separation of carbon dioxide
from syngas in a coal red pilot plant. Two contactors, based on polypropylene (PP) and polytetrauoroethylene (PTFE), are trialed with two solvents, 30 wt.% monoethanolamine (MEA) and 30 wt.% potassium
carbonate (K2CO3) solutions. To validate performance, results were also obtained with a mixture of 10%
CO2 in N2 in the laboratory. All contactorsolvent systems tested in the laboratory behaved in accordance
with membrane contactor models with only minor pore wetting observed. Mass transfer coefcients
were improved when solvent owed on the shell side of the contactor due to increased turbulence
and reduced pore wetting relative to the lumen side. In contrast, for the pilot plant trials with syngas,
only the PPK2CO3 and PTFEMEA systems provided mass transfer coefcients similar to those determined in the laboratory. For the PTFEK2CO3 system, additional pore wetting resulted in reduced overall
mass transfer coefcients. The PTFEMEA system retained the best overall mass transfer performance,
due to reduced pore wetting and greater reaction enhancement.
2012 Elsevier B.V. All rights reserved.

1. Introduction
The demonstration of carbon capture technologies is becoming
increasingly important, as solutions to reduce anthropogenic carbon emissions are sought. Two viable carbon capture technologies
are membrane gas separation and reversible solvent absorption
[1], both of which are currently commercialized in natural gas processing. Hybrid membranesolvent systems, known as membrane
gas absorption, seek to exploit the advantages of both membrane
gas separation and solvent absorption technologies [2]. The process
involves the transfer of CO2 from the process gas through a nonselective porous hollow-ber membrane where it is chemically absorbed into a solvent. This takes advantage of the highly selective
nature of solvent technology, while incorporating the benets of
membrane technology in terms of reduced equipment size, the
modular nature of the equipment, and exibility in orientation
[3]. A membrane contactor can achieve much greater mass transfer
area per unit volume than conventional solvent absorption column
Corresponding author. Tel.: +61 3 8344 6682; fax: +61 3 8344 4153.
E-mail address: sandraek@unimelb.edu.au (S.E. Kentish).
1385-8947/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.04.034

technology. Reed et al. [4] indicate that 500600 m2/m3 can be


achieved in a membrane contactor compared to 100250 m2/m3
in a traditional column. Similarly, Falk-Pedersen et al. [3] indicate
that the reduced specic area of a membrane contactor allows for
a 6575% reduction in weight and size compared to conventional
towers. Further, the membrane acts to physically separate the
liquid and gas ows, which eliminates foaming and reduces liquid
channeling, two major operating issues in solvent absorption
columns [2].
There are three main strategies for carbon capture from combustion processes, post-combustion capture, pre-combustion capture and oxy-red combustion [1]. In pre-combustion capture,
fossil fuels are reformed into synthesis gas (syngas) comprised
mainly of hydrogen and carbon monoxide [5]. More hydrogen is
produced by further converting CO through the water gas-shift
reaction, resulting in high pressure CO2 and H2 [6]. Separation of
these two components allows for the storage of CO2, while H2
can be used for a number of purposes, such as power generation
[1]. Pre-combustion processes can be further classied into those
that use oxygen-blown gasication and those that use an airblown gasier [7]. In the former case, the shifted syngas is a simple

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

189

Nomenclature
d
dnw
dw

e
s
Ai
Ao
C
CLM
CMEA
din
dh
DCO2
DG
Dl
Ds
E
G
Gz
Ke

membrane thickness (m)


membrane pore thickness that is not wetted (m)
membrane pore thickness that is wetted (m)
membrane porosity
tortuosity
mass transfer area based on inner ber diameter (m2)
mass transfer area based on outer ber diameter (m2)
equilibrium concentration of the bulk liquid (M)
log mean average of the inlet and outlet concentrations
in the bulk gas phase (M)
concentration of the MEA solvent (M)
inner diameter of the tube (m)
diameter of contactor shell side (m)
diffusion coefcient of CO2 in the liquid phase (m/s)
diffusivity in the gas phase (m/s)
diffusivity of CO2 in the lumen side (m/s)
diffusivity of CO2 in the shell side (m/s)
enhancement factor
inert gas owrate (mol/s)
Graetz dimensionless number
equilibrium constant for the reaction

mixture of CO2 and H2. In air-blown gasication the syngas is diluted by a signicant stream of nitrogen, complicating the CO2 capture operation. In particular there is need for separation
technologies that will remove CO2 from the syngas stream while
leaving H2 and N2 in the process gas.
A range of porous contactors and solvents have been trialed for
CO2 removal in such applications on the laboratory scale. The most
common membrane materials are polypropylene (PP), polyethylene (PE) and polytetrauoroethylene (PTFE) while water, amines
such as monoethanolamine, amino acid salts and NaOH have all
been tested as solvents [819]. Many of these studies have shown
that the potential for membrane gas contactors is limited by the
wetting of the membrane pores, which reduces the overall mass
transfer coefcient [8,20,21].
To date, the only signicant pilot plant trials of porous membrane contactors were those undertaken by Kvaerner for natural
gas sweetening in 19981999 [3,22]. Initial pilot plant trials used
a PTFE contactor with activated MDEA, while later trials employed
a physical solvent (Morphysorb). These pilot scale trials also identied membrane pore wetting as a major issue. Accurate pressure
regulation across the contactor was also essential to protect the
membrane hollow bers from rupture and collapse.
In this work, we report the performance of two porous hollow
ber membrane contactors for the separation of CO2 from airblown syngas in a pilot capture plant. These trials were conducted
as part of the CO2CRC Mulgrave capture project [23,24]. In support
of the pilot capture plant ndings, laboratory measurements of CO2
separation with similar contactors from a N2CO2 gas mixture are
also reported. The two contactors are made from PP and PTFE; both
of which have been reported in the literature as contactors for CO2
separation. Two solvent systems are studied; 30 wt.% monoethanolamine (MEA) and 30 wt.% potassium carbonate (K2CO3) solution. The rst solvent is considered a standard amine approach
for CO2 separation, while the second solvent has been widely applied in the Beneld process for CO2 separation [25].
2. Theory
The CO2 molar ux (N) through a membrane contactor into the
solvent is given by:

kg
kl
km
km,nw
km,w
kr;CO2
kt
K
l
m
N
Re
Sc
Sh
xo
Y

gas phase mass transfer coefcient (m/s)


liquid phase mass transfer coefcient (m/s)
membrane mass transfer coefcient (m/s)
mass transfer coefcients within the non-wetted membrane pore (m/s)
mass transfer coefcient with the wetted membrane
pore (m/s)
reaction rate coefcient between CO2 and the solvent
(L3/mol s)
mass transfer coefcient for the tube (lumen) side (m/s)
overall mass transfer coefcient based on the internal
diameter (m/s)
length of contactor (m)
partition coefcient
molar ux (mol/m2 s)
Reynolds dimensionless number
Schmidt dimensionless number
Sherwood dimensionless number
solvent loading
mole ratio of CO2 in the gas phase

Y in  Y out G
K i C LM
Ai

where G is the inert gas owrate, Ai is the mass transfer area based
on the internal ber diameter, K the overall mass transfer coefcient based on the internal diameter and Y is the mole ratio of
CO2 in the gas phase:

PCO2
P  PCO2

CLM represents the log mean average of the concentration driving force between the bulk gas phase and the liquid. As the reaction
of CO2 with the solvent is rapid, the equilibrium concentration of
CO2 that would be in equilibrium with the bulk liquid (C) can be
assumed to be zero [2]. This means that the log mean driving force
can be approximated simply by the log mean average of the inlet
and outlet concentrations of CO2 in the bulk gas phase:

C LM

C in  C out
in
ln CCout

The overall mass transfer across the membrane consists of three


mass transfer stages, the transfer of CO2 across the gas boundary
layer, the transfer of CO2 through the membrane pore and nally
the transfer of CO2 across the solvent boundary layer. Each of these
stages acts as a resistance to mass transfer, and is usually expressed as a series [25]. For solvent ow through the lumen side:

1
Ai
Ai
1

ki Ao kg ALM km mEkl

where kg, km and kl are the gas, membrane and liquid side physical
mass transfer coefcients respectively, Ai is the inner diameter area,
ALM the log mean area, m is the partition coefcient (dened as the
ratio of liquid to gas concentrations at equilibrium) and E is the
enhancement factor due to the chemical reaction in the solvent.
The partition coefcient for MEA is 0.76 at 25 C and K2CO3 at
35 C is 1.75 [26]. For non-wetted pores, kg and km  kl, and the
overall mass transfer coefcient can be approximated by:

K i mEkl

For the MEA system, the Enhancement Factor can be modeled


by a pseudo rst order system [27]:

190

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

ECO2

p
kr;CO2  C MEA  DCO2

kl

where kr;CO2 is the reaction rate coefcient of the reaction between


CO2 and the solvent, CMEA is the concentration of the MEA solvent,
and DCO2 is the diffusion coefcient of CO2 in the liquid phase. Both
kr;CO2 and DCO2 are functions of temperature, pressure and concentration. For K2CO3, the enhancement factor is given by [28]:

p
DCO2 kr;CO2 K e 1  xo =2xo
kl

where Ke is the equilibrium constant for the reaction and xo is the


loading dened by:

X0

molCO2
HCO3

molsolvent
K

and a function of CO2 partial pressure and temperature.


The mass transfer across the lumen side boundary layer is wellknown to be modeled by the GraetzLeveque correlation [29]:

Sht


 1=3
kt din
din
1:62 ScRe
Dl
l

Gz > 6

Or alternatively,

Sht

 
kt din
din
0:5  ScRe
Gz < 6
Dl
l

10

where Sh, Re, Sc and Gz are the dimensionless Sherwood, Reynolds,


Schmidt and Graetz numbers respectively [30]. The Graetz number
is:

Gz Sc  Re 

 
din
1

11

kt is the mass transfer coefcient for the tube (lumen) side, din the
inner diameter of the tube, Dl the diffusivity of CO2 in the lumen
phase, and l the length of the contactor.
A number of empirical correlations have been proposed for
modeling mass transfer through the shell side of membrane contactors [31]. The correlations that are most appropriate for the
experimental conditions experienced here is that developed by
Yang and Cussler [29,31] for randomly packed modules:

ShS


0:93
ks dh
Re  dh
1:25
Sc0:33
Ds
l

12

where ks is the shell side mass transfer coefcient and dh is the shell
side diameter. The Sherwood number is an effective ratio of the
convective to diffusive mass transport processes. The Reynolds
number is based on the randomly packed denition [32,33].
For gas-lled pores, the membrane pore resistance is the inverse of the mass transfer coefcient, given by [20]:

1
ds

km;nw DG e

13

Where d is the thickness, s is the tortuosity, e is the porosity of


the membrane and DG is the diffusivity in the gas phase. When the
pores become wetted, a more complex expression is required:

1
1
1

km km;nw mEkm;w

14

where km,nw and km,w represent the mass-transfer coefcients within


the non-wetted and wetted lengths of the pores (dnw + dw = d) [34]:

1
dw s

km;w DL e

15

1
dnw s

km;nw
DG e

16

Given all other parameters are known, the wetted length (dw)
can be evaluated from the experimentally determined overall mass
transfer coefcient. The pore wetting fraction is then give by:

Pore wetting

dw
d

17

The overall accuracy of this approach is 20% [31].


