You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/264615211

Highly permeable membrane materials for CO2 capture

Article  in  Journal of Materials Chemistry A · October 2013


DOI: 10.1039/C3TA13066E

CITATIONS READS

52 1,769

7 authors, including:

Qiang Fu Andri Halim


University of Melbourne University of Melbourne
60 PUBLICATIONS   2,561 CITATIONS    6 PUBLICATIONS   266 CITATIONS   

SEE PROFILE SEE PROFILE

Jinguk Kim Joel M. P. Scofield


Singapore University of Technology and Design University of Melbourne
16 PUBLICATIONS   579 CITATIONS    13 PUBLICATIONS   465 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Gas separation by membrane technology View project

Valorization of mine wastes from Québec by mineral carbonation View project

All content following this page was uploaded by Paul Gurr on 18 June 2015.

The user has requested enhancement of the downloaded file.


Journal of
Materials Chemistry A
View Article Online
PAPER View Journal | View Issue
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

Highly permeable membrane materials for CO2 capture†


Cite this: J. Mater. Chem. A, 2013, 1,
Qiang Fu,ab Andri Halim,ab Jinguk Kim,ab Joel M. P. Scofield,ab Paul A. Gurr,ab
13769 Sandra E. Kentisha and Greg G. Qiao*ab

The release of large quantities of CO2 into the atmosphere has been linked to global warming and climate
anomalies. Membrane processes offer a potentially viable energy-saving alternative for CO2 capture in
comparison with conventional technologies such as amine absorption. However, gas separation
membranes that are currently available have insufficiently high permeance (flux) for large scale
applications such as the treatment of high volume flue gas with low concentration of CO2. Here we
demonstrate a class of thin film composite (TFC) membranes, consisting of a high molecular weight
amorphous poly(ethylene oxide)/poly(ether-block-amide) (HMA-PEO/Pebax 2533) selective layer and a
highly permeable polydimethylsiloxane (PDMS) intermediate layer which was pre-coated onto a
polyacrylonitrile (PAN) microporous substrate. In contrast to the performance of conventional materials,
Received 5th August 2013
Accepted 13th September 2013
the selective layer of TFC membranes shows super-permeable characteristics and outstanding CO2
separation performance. This unprecedented result arises from the introduction of HMA-PEOs into the
DOI: 10.1039/c3ta13066e
Pebax 2533 matrix, leading to high CO2 permeability and flux. These results provide an encouraging
www.rsc.org/MaterialsA direction to further develop TFC membranes for efficient CO2 capture processes.

1. Introduction Polyethylene oxide (PEO) has attracted much attention in


CO2 separation membranes as the ethylene oxide (EO) unit can
The use of polymeric membranes has been considered as an exhibit favorable interactions with CO2 compared to light gases,
effective tool to separate CO2 from a mixture of gases due to i.e. H2, He, N2 and CH4.13–18 Recent studies have been focused on
their adaptability, operational simplicity, low energy a facile approach to enhance the CO2 separation performance
consumption, low capital cost and environmental friendliness by blending PEO polymers into poly(ether-block-amide)
when compared to conventional separation methods such as (Pebax) copolymers.19–21 Pebax is a thermoplastic elastomer
absorption with amine solutions.1,2 However, current commer- combining linear chains of rigid polyamide (PA) segments
cial membranes are not competitive for this application as their interspaced with exible polyether (PE) segments.22–24 It is
permeance (ux) for CO2 is low.3 The development of novel thin believed that the hard PA block provides the mechanical
lm composite (TFC) membranes has gained much attention strength, whereas gas transport occurs primarily through the
due to their potential for increased ux and hence higher energy so PE block.20,21,25 The blend membranes of Pebax 1657 and
efficiency for gas separation.3–10 In order to manufacture defect- gra copolymers containing PEO segments exhibit high CO2
free membranes, the TFC membrane generally should have a permeability and high CO2/N2 and CO2/H2 selectivity.26
thin (<300 nm) polydimethylsiloxane (PDMS) lm as an inter- Blending PEO200 with Pebax 1657 nearly doubles the CO2
mediate layer between the substrate and the selective layer. This permeability.20,21 However, the PEO additives have been limited
concept was developed by Lundy11 and demonstrated in to low molecular weight PEOs (ca. Mn < 300 Da) as high
previous studies.3,4,12 The highly permeable PDMS intermediate molecular weight PEOs cause crystallinity, which leads to a
layer works as a protective coating which prevents the pene- decrease of free volume and lower permeability.19 Yave and Car
tration of diluted polymer solution into the porous structure blended the poly(ethylene oxide)-poly(butylene terephthalate)
and renders the entire membrane surface smoother.4,6 (PBT-PEO) copolymer with low molecular weight PEO200, and
they observed an increase in CO2 permeability (from 150 to 750
Barrers) without loss of selectivity.7 Although high molecular
a
Cooperative Research Centre for Greenhouse Gas Technologies, Department of weight amorphous PEO (HMA-PEO) is expected to signicantly
Chemical and Biomolecular Engineering, The University of Melbourne, VIC 3010, increase the proportion of EO moieties,27 there is no existing
Australia. E-mail: gregghq@unimelb.edu.au; Fax: +61 3 8344 4153; Tel: +61 3 8344
method to synthesize HMA-PEOs and therefore there is no study
8665
b
on the effects of blending HMA-PEO into membrane systems.
Polymer Science Group, Department of Chemical and Biomolecular Engineering, The
University of Melbourne, VIC 3010, Australia In the present work, we introduce a new approach of
† Electronic supplementary information (ESI) available. See DOI: blending high molecular weight amorphous PEO with high EO
10.1039/c3ta13066e moieties into the existing membrane system. Pebax 2533,

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 13769–13778 | 13769
View Article Online

Journal of Materials Chemistry A Paper

which is composed of 20% amide groups and 80% ether link- 2.2 Measurements
ages (Scheme 1), was selected as the base polymer for the
Gel permeation chromatography (GPC) was performed on a
membrane matrix. A multi-block HMA-PEO copolymer is
Shimadzu liquid chromatography system tted with a Wyatt
synthesized and the multi PEO blocks are interspaced with a
DAWN HELEOS LS detector (l ¼ 658 nm), Shimadzu RID-10
specially designed terephthalic spacer, which reduces the PEO
refractometer (l ¼ 633 nm) and Shimadzu SPD-20A UV-vis
crystallinity. The incorporation of HMA-PEO into the Pebax
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

detector, using three identical Polymer Laboratories' PLgel


matrix to form the selective layer displayed unprecedented CO2
columns (5 mm, MIXED-C) and DMF with 0.05 M LiBr (70  C,
separation ability (permeance >900 GPU, 1 GPU ¼ 106 cm3 1 mL min1) as the mobile phase. Astra soware (Wyatt Tech-
(STP) cm2 s1 cm Hg1). In addition, we also found that this nology Corp.) was used to process the GPC data and determine
blending method successfully suppresses any possible crystal-
molecular weight characteristics based on the assumption of
linity caused by HMA-PEO copolymers. The synthesized HMA-
100% mass recovery of the polymer. Proton magnetic resonance
PEOs are fully characterized by gel permeation chromatography
spectroscopy (1H NMR) was performed using a Varian Unity
(GPC) and proton magnetic resonance spectroscopy (1H NMR).
(400 MHz) spectrometer, where the samples were dissolved in
Pebax/HMA-PEO TFC blend membranes are characterized by
CDCl3. X-ray diffraction (XRD) patterns of the samples were
scanning electron microscopy (SEM), X-ray photoelectron
recorded on a Bruker D8 Advance instrument with Cu Ka
spectroscopy (XPS) and X-ray diffraction (XRD) measurements. radiation (40 kV, 40 mA) and a nickel lter, and the samples
Calculations show that the selective layer presented were exposed at a scanning rate of 2q ¼ 0.020 s1 in the range
outstanding CO2 separation properties with a CO2 permeability
of 3–70 . SEM images were acquired using a FEI Quanta 200
of 780 Barrer, a CO2 permeance of 2000 GPU and a CO2/N2
ESEM FEG. Samples were pre-coated with gold using a Dynavac
selectivity of 40. These results surpassed the most recent upper
Mini Sputter Coater prior to imaging. X-ray photoelectron
bounds and make this simple HMA-PEO copolymer an attrac-
spectroscopy (XPS) analysis was performed on a VG ESCALAB
tive additive for advanced CO2 separation membranes.
220i-XL spectrometer under an ultra-high vacuum (6  109
mbar) to reveal the surface composition of the TFC membrane.
2. Experimental A xed photon energy (Al Ka 1486.6 eV) was used. A survey scan
2.1 Materials was performed between 0 and 1200 eV with a resolution of
1.0 eV and a pass energy of 100 eV. High resolution scans for C
1,3,5-Benzenetricarbonyl trichloride (TMC, 98%), terephthaloyl
1s (276 to 296 eV) and O 1s (522 to 542 eV) were also conducted
chloride (TCL, >99%), and poly(ethylene oxide) (PEO400, Mn ¼
with a resolution of 0.05 eV and a pass energy of 20 eV.
400 Da and PEO1k, Mn ¼ 1.0 kDa) were purchased from Sigma-
Aldrich and used as received. Aminopropyl terminated polydi-
methylsiloxane (NH2–PDMS–NH2, Mn ¼ 5.0 kDa) was purchased 2.3 Synthesis of multi-block HMA-PEO copolymers
from Gelest Inc. and used as received. Deuterated chloroform The condensation polymerization of the PEO macromonomer
(CDCl3, 99.9%) was purchased from Cambridge Isotope Labo- with terephthaloyl chloride (TCL) occurred under ambient
ratories, Inc. Triethylamine (TEA) was distilled from calcium conditions. A typical example can be described as follows.
hydride under argon. Tetrahydrofuran (THF) was distilled from PEO400 (2.0 g, 5 mmol; 10 mmol of hydroxyl groups) was dis-
benzophenone and sodium metal under argon. AR grade solved in a mixture of 50 mL of anhydrous THF and 4.2 mL of
dichloromethane (DCM), diethyl ether (DEE), 2-propanol, anhydrous TEA (3.05 g, 30 mmol) under dry nitrogen. TCL
n-butanol, n-hexane and other solvents were purchased from (0.952 g, 4.69 mmol) was dissolved in 10 mL of anhydrous THF
Chem-Supply Pty. Ltd. and used as received. The microporous and the solution was added dropwise to PEO solution at 0  C
polyacrylonitrile (PAN) substrate was purchased from SolSep BV over 30 min under vigorous stirring. The mixture was then
and used without any further treatment. warmed to room temperature and stirred for a further 24 h. The