3. Experimental
The PP contactor used in all syngas capture pilot plant studies
was a LiquiCel MiniModule with 7400 bers (Membrana). Laboratory studies with K2CO3 used a similar LiquiCel MiniModule with
2300 bers. It was not possible to complete laboratory studies with
MEA, as the seals and plastic components of the smaller LiquiCel
module failed upon extended exposure to this solvent. Similarly,
the larger PP contactor could not be operated at 65 C in the pilot
plant trials because of a failure of the contactor seals at this temperature under the syngas conditions. The PTFE contactor for both
laboratory and syngas pilot plant studies was in-house built using
PTFE bers obtained from Markel Corporation. Details of the contactor dimensions and ber details are provided in Table 1 and a
schematic showing the arrangement in these contactors for a solvent-in-shell ow arrangement is provided in Fig. 2. MEA was supplied by Orica Chemicals Australia and made up to a 30 wt.%
solution, while potassium carbonate was supplied by SigmaAldrich and made up to a 30 wt.% solution. Details of these solvent
systems are provided in Table 2.
For the laboratory studies, a 90% N210% CO2 gas mixture was
supplied by BOC Gas Ltd. For the pilot capture plant studies, syngas
was supplied from an air-blown research gasier as part of the
CO2CRC Mulgrave Capture project [23,24]. The feed gas composition was provided by the gasication operator, with the average
composition across the full operational time in Table 3. However,
this composition varied considerably during and between days.
The feed syngas was cooled to ambient temperature through uninsulated pipework, before a separator vessel was used to remove
condensed water and heavy hydrocarbons. A lter was also present
to prevent any dust entering the process lines to the membrane
modules. The syngas was then reheated to either 35 or 65 C as required before entering the contactor module.
Both syngas capture pilot plant trials and laboratory measurements used the same membrane pilot plant (Fig. 1). This pilot capture plant was housed within a custom built cabinet. All valves and
ttings were made from stainless steel (Swagelok). Pressure
gauges (Swagelok) ranged between 0 at 12 bar and pressure transmitters (GEMS sensors and control Basingstoke England) with
displays (PR electronics) ranged between 0 and 12 bar for all lines.
A back-pressure regulator (Porter 08011) controlled the pressure
within the modules, with feed gas owrate measured by gas
rotameters (010 L/min Kytala (Muurame Finland)) and the retentate owrate measured by a gas meter (Ampy Email metering
Model 750) or universal owmeter (Agilent Technologies
ADM3000). All temperature measurements were by K-type thermocouples (ECE Fasil), with electronic displays (PR electronics).
The solvent (30 wt.% K2CO3 or 30 wt.% MEA), was stored in a
20 L tank and pumped (Micropump with Ismatec controller)
through the solvent circuit. Flowrate was measured on a rotameter
(Kyala) and controlled by the pump speed, with the differential
pressure across the module controlled manually on the solvent
side by a needle valve. To trap any solvent breakthrough on the
contactor and prevent damage to downstream instrumentation, a
separator was positioned on the retentate output of the module.
The exit gas from all lines was recombined before exhaust, with
non-return valves to prevent back ow. During syngas capture pilot plant studies, to prevent hazardous conditions downstream,

191

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197


Table 1
Specications of the four porous hollow-ber contactors.

Supplier
Shell diameter (m)
Length (m)
Outer ber diameter (m)
Inner ber diameter (m)
No. of bers
Average pore size (lm)
Porosity (%)
Mass transfer area (m2)

PP contactor
pilot plant

PP contactor laboratory
K2CO3 measurements

PTFE
contactor

Membrana
0.0425
0.14
0.0003
0.00022
7400
0.1
40
0.716

Membrana
0.018
0.10
0.0003
0.00022
2300
0.1
40
0.159

Markel
0.045
0.147
0.002
0.0016
19
0.16
22.5
0.014

Table 2
Characteristics of K2CO3 and MEA.
K2CO3 (30 wt.%)

MEA (30 wt.% at 35 C)

35 C

65 C

Density (g/L)
Viscosity (cP)
Diffusivity (m/s)
Interfacial tension

1286
3.43
2.62  109
81 mN/m [47,32]

1269 [43]
1.80 [45]
4.47  109 [46]

Enhancement factor at 10% loading


Sc No.

1.2
1017

1.6
318

Table 3
Average unshifted syngas feed composition (mol %) to the membrane pilot capture
plant.