Scheme 1 The fabrication of a thin film composite blend membrane. (I) The PDMS gutter layer was formed by the cross-linking of amino-terminated PDMS and 1,3,5-
benzenetricarbonyl trichloride (TMC). (II) The thin film composite blend membrane was prepared by spin-coating the mixture of Pebax 2533 and HMA-PEO onto the
PDMS gutter layer.

13770 | J. Mater. Chem. A, 2013, 1, 13769–13778 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry A

precipitate of triethylamine hydrochloride was removed by 2.5 Gas permeation measurements


ltration. The solution was then concentrated in vacuo (1 mbar)
Single gas permeation measurements of the TFC membranes
and the product was re-dissolved in 50 mL of DCM. The organic
were made using an in-house built constant pressure variable
solution was then washed with H2O (3  50 mL), dried (anhy-
volume (CPVV) apparatus shown in Fig. 1. The TFC blend
drous MgSO4), ltered, concentrated and precipitated into cold
membranes were installed in a stainless cell and were tested for
DEE. Aer removal of DEE, the gel-like polymer was dried in
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

single gases (N2 and CO2) under a range of pressures (340 to


vacuo (1 mbar) at 30  C. A colorless product, multi-block HMA-
1000 kPa) and temperatures (35 to 80  C). All data presented in
PEO400 (P1), with an isolated yield of 95% was obtained. this work were collected from at least three TFC membranes.
[PEO400]0/[TCL]0 ¼ 16/15; MnTheo ¼ 9.4 kDa; MnGPC ¼ 9.7 kDa, PAN substrates that were pre-coated with the PDMS gutter layer
Mw/Mn ¼ 1.28. For multi-block HMA-PEO1k (P2) [PEO1k]0/[TCL]0
were also tested with the same gases at 35  C and a pressure of
¼ 10/9; MnTheo ¼ 11.8 kDa; Mn GPC ¼ 12.4 kDa, Mw/Mn ¼ 1.36. 1H
340 kPa to detect any leakages. The ow rate of the single gases
NMR (400 MHz, CDCl3, dH ppm): 8.04–8.13 (m, 4H, ArH), 4.41–
was recorded manually with a digital ow meter (Agilent
4.51 (s, 2H, –CH2O(C]O)–), 3.50–3.74 (m, 4H, –CH2CH2O–).
Technologies ADM 2000).
The mass transport in non-porous polymeric membranes
2.4 Membrane preparation follows the solution-diffusion mechanism that has been well
documented in the literature.26 The ux (J) of a single gas A can
The cross-linked PDMS gutter layer was coated onto the poly-
be calculated from eqn (1):
acrylonitrile (PAN) microporous substrate as shown in Scheme
 