CO2
H2
CO
N2
CH4
Heavy hydrocarbons
H2O

Mol%

Error

14.0
11.7
11.5
60.0
2.5
0.21
0.11

1.0
1.8
3.3
4.3
0.2
0.04
0.04

this exit gas was diluted with N2, and sent to the onsite air extraction system.
For syngas capture pilot plant studies, gas sampling was
achieved by connection of 10 mL sampling bombs at the analysis
points. The sample bomb was ushed three times with the process
gas before the sample was taken. All gas compositions were measured by gas chromatography (Hewett Packard, with a Molecular
Sieve and HP-PLOT Q column in series and carrier gas Helium).
For laboratory measurements, a dedicated CO2 analyzer (Horiba
VA3000) was used. Solvent sampling was achieved by collecting
the loaded solvent (10 mL) in sampling vials from the solvent
sample analysis point with analysis through standard methodologies [35].
The syngas feed to the module was at 1.3 bar(a) while the N2
CO2 gas mixture in the laboratory was at 1.05 bar(a). The pressure of the syngas feed was dictated by the gasier operator and
is considerably lower than the standard gasication pressure of
30 bar. Operating the membrane contactor at this lower pressure
is expected to have little inuence on the mass transfer through
the solvent and gas phases, but could limit the amount of pore wetting observed. The solvent pressure was held greater than the gas
pressure to prevent bubble formation in the solvent line. For each
owrate and temperature condition, the contactor was operated
for 6090 min before sampling was undertaken, to ensure steady-state conditions had been achieved. For the laboratory measurements, consecutive experiments were undertaken with the
solvent on both the lumen or shell side of the contactors. For the

999 [44]
1.90 [44]
2.44  109 [26]
56 mN/m [48]
63 mN/m [25]
82
9.6

capture pilot plant measurements under syngas, the solvent was


passed only on the lumen side of both contactors due to operational constraints. The average differential pressure across the
contactor through the pilot plant measurements was kept between
0.04 and 0.12 bar, though upon start-up a differential pressure
spike of 0.5 bar was often observed.
4. Results and discussion
The laboratory measurements under N2CO2 mixed gas conditions (post-combustion) are presented rst for both contactors
and solvents, followed by the performance of both contactors
and solvents in the syngas (pre-combustion) capture pilot plant
trials.
4.1. Laboratory results polypropylene contactor
The overall mass transfer coefcient for CO2 sorption through
the polypropylene contactor into K2CO3 is provided in Fig. 3 for solvent in the lumen side and Fig. 4 for solvent in the shell side. In all
cases there is a trend of increasing mass transfer coefcient with
increasing solvent owrate. This is associated with the higher solvent owrate increasing turbulence in the solvent side boundary
layer, reducing the resistance to CO2 transfer [2]. Operating at a
higher temperature is expected to lead to an increase in the mass
transfer coefcient because of an increase in the rate of reaction between carbonate and CO2 in the solvent boundary layer and a
reduction in solution viscosity, as well as an increase in diffusion
coefcient. However, for the solvent-in-lumen measurement no
difference in the mass transfer coefcient is observed with
temperature, while for the solvent-in-shell measurement, only a
slight increase is observed as temperature increases. This behavior
may reect the minimal change in the Enhancement factor between
these two temperatures [28] due to slow reaction kinetics (see
Table 2). The other reason for this behavior may relate to the reduction in interfacial tension that occurs as temperature increases. This
reduction causes a coincident reduction in the pore breakthrough
pressure and hence greater pore wetting [25]. The pore wetting
fraction can indeed be calculated from the overall mass transfer

192

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

Fig. 2. Schematic and photo of membrane gassolvent contactor pilot capture


plant. FIC = ow indicator and control, FI = ow indicator, TI = temperature indicator, PI = pressure indicator and PIC = pressure indicator and control.

Fig. 3. The CO2 overall mass transfer coefcient for a polypropylene contactor with
30 wt.% K2CO3 (at 35 and 65 C) on the lumen side upon exposure to 10% CO2 in N2
in the laboratory.

Fig. 1. Schematic of shell side solvent ow and lumen side gas ow.

coefcient using Eqs. (1)(17). The results reect this phenomenon,


with pore wetting greater at higher temperatures (Table 4).
Greater pore wetting is also observed for solvent-in-lumen ow
compared to solvent-in-shell ow (Table 4). This is due to a greater
solvent pressure drop along the length of the contactor. Solvent

ow through the lumen side generates a pressure drop of around


0.3 kPa at 35 C and 0.15 kPa at 65 C in this commercial module.
The pressure drop across the shell side for the solvent was
0.02 kPa irrespective of the solvent or the contactor. Hence, for
solvent-in-lumen ow, the higher feed pressure at the entry into
the lumen is more likely to force solvent into the pores, than when
solvent ow is on the shell side. For this lumen solvent ow, the
pore wetting increases as solvent owrate increases (Table 4),
reecting the increasing pressure drop and hence increasing feed
pressure as this owrate changes.
For K2CO3, the solvent-in-shell overall mass transfer coefcient
(Fig. 4) is an order of magnitude greater than that observed for the
solvent-in-lumen side (Fig. 3). This behavior reects both the
reduced pore wetting discussed above and greater uid mixing
[36]. Chun and Lee [11] report very similar mass transfer
coefcients (2  104 to 1  103 cm/s) at comparable solvent

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

193

2.6  103 cm/s respectively with solvent on the lumen side. These
higher coefcients reect considerably higher solvent owrates, 18
and 6 cm/s respectively, compared to the maximum solvent owrate of 0.44 cm/s reported here. Operating in more turbulent ow
conditions produces greater mass transfer. However, the shorter
residence time that results from a greater solvent ow results in
lower solvent loadings (Fig. 5). At an industrial scale this must be
compensated for by multiple passes through membrane modules
in series, to ensure that a full solvent loading is achieved. Differential pressures would also need to be carefully managed to avoid solvent breakthrough.
4.2. Laboratory results PTFE contactor

Fig. 4. The CO2 overall mass transfer coefcient for a polypropylene contactor with
30 wt.% K2CO3 (at 35 and 65 C) on the shell side upon exposure to 10% CO2 in N2 in
the laboratory.