S1.† A typical example can be described as follows. In separate Dp
JA ¼ PA (1)
vials, TMC (0.007 g, 0.0267 mmol) was dissolved in 0.35 mL of l
hexane (2.0% w/v) and NH2–PDMS–NH2 (0.2 g, 0.04 mmol, 1
where PA is the permeability of gas A in Barrer, Dp is the pres-
equiv.) was dissolved in 10 mL of hexane (2.0% w/v). The two
sure difference across the membrane in bar and l is the
solutions were then mixed for 30 s and 1 mL of the solution was
membrane thickness in mm.
then spin-coated (1k rpm, 10 s) onto each PAN substrate
The membrane permeance is dened as the permeability
(19.63 cm2) to prepare 10 substrates coated with the cross-
divided by the membrane thickness and has a unit of GPU. This
linked PDMS gutter layer. The coated PAN substrates were dried
permeance can also be expressed in terms of total resistance to
in vacuo (1 mbar) at 25  C overnight and tested for their gas
ow (RT) by
transport properties to ensure that there were no leakages  
before coating the next layer, the selective layer. PA Dp
JA ¼ Dp ¼ (2)
Pebax 2533 was dissolved in a mixture of IPA/n-BuOH (3 : 1 l RT
vol.%) at 80  C for 24 h to prepare the polymer solution (2.0 wt%
In turn, the total resistance (RT) to permeation through TFC
of Pebax 2533). Aer cooling down the solution to room
membranes can be expressed as the sum of the resistances from
temperature, different amounts of the multi-block HMA-PEO (P1
the feed side boundary layer (RF), the permeate side boundary
and P2) were added (0 to 66 wt% relative to the amount of Pebax
layer (RP), the membrane selective layer (RSL) and the gutter
2533) and stirred at 50  C for 30 min. The IPA/n-BuOH is a binary
layer (RG) coated on the microporous substrate:
solvent which dissolves Pebax 2533 and HMA-PEOs but does not
damage the PDMS gutter layer or the microporous substrate. RT ¼ RF + RP + RSL + RG (3)
1.5 mL of the mixture was then spin coated (1.5k rpm, 20 s) onto
the substrate that had been pre-coated with the PDMS gutter layer. The boundary layer resistances (RF and RP) arise from
The TFC blend membranes were then dried in vacuo (1 mbar) at concentration gradients that are formed at the surface of the
25  C for 24 h and tested for their gas transport properties.

Fig. 1 Schematic diagram of the apparatus for measurement of gas flow rate.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 13769–13778 | 13771
View Article Online

Journal of Materials Chemistry A Paper

membrane in mixed gas systems. However, such concentration 1000 Da, respectively) were employed, allowing the PEO
polarization is not possible for single gas permeation. Thus the contents of the multi-block HMA-PEOs to be varied. The resul-
total resistance to ow is related only to the respective thick- tant HMA-PEOs (P1 and P2) were characterized by GPC for their
nesses of the selective and gutter layers and their respective molecular weights (MnGPC) assuming 100% mass recovery of
permeability:28 polymers (Table 1). The chemical structure of P1 and P2 was
lSL lG characterized by 1H NMR analysis. The molar ratio of the PEO
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