Table 4
Pore wetting percentage (%) of PPK2CO3 determined from laboratory results.
Average owrate (L/min)

Lumen
35 C

Polypropylene and K2CO3 pore wetting (%)


0.005
3
0.008
4
0.012
5
0.017
5
0.018
6
0.022
9
0.024
13
0.028
19

Shell
65 C

35 C

65 C

5
6
7
9
11
14
12
18

0.6
0.3
0.1
0
0
0
0
0

0.6
0.8
0.8
1.5
0.8
1.6
1.5
1.0

The overall mass transfer coefcients for the PTFE contactor


with MEA in both the lumen and shell sides are provided in
Fig. 6. Again, the overall mass transfer coefcient increases with
solvent owrate because of increased turbulence in the solvent
boundary layer. There is also again clear evidence of increased
mass transfer when the solvent is on the shell side. The magnitude
of the mass transfer coefcient for the PTFEMEA system is larger
than that reported in the literature; Yeon et al. [38] reports a value
of 1.34  103 cm/s with 5 wt.% MEA and Kim and Yang [39] have a
maximum transfer coefcient of 2.5  103 cm/s for 4 wt.% MEA, at
a comparable solvent owrate. The differences probably reect the
more concentrated MEA used in the current work (30 wt.%) which
will increase reaction enhancement.
For both the PTFE and PP contactor, the pore wetting fractions
(Tables 4 and 5) are generally lower than those observed in our
prior work, which used at sheet membranes with MEA solvent.
In this case, pore wetting fractions from 24% to over 100% were observed [34]. The higher values in this case may reect the at sheet
format, which used overhead stirring, rather than a crossow
arrangement.
The overall mass transfer coefcients for PTFE with K2CO3 at 35
and 65 C are provided in Fig. 7 for solvent on the lumen side and
Fig. 8 for solvent on the shell side. The mass transfer coefcients
observed are an order of magnitude lower than for the MEA system, reecting the low rate of reaction for the carbonateCO2 reaction at these temperatures and hence the lower reaction
enhancement (see Table 2). In these systems there is a clear distinction that the higher temperature does enhance the mass trans-

Fig. 5. The CO2 loading in 30 wt.% K2CO3 as a function of the ratio of solvent to gas
owrate for PP contactor, solvent on the lumen side at 35 (d) and 65 C (j), and on
the shell side at 35 (d) upon exposure to 10% CO2 in N2 in the laboratory.

owrates for 15% K2CO3 and solvent in the shell. In contrast, both
Dindore et al. [37] as well as Karoor and Sirkar [14] report higher
mass transfer coefcients for PPH2O systems of 2.3  103 and

Fig. 6. The CO2 overall mass transfer coefcient for a PTFE contactor with 30 wt.%
MEA, on both lumen and shell side at 35 C upon exposure to 10% CO2 in N2 in the
laboratory.

194

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197


Table 5
Pore wetting percentage (%) of PTFEMEA and K2CO3 determined from laboratory
results.
Average owrate (L/min)

Lumen
MEA

Shell
K2CO3
35 C

K2CO3
65 C

PTFE with MEA and K2CO3 pore wetting (%)


0.005
0
13
16
0.007
25
17
0.011
1
21
18
0.017
1
26
19
0.020
3
21
0.025
22
21
0.026
3
25
0.031
3
21
0.034
0.046
3

Fig. 7. The CO2 overall mass transfer coefcient for a PTFE contactor with 30 wt.%
K2CO3 (at 35 and 65 C) on lumen side upon exposure to 10% CO2 in N2 in the
laboratory.

MEA

K2CO3
35 C

2
2
2
2
2
2

2
3
3

K2CO3
65 C

3
2
2
2

6
6
6

for solvent ow through the bers, on average 0.47 kPa compared


to 0.02 kPa for the shell, For the solvent-in shell arrangement, the
overall mass transfer coefcient for PTFEK2CO3 is of similar magnitude to the PPK2CO3 system at 35 C.
In this laboratory work, pore wetting in the PTFE contactor is
similar to that in the PP contactor when the solvent is on the lumen
side. While PTFE is more hydrophobic [41], the pore diameter in
these membrane bers is larger (0.16 lm) than in the PP contactor
(0.1 lm) and this larger pore size offsets the changes in hydrophobicity. There is a clear difference in the overall mass transfer coefcient, with MEA providing an order of magnitude faster mass
transfer than K2CO3. This relates to the much faster reaction rate
of the MEA solvent with CO2. The criteria for solvent selection must
include consideration of this reaction rate, as it strongly dictates
the observed overall mass transfer coefcient. However, solvent
resistance is also important, especially for corrosive amine solvents
such as MEA.

4.3. Pilot capture plant with syngas


For the PPK2CO3 system there is good agreement between the
pilot capture plant results and those determined under laboratory
conditions (Fig. 9), with a consistent trend as the Reynolds number
increases.
Fig. 8. The CO2 overall mass transfer coefcient for a PTFE contactor with 30 wt.%
K2CO3 (at 35 and 65 C) on the shell side upon exposure to 10% CO2 in N2 in the
laboratory.

fer rate, for both lumen and shell solvent ow. Indeed for the shell
side, there is an enhancement of 1.7 times for the 65 C system
compared to the 35 C. The mass transfer coefcients are in good
agreement with the literature, Nii and Takeuchi [40] report for
PTFE with 2 M K2CO3 solution a maximum mass transfer coefcient of 5.6  104 cm/s at ambient temperature and with solvent
on the lumen side, which is of similar magnitude to that measured
in Fig. 7.
The solvent-in-lumen result for the mass transfer coefcient for
K2CO3 in the PTFE contactor is an order of magnitude greater than
that in the PP contactor. This is believed to be associated with the
signicantly smaller number of bers in the PTFE contactor. For a
comparable solvent owrate, much higher solvent velocities are
obtained in the PTFE contactor leading to higher Reynolds numbers. This contactor has a maximum Reynolds number of 9 on
the lumen side, while for PP this value is 0.14. Again, greater pore
wetting is observed on the lumen side than on the shell side
(Table 5) and this is attributed to the greater pressure required

Fig. 9. The CO2 overall mass transfer coefcient for a polypropylene contactor with
30 wt.% K2CO3, at 35 C on lumen side, for laboratory (d) and pilot plant
measurements (j).