RT ¼ RSL þ RG ¼ þ (4) segment to the TCL unit (l) was determined by the integral ratio
PSL PG
of the methylene protons (Hc, dH 4.41–4.51 ppm) adjacent to the
The resistance to ow through the gutter layer (RG ¼ lG/PG) ester group and TCL aromatic protons (Ha, dH 8.04–8.13 ppm)
can be determined by measuring the ux through this layer (Fig. 2 and Table 1). The l values determined by 1H NMR
prior to deposition of the selective layer. Eqn (2) and (4) can analysis were slightly higher than the theoretical value from the
then be used to determine the permeance and permeability feed; the relative difference was reasonably small and within the
through the selective layer, by measuring the ow through the error margin. Based on the 1H NMR spectroscopic analysis, P2
combined TFC membrane. However, it should be noted that has more EO units per polymer chain than P1.
the estimation of the permeability is highly dependent on the
accuracy in determining the thickness of the selective layer. 3.2 Preparation of the cross-linked PDMS gutter layer
The ideal single gas selectivity (aA/B) between two gases A and
The PDMS gutter layer was prepared by spin coating the cross-
B can be expressed by the following equation:
linkable PDMS solution onto the PAN substrate. The successful
PA cross-linking of the amino groups was conrmed by XPS
aA=B ¼ (5)
PB measurements. The deconvoluted C 1s, O 1s, N 1s and Si 2p
spectra (Fig. S1, ESI†) provided detailed information about
surface functional groups presenting on PDMS, of which all are
3. Results and discussion in good agreement with the previous literature.29 In order to
estimate the thickness of the PDMS layer, SEM analysis was
In the present study, a commercially available PAN microporous
carried out on a number of different samples and one image
substrate was employed as a porous support to construct the
example is shown in Fig. 3A. The thickness of the PDMS coating
thin lm composite (TFC) blend membrane. The solution of
is less than 200 nm. The gas separation performance of the
amino-terminated PDMS and TMC was spin-coated onto the
highly permeable PDMS gutter layer was tested using the
PAN substrate to form a cross-linked gutter layer with high
apparatus shown in Fig. 1. The average CO2 permeance reached
permeability to CO2 (Scheme 1, step I). This PDMS intermediate
up to 2000  100 GPU and the CO2/N2 selectivity was approxi-
layer could prevent the penetration of diluted polymer solution
mately 9, which are in agreement with the previously published
into the PAN structure as well as render the entire membrane
data30 and demonstrate the successful formation of a PDMS
surface smoother. Thereaer, the pre-synthesized HMA-PEO
gutter layer without any leakage.
was blended into the Pebax 2533 matrix and the mixture was
then spin-coated onto the PDMS gutter layer to form the TFC
3.3 Formation and characterization of the selective layer
blend membrane (step II). Its ability to selectively separate CO2
from N2 was tested by utilizing an in-house built constant Pebax 2533 and Pebax 2533/HMA-PEO TFC blend
pressure variable volume (CPVV) apparatus (see Fig. 1). membranes (P1-25, P1-33, P1-50 and P1-66 in Table 2) were
prepared by spin-coating respective polymer solutions onto the
PAN substrates which were pre-coated with the PDMS gutter
3.1 Synthesis of the macromolecular precursor
layer. The cross-sectional morphologies of the Pebax 2533 and
The multi-block HMA-PEO copolymers were prepared via P1-50 TFC blend (50 wt% of HMA-PEO) membranes observed by
condensation polymerization of PEOs with TCL, which acts as a SEM analysis are illustrated in Fig. 3B and C, respectively. The
spacer, at room temperature (Scheme 2). PEO crystallinity, thickness of the selective layer was calculated by subtracting the
which is typically observed in high molecular weight PEO, is thickness of the PDMS gutter layer and it was estimated to be
suppressed by the introduction of the terephthalic spacer. The approximately 420 nm in all cases. The chemical changes that
low molecular weight PEO block provides exibility which occurred as a result of the coating of the selective layer on the
enables homogeneous dispersion in the Pebax matrix. Two PDMS gutter layer were further ascertained by X-ray photo-
types of PEOs with different molecular weights (Mn ¼ 400 and electron spectroscopy (XPS) analysis. Fig. S2 (ESI†) shows the
high resolution XPS spectra of Pebax 2533 (a, c and e) and P1-
50 (b, d and f). The C 1s and O 1s signals have been altered aer
the coating of the selective layer. For example, new chemical
bonds from the Pebax/HMA-PEO coating were detected. The C
1s spectra signal at 284.8 eV is attributed to the carbon in the
C–C bond, the signal at 286.4 eV is attributed to the carbon
Scheme 2 The synthesis of HMA-PEO copolymers via condensation ether bond C–O, the signal at 287.3 eV is attributed to the
polymerization. carbonyl bond C]O, and the signal at 289.1 eV is attributed to

13772 | J. Mater. Chem. A, 2013, 1, 13769–13778 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry A

Table 1 Summary of characterization results of multi-block HMA-PEOs P1 and P2

Sample Yield (%) MnTheoa (kDa) MnGPCb (kDa) Mw/Mnb lc (mPEO/mTCL)

P1 95 9.4 9.7 1.28 1.07


P2 92 11.8 12.4 1.36 1.12
a
Determined by feed ratio. b Determined by GPC. c l represents the molar ratio of the PEO segment to the TCL unit obtained by 1H NMR data using
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

Ic
the equation l ¼ , where Ia and Ic represent the integral areas of the resonance “a” for aromatic protons of TCL and the resonance “c” for
Ia
methylene group protons (–CH2O(C]O)–) of PEO, respectively.