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

Fig. 10. Overall CO2 mass transfer coefcient for the polypropylene contactor with
30 wt.% MEA, at 35 C on the lumen side for pilot plant measurements (j), as a
function of the solvent Reynolds Number.

For the PPMEA system (Fig. 10), the pilot capture plant results
indicate reduced performance under syngas conditions, with the
overall mass transfer coefcient falling by around an order of magnitude compared to that expected from the literature. Yeon et al.
[38] report a value for a polyvinylideneuoride contactor with 5%
MEA of 2.1  103 cm/s, while Kosaraju et al. [42] for a poly(4methyl-1-pentene) contactor with 11.6% MEA reports a slightly
lower mass transfer coefcient of 0.96  103 cm/s. In our own
published work with PP and 30 wt.% MEA, a value of 0.05 m/s is obtained at slightly higher Reynolds numbers [25]. Indeed, the values
for this system under pilot capture plant conditions are comparable with K2CO3 even though this solvent has a much smaller reaction enhancement (Table 2).
This reduction in mass transfer coefcient is associated with
greater pore wetting by the MEA solvent (Table 6). The cause of
this wetting may be associated with operational issues on startup, where it was difcult to control the differential pressure across
the membrane as initial syngas entered the system. Furthermore,
the syngas supply pressure varied considerably during operation
(0.2 bar). The response time lag in changing solvent pressure
was slow compared to these gas pressure uctuations making differential pressure control extremely difcult. This will be a critical
issue in operating membrane gassolvent contactors on a large
scale.
The CO2 loading over the Reynolds number range was a relatively constant low value of 0.070.10 mol CO2 per mol of MEA.
Again, much longer ow paths will be required in a full scale operation to ensure greater solvent loadings.
Figs. 11 and 12 compare pilot capture plant and laboratory results for the PTFEK2CO3 system at 35 and 65 C respectively. It
is clear that the pilot capture plant performance of the PTFE contactor is again reduced compared to the laboratory case. This can

195

Fig. 11. Overall CO2 mass transfer coefcient for the PTFE contactor with 30 wt.%
K2CO3, at 35 C on the lumen side, for laboratory (d) and pilot plant measurements
(j) as a function of the solvent Reynolds number.

Fig. 12. Overall CO2 mass transfer coefcient for the PTFE contactor with 30 wt.%
K2CO3, at 65 C on the lumen side, for laboratory (d) and pilot plant measurements
(j) as a function of the solvent Reynolds number.

Table 6
Pore wetting percentage (%) for PP and PTFE pilot plant trials.

Solvent in

Polypropylene
Pilot plant
Lumen

Poly tetrauoroethylene
Pilot plant
Lumen

K2CO3, 35 C
K2CO3, 65 C
MEA, 35 C

33

41

12
9
32

Fig. 13. Solvent breakthrough on the PTFE contactor under syngas conditions, as
evidenced by the beads of solvent appearing at the left (feed entry) end.

196

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197

limiting the experiments that could be conducted with this


solvent.
Acknowledgements
The authors would like to thank HRL for access to equipment, as
well as the contributions of Joannelle Bacus, Wen Tao, George Chen
and Gang Li. Funding for this project is provided by the CRC for
Greenhouse Gas Technologies (CO2CRC) through the Australian
Government Cooperative Research Centre program and facilities
from the Particulate Fluids Processing Centre of the University of
Melbourne.
References

Fig. 14. Overall CO2 mass transfer coefcient for the PTFE contactor with 30 wt.%
MEA, at 35 C on the lumen side, for laboratory (d) and pilot plant measurements
(j) as a function of the solvent Reynolds number.

again be attributed to additional pore wetting of the membrane


(Table 6). Indeed, solvent breakthrough was visually observed at
the feed end of this contactor (Fig. 13) where the solvent pressure
was highest. The corresponding solvent loading at both temperatures was also low because of the lower packing density and small
contactor area, with values between 0.01 and 0.05 mol CO2/mol
K2CO3 observed.
For the PTFEMEA system, the overall mass transfer coefcient
observed under syngas conditions is equivalent to that measured
in the laboratory (Fig. 14). This implies that there is no additional
pore wetting in this system and optimal mass transfer is occurring.
This result is somewhat surprising as MEA would be expected to
wet the PTFE pores more readily than K2CO3 due to its lower surface tension (Table 2). The results in this case may represent a better control of transmembrane pressure drop during startup. The
CO2 loading in MEA for the syngas experiments was very low at
0.020.05 over the Reynolds number range, mostly due to the
low mass transfer area for this contactor.

5. Conclusion
CO2 separation from syngas through membrane gassolvent
contactors has been demonstrated. Pilot capture plant results show
that a PTFE contactor with 30 wt.% MEA solution provided the best
overall mass transfer coefcient, consistent with literature expectations. The presence of pore wetting in the PTFE contactor with
K2CO3 solvent and the PP contactor with MEA solvent decreased
the performance of these contactorsolvent systems. In addition,
the low rate of reaction for the K2CO3 solvent at the temperatures
used meant that lower overall mass transfer coefcients were observed, even when pore wetting was minor.
The application of membrane gassolvent contactors to carbon
capture from syngas will rely on the development of membrane
materials and solvent systems that resist pore wetting. However,
this work has also shown the importance of developing process
control systems that can minimize pressure differential spikes between the solvent and gas sides of the contactor, which force solvent into the pores. This is consistent with the ndings from the
original Kvaerner work [3,22]. The work has also shown the importance of careful materials selection in module design, with degradation to the module housing and seals by the MEA solvent

[1] K. Thambimuthu, M. Soltanieh, J.C. Abandas, IPCC Special Report on Carbon


Dioxide Capture and Storage, Cambridge University Press, Cambridge, 2005.
[2] J. Franco, D. deMontigny, S. Kentish, J. Perera, G. Stevens, A study of the mass
transfer of CO2 through different membrane materials in the membrane gas
absorption process, Sep. Sci. Technol. 43 (2008) 225244.
[3] O. Falk-Pedersen, M.S. Gronvold, P. Nokleby, F. Bjerve, H.F. Svendsen, CO2
capture with membrane contactors, Int. J. Green Energy 2 (2005) 157165.
[4] B.W. Reed, M.J. Semmens, E.L. Cussler, Membrane contactors, in: R.D. Noble,
S.A. Stern (Eds.), Membrane Separations Technology: Principles and
Applications, Elsevier Science, New York, 1995.
[5] H. Audus, O. Kaarstad, G. Skinner, CO2 Capture by Pre-Combustion
Decarbonisation of Natural Gas, Interlaken, Switzerland, 1998.
[6] C. Higman, M. van der Burgt, Gasication, Gulf Professional Publications/
Elsevier Science, Amsterdam, 2008.
[7] C.A. Scholes, K.H. Smith, S.E. Kentish, G.W. Stevens, CO2 capture from precombustion processes strategies for membrane gas separation, Int. J.
Greenhouse Gas Control 4 (2010) 739755.
[8] R. Wang, H.Y. Zhang, P.H.M. Feron, D.T. Liang, Inuence of membrane wetting
on CO2 capture in microporous hollow ber membrane contactors, Sep. Purif.
Technol. 46 (2005) 3340.
[9] K.A. Hoff, O. Juliussen, O. Falk-Pedersen, H.F. Svendsen, Modeling and
experimental study of carbon dioxide absorption in aqueous alkanolamine
solutions using a membrane contactor, Ind. Eng. Chem. Res. 43 (2004) 4908
4921.
[10] H.B. Al-Saffar, B. Ozturk, R. Hughes, A comparison of porous and non-porous
gasliquid membrane contactors for gas separation, Chem. Eng. Res. Des. 75
(1997) 685692.
[11] M.-S. Chun, K.-H. Lee, Analysis on a hydrophobic hollow-ber membrane
absorber and experimental observations of CO2 removal by enhanced
absorption, Sep. Sci. Technol. 32 (1997) 24452466.
[12] K. Li, J.F. Kong, X. Tan, Design of hollow bre membrane modules for soluble
gas removal, Chem. Eng. Sci. 55 (2000) 55795588.
[13] P. Feron, A. Jansen, CO2 separation with polyolen membrane contactors and
dedicated absorption liquids: performances and prospects, Sep. Purif. Technol.
27 (2002) 231242.
[14] S. Karoor, K.K. Sirkar, Gas absorption studies in microporous hollow ber
membrane modules, Ind. Eng. Chem. Res. 32 (1993) 674684.
[15] H. Kreulen, C. Smolders, G. Versteeg, W. van Swaaij, Microporous hollow bre
membrane modules as gas liquid contactors. Part 2. Mass transfer with
chemical reaction, J. Membr. Sci. 78 (1993) 217238.
[16] H. Matsumoto, H. Kitamura, T. Kamata, M. Ishibashji, H. Ota, Effect of
membrane properties of microporous hollow-ber gasliquid contactor on
CO2 removal from thermal power plant ue gas, J. Chem. Eng. Jpn. 28 (1995)
125128.
[17] M. Mavroudi, S.P. Kaldis, G.P. Sakellaropoulos, Reduction of CO2 emissions by a
membrane contacting process, Fuel 82 (2003) 21532159.
[18] Z. Qi, E.L. Cussler, Microporous hollow bres for gas absorption I: mass transfer
in the liquid, J. Membr. Sci. 23 (1985) 321332.
[19] P.H.M. Feron, A.E. Jansen, CO2 separation with polyolen membrane contactors
and dedicated absorption liquids: performances and prospects, Sep. Purif.
Technol. 27 (2002) 231242.
[20] S. Khaisri, D. DeMontigny, P. Tontiwachwuthikul, R. Jirarata-nanon, A
mathematical model for gas absorption membrane contactors that studies
the effect of partially wetted membranes, J. Membr. Sci. 347 (2010) 228239.
[21] D. deMontigny, P. Tontiwachwuthikul, A. Chakma, Using polypropylene and
polytetrauoroethylene membranes in a membrane contactor for CO2
absorption, J. Membr. Sci. 277 (2006) 99107.
[22] H. Herzog, O. Falk-Pedersen, The Kvaerner membrane contactor: lessons from
a case study in how to reduce capture costs, in: 5th International Conference
on Greenhouse Gas Control Technologies, Cairns, Australia, 2000.
[23] C. Anderson, C. Scholes, A. Lee, K. Smith, S. Kentish, G. Stevens, P. Webley, A.
Qader, B. Hoooper, Novel pre-combustion capture technologies in action
results of the CO2CRC/HRL Mulgrave capture project, Energy Proc. (2011)
11921198.
[24] A. Qader, Demonstrating carbon capture, Chem. Eng. (November) (2009) 52
53.