Pebax 2533) show similar XRD patterns to Pebax 2533 and


exhibit no PEO crystalline peaks around 19.2 and 23.6 , indi-
cating a high degree of blending and strong interaction (e.g.
hydrogen bonding and entanglement) between the PEO and
Pebax chain. The XRD investigation provided strong evidence
for the successful suppression of PEO crystallinity by the
condensation polymerization of PEO400. This also indicates that
the crystalline phase of PEO in P2 was eliminated in the blend
system.
It is known that one drawback of using pure PEO in separation
Fig. 2 1
H NMR spectrum of the multi-block HMA-PEO copolymer P1 in CDCl3. membranes is its strong tendency to crystallize. Previous litera-
ture14 has reported a number of techniques to reduce the crys-
tallinity of PEO, such as utilizing low molecular weight liquid
the carbon in the N–C]O bond. The O 1s spectra signal at PEO or designing phase separated block copolymers. In the
533.0 eV is due to the oxygen in the C–O bond, and the signals at present work, we have successfully developed a simple approach
531.2 eV and 534.0 eV are assigned to the oxygen of C]O and to suppress the crystallization of PEO by condensation polymer-
O–C]O bonds, respectively. In the Pebax selective layer, the ization of low molecular weight liquid PEO and TCL.
intensity ratio of C–O to C–C signal is 0.73, while the intensity
ratio in the TFC blend membranes, e.g. in P1-50, is 0.85. This 3.4 Gas transport properties
indicates that the selective layer of P1-50 contains a larger All the gas transport properties of the TFC blend membranes as
proportion of EO moieties, and thus it is expected to have a function of the HMA-PEO mass fraction (0–66 wt%) have been
higher CO2 permeance. measured at a temperature of 35  C and a feed pressure of
The crystallinity of the multi-block HMA-PEO copolymers 340 kPa. Fig. 5 shows the CO2 gas permeance and CO2/N2
and the TFC membranes was investigated by X-ray diffraction selectivity for the TFC blend membranes up to 66 wt% HMA-
(XRD) as shown in Fig. 4. The XRD pattern obtained for the PEO content relative to Pebax 2533.
Pebax 2533 membrane shows three crystalline peaks, which As shown in Table 2, the CO2 permeance of these TFC blend
are attributed to the crystallization of the PA (2q ¼ 22.9 ) and PE membranes strongly increases with the increase in the HMA-
(2q ¼ 17.8 and 26.0 ) segments, respectively. Conversely, the PEO mass fraction. We hypothesize that the signicant increase
HMA-PEO copolymer P1 consisted of tens of liquid PEO400 of gas permeance can be attributed to three separate effects.
segments and presented as an amorphous phase. Thus its XRD Firstly, the increase of the HMA-PEO mass fraction results in a
pattern shows a broad peak around 22.0 . On the other hand, P2 higher overall concentration of EO moieties in the selective
displayed a strong tendency to crystallize due to the crystalli- layer. This in turn increases the CO2 solubility for permeation
zation of its repeat unit, i.e. PEO1k. Its XRD pattern shows a due to the plasticizing effect of the additive.21 Secondly, the
typical PEO crystalline peak at 2q ¼ 19.2 and 23.6 .31 The P1-50 terephthalic group, which acts as a spacer in the HMA-PEO
and P2-50 TFC blend membranes (P1 or P2 50 wt% blend with polymer, results in a reduction of the PEO crystallinity. Thirdly,

Fig. 3 SEM images of the cross-section of the membranes: (A) PDMS gutter layer, (B) Pebax 2533 TFC membrane and (C) P1-50 TFC blend membrane.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 13769–13778 | 13773
View Article Online

Journal of Materials Chemistry A Paper

Table 2 Gas separation performance of Pebax 2533/HMA-PEO TFC blend membranes and their selective layers tested at 35  C and 340 kPa

TFC blend membranesa Selectivity layerb

PEO content CO2 permeance CO2 permeability


Sample (%) CO2 permeance (GPU) CO2/N2 selectivity (GPU) (Barrer) CO2/N2 selectivity lc (nm)
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

Pebax 2533 0 305 23 373 157 26 400


P1-25 25 443 24 602 253 28 430
P1-33 33 574 23 914 384 31 440
P1-50 50 899 24 1850 778 40 420
P1-66 66 1130 14 — — — —
P2-25 25 479 23 685 288 29 420
P2-33 33 675 23 1080 452 31 420
P2-50 50 1070 22 2290 960 38 410
P2-66 66 1190 14 — — — —
a
The gas separation permeance of the TFC blend membranes. b The CO2 permeance, permeability and CO2/N2 selectivity of the selectivity layers. c l
is the thickness of the selective layer determined by SEM measurements.

Fig. 4 X-ray diffraction diagrams of Pebax 2533, P1, P1-50, P2 and P2-50.

the HMA-PEO polymer could be readily dispersed in the PA


phase improving the CO2-philic nature of this phase and
increasing the inter-chain distance, which greatly enhanced the
CO2 permeance. The TFC blend membranes containing
between 0 and 50 wt% HMA-PEO copolymers presented a 3 fold Fig. 5 CO2 permeance and CO2/N2 selectivity of the TFC blend membranes as a
increase in CO2 permeance while maintaining a constant CO2/ function of HMA-PEO copolymer (P1 and P2) content tested at 35  C and 340 kPa.
N2 selectivity of ca. 23. However, the TFC blend membranes
prepared with 66 wt% HMA-PEOs showed lower CO2/N2 selec-
tivity of ca. 14 with higher CO2 permeance. We believe that this This result is attributed to both structural (increase of EO
loss in selectivity resulted from the lack of mechanical stability concentration) and morphological (decrease of crystallinity)
of the selective layer for gas permeation experiments. changes in the selective layer matrix.
These results also indicate that there is an optimal content of
the HMA-PEO mass fraction (50 wt%) to fabricate TFC
membranes (P1-50 and P2-50) with desirable CO2 separation 3.5 Thermal properties
performance. In addition, the P2 series of TFC blend The effect of gas temperature on membrane performance was
membranes presented higher CO2 permeance than their P1 studied in the temperature range of 35–80  C and at a feed
counterparts of 50 wt% HMA-PEO blends. The CO2 permeance pressure of 340 kPa. The CO2 permeance for Pebax and P1-50
increases from 305 to 1070 GPU, ca. 3.5-fold increase, without TFC blend membranes versus the experimental temperature is
loss of selectivity. This is mainly attributed to the higher EO presented in Fig. 6 and Table 3. Higher temperatures can
concentration (see Table 1) in the P2 series, which further enhance the gas permeance due to the increase in polymer
enhances the CO2 solubility of the membrane. The incorpora- chain mobility and hence the increase in gas diffusivity.32 The
tion of EO units as HMA-PEOs into the Pebax 2533 matrix CO2 permeance was increased by 330 GPU in the pure Pebax
clearly resulted in an improvement of gas transport properties. TFC membrane and 420 GPU in the P1-50 TFC blend membrane

13774 | J. Mater. Chem. A, 2013, 1, 13769–13778 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry A

increasing feed gas pressure. The increasing permeance and


decreasing selectivity are due to CO2-induced plasticization.34
That is, CO2 molecules cause the polymer matrix to swell,
leading to increased FFV and hence increased diffusivity of
gases. However, the change in diffusivity is greater for the larger
gas molecule (N2) resulting in a reduction in the selectivity.
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

Although plasticization occurred at high feed pressures, the


CO2/N2 selectivity of the TFC blend membrane still maintained
a value of approximately 21 at 1000 kPa. As shown in Fig. 7
(bottom), the change in selectivity of the P1-50 TFC blend
membrane is similar to that of Pebax 2533, which indicates
that the plasticization effect is mainly attributed to the plasti-
cization of the Pebax copolymer and the multi-block HMA-PEO
additive did not induce any further plasticization of the selec-
tive layer.

3.7 Gas permeance and permselectivity of the selective layer


The permeance and selectivity through the selective layer can be
determined by deducting the transport resistances of the gutter
layer. In this study, each of the PAN substrates pre-coated with
the cross-linked PDMS gutter layer was tested for its gas trans-
Fig. 6 CO2 and N2 permeance and CO2/N2 selectivity of the Pebax 2533 and port properties. The average CO2/N2 selectivity was 9. The ow
P1-50 TFC blend membranes as a function of temperature tested at 340 kPa. resistance through this gutter layer (RG) was determined by
dividing the pressure difference applied across the membrane
by the average ux of CO2 and N2 (eqn (4)). The ow resistance
when the operating temperature was increased from 35 to 80  C. through the selective layer could then be determined by sub-
This behavior is typical of non-porous polymeric membranes as tracting this resistance from the total resistance calculated
higher gas ux can be recorded with higher temperature.33 using the same approach with the TFC blend membrane. The
Although the CO2 permeance was enhanced with increasing thickness of the selective layers was estimated by SEM analysis
temperature, the CO2/N2 selectivity of both the Pebax and and the values were in the range of 400–440 nm (Table 2). The
P1-50 TFC blend membranes decreased dramatically at high CO2 permeance and permeability of the selective layer were then
temperatures. These effects are typical of polymeric calculated and are shown in Fig. 8.
membranes: the reduction in CO2/N2 selectivity is due to the Fig. 8 shows the CO2 permeability, permeance and CO2/N2
signicant decrease of both diffusivity and solubility selectivity of the selective layer in the fabricated TFC blend
selectivity.24 membranes. Both the CO2 permeability and permeance are
enhanced up to six-fold by increasing the HMA-PEO mass
fraction relative to Pebax 2533 (Fig. 8A and B). For example, the
3.6 Pressure dependence CO2 permeability of the selective layer is increased from 156 to
As described in the experimental section, single gas (CO2 and 788 Barrer for the P1 series and to 960 Barrer for the P2 series.
N2) measurements were also carried out across a range of feed The permeability value of the selective layer, P1-50SL (or P2-
pressures at 35  C. Fig. 7 and Table 4 present the CO2 per- 50SL), is much higher than that reported for either cross-linked
meance and CO2/N2 selectivity as a function of the feed gas PEO (P ¼ 150–270 Barrer)35 or Pebax 2533 (P ¼ 160–180
pressure. The CO2 permeance was increased by ca. 120 GPU in Barrer).36 This phenomenon can be assigned to the increase in
Pebax and by 200 GPU in the P1-50 TFC blend membrane the proportion of EO moieties, the increase of interaction
when the feed pressure was increased from 340 to 1000 kPa. The between HMA-PEO and Pebax chains, and the decrease of
CO2/N2 selectivity of Pebax and P1-50 slightly decreased with crystallinity in the selective layer matrix. In addition, the CO2/N2