C.A. Scholes et al. / Chemical Engineering Journal 195196 (2012) 188197


[25] J.A. Franco, D. DeMontigny, S.E. Kentish, J.M. Perera, G.W. Stevens,
Poly(tetrauoroethylene) sputtered polypropylene membranes for carbon
dioxide separation in membrane gas absorption: hollow ber conguration,
Ind. Eng. Chem. Res. 51 (2012) 13671382.
[26] J. Franco, CO2 Separation Using a Modied Polypropylene Gas Absorption
Membrane, PhD Thesis, Department of Chemical and Biomolecular
Engineering, University of Melbourne, 2007.
[27] P.V. Danchwerts, Gas Liquid Reactions, McGraw Hill, New York, 1970.
[28] D.W. Savage, G. Astarita, S. Joshi, Chemical absorption and desorption of
carbon dioxide from hot carbonate solutions, Chem. Eng. Sci. 35 (1980) 1513
1522.
[29] M. Yang, E.L. Cussler, Designing hollow-ber contactors, AIChE J. 32 (1986)
19101916.
[30] S. Shen, K.H. Smith, S.E. Kentish, G.W. Stevens, Comparison of shell side mass
transfer correlations in randomly packed hollow ber membrane modules,
Desal. Water Treatm. 17 (2010) 739755.
[31] S. Shen, S.E. Kentish, G.W. Stevens, Shell-side mass-transfer performance in
hollow-ber membrane contactors, Solv. Extr. Ion Exchange 28 (2010) 817
844.
[32] M.J. Costello, A.G. Fane, P.A. Hogan, R.W. Schoeld, The effect of shell-side
hydrodynamics on the performance of axial ow hollow ber modules, J.
Membr. Sci. 80 (1993) 111.
[33] R. Prasad, K.K. Sirkar, Dispersion-free solvent extraction with microporous
hollow-ber moedules, AIChE J. 34 (1988) 177188.
[34] J.A. Franco, S.E. Kentish, J.M. Perera, G.W. Stevens, Poly(tetrauoroethylene)
sputtered polypropylene membranes for carbon dioxide separation in
membrane gas absorption, Ind. Eng. Chem. Res. 50 (2011) 40114020.
[35] W. Horowitz, Ofcial Methods of Analysis of the Association of Ofcial
Analytical Chemists, 12th ed., Washington Association of Ofcial Analytical
Chemists, Washington, DC, 1974.
[36] S. Majumdar, K.K. Sirkar, Hollow-ber contained liquid membrane, in: W.S.
Winston, K.K. Sirkar (Eds.), Membrane Handbook, Kluwer Academic
Publishers, Norwell, Massachusetts, 2001, pp. 764808.

197

[37] V.Y. Dindore, D.W.F. Brilman, G.F. Versteeg, Hollow ber membrane contactor
as a gasliquid model contactor, Chem. Eng. Sci. 60 (2005) 467479.
[38] S.-H. Yeon, B. Sea, Y.-I. Park, K.-H. Lee, Determination of mass transfer rates in
PVDF and PTFE hollow ber membranes for CO2 absorption, Sep. Sci. Technol.
38 (2003) 271293.
[39] Y.-S. Kim, S.-M. Yang, Absorption of carbon dioxide through hollow ber
membranes using various aqueous absorbents, Sep. Purif. Technol. 21 (2000)
101109.
[40] S. Nii, H. Takeuchi, Removal of CO2 and/or SO2 from gas streams by a
membrane absorption method, Gas Sep. Purif. 8 (1994) 107114.
[41] Z.-Y. Xi, Modication of polytetrauoroethylene porous membranes by
electron beam initiated surface grafting of binary monomers, J. Membr. Sci.
339 (2009) 3338.
[42] P. Kosaraju, A.S. Kovvali, A. Korikov, K.K. Sirkar, Hollow ber membrane
contactor based CO2 absorption stripping using novel solvents and
membranes, Ind. Eng. Chem. Res. 44 (2005) 12501258.
[43] M. Simioni, Membrane Stripping: Desorption of Carbon Dioxide from Alkali
Solvents, PhD Thesis, Department of Chemical and Biomolecular Engineering,
University of Melbourne, 2010.
[44] M.-J. Lee, T.-K. Lin, Density and viscosity for monoethanolamine + water, +
ethanol, and + 2-propanol, J. Chem. Eng. Data 40 (1995) 336339.
[45] Z. Palaty, Viscosity of aqueous solutions of potassium carbonate/potassium
bicarbonate (K2CO3/KHCO3), Chem. Biochem. Eng. Quart. 7 (1993) 155159.
[46] G.A. Ratcliff, J.G. Holdcroft, Diffusivities of gases in aqueous electrolyte
solutions, Trans. Inst. Chem. Eng. 41 (1963) 315319.
[47] R. Pohorecki, W. Moniuk, Calculation of densities, viscosities and surface
tensions of aqueous solutions of potassium and sodium hydroxides and
carbonates, Prace Instytutu Inzynierii Chemicznej Politechniki Warszawskiej
11 (1982) 127140.
[48] G. Vazquez, E. Alvarez, J.M. Navaza, R. Rendo, E. Romero, Surface tension of
binary mixtures of water + monoethanolamine and water + 2-amino-2methyl-1-propanol and tertiary mixtures of these amines with water from 25
C to 50 C, J. Chem. Eng. Data 42 (1997) 5759.

Vous aimerez peut-être aussi