Table 3 Gas separation performance of Pebax 2533 and P1-50 TFC blend membranes tested at 340 kPa

Pebax 2533 P1-50

Temperature ( C) CO2 (GPU) CO2/N2 selectivity CO2 (GPU) CO2/N2 selectivity

35 329 25 898 25
50 430 18 986 18
65 563 14 1170 13
80 662 11 1320 11

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 13769–13778 | 13775
View Article Online

Journal of Materials Chemistry A Paper


Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

Fig. 7 CO2 and N2 permeance and CO2/N2 selectivity of the Pebax 2533 and
P1-50 TFC blend membranes as a function of feed gas pressure.

selectivity increased from 25 to 40 (Fig. 8C) with the blending of


the HMA-PEO copolymer up to 50 wt% relative to Pebax 2533 Fig. 8 CO2 permeability (A), permeance (B), and CO2/N2 selectivity (C) of the
due to the increase of CO2 solubility selectivity in the selective selectivity layer as a function of HMA-PEO copolymer (P1 and P2) content tested
layer. These unprecedented results, the concurrent increase in at 35  C and 340 kPa.
CO2 permeability and CO2/N2 selectivity, present a new direc-
tion for the fabrication of polymeric membranes that surpass
the upper performance limit. of the selective layer. However, the Pebax/HMA-PEO selective
The CO2 permeability of some polymeric membranes layer has less FFV when compared to the TR and TZPIM
reported in the literature is indicated in Fig. 9.37–39 Fig. 9 also membranes, which results in lower CO2 diffusivity and hence
shows the relationship between CO2 permeability and CO2/N2 lower CO2 permeability. Nevertheless, the high CO2 perme-
selectivity of the selective layer studied in this work when ability (778–960 Barrer) and the reasonable CO2/N2 selectivity
compared to the data of Robeson's 2008 plots.40 Both the (40–38) for P1-50 and P2-50, respectively, make these polymeric
selective layers P1-50SL and P2-50SL demonstrate exceptional gas blends remarkable materials for CO2 separation membranes.
separation performance, surpassing the most recent Robeson Fig. 10 shows a trade-off plot of CO2/N2 selectivity versus CO2
upper bounds of conventional and state-of-the-art polymeric permeance.8,12 The target area for post combustion capture
membranes for CO2/N2 separation. The P1-50SL and P2-50SL also proposed by Merkel et al.8 is also included in the plot for
present higher CO2/N2 selectivity in comparison with the ther- comparison. Membranes require high CO2 permeance ($1000
mally rearranged (TR) polymeric membranes38,41 (hexagon) and GPU) and modest selectivity ($20) to fall in this target perfor-
membranes derived from tetrazole modied polymers of mance area. Membranes with lower CO2 permeances will
intrinsic microporosity (TZPIMs)37 (circle). In the TFC blend require a large membrane area. In addition, membranes with a
membranes, the addition of the HMA-PEO copolymer improved selectivity of less than 20 are unlikely to be selective enough,
the CO2 solubility and hence increased the solubility selectivity regardless of their permeance. On the other hand, membranes

Table 4 Gas separation performance of Pebax 2533 and P1-50 TFC blend membranes tested at 35  C

Pebax 2533 P1-50

Pressure (kPa) CO2 (GPU) CO2/N2 selectivity CO2 (GPU) CO2/N2 selectivity

340 329 25 898 25


500 364 22 961 22
700 404 22 1030 21
1000 458 21 1080 21

13776 | J. Mater. Chem. A, 2013, 1, 13769–13778 This journal is ª The Royal Society of Chemistry 2013
View Article Online

Paper Journal of Materials Chemistry A

performance was close to the target performance area.


Furthermore, the selective layers P1-50SL and P2-50SL displayed
much higher CO2 permeance (ca. 2000 GPU) with a modest
selectivity (ca. 40). This makes the Pebax 2533/HMA-PEO
copolymer mixture in this work an attractive material with
optimum membrane properties for efficient CO2 separation.
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

It should be pointed out that although the total gas separa-


tion performance of these TFC blend membranes only falls at
the edge of the target area, better separation performance could
be obtained by reducing the thickness of the PDMS gutter layer
or the selective layer. This is an area of our further research.

4. Conclusion
Fig. 9 Relationship between CO2 permeability and CO2/N2 selectivity of the
In summary we report a facile approach to prepare Pebax 2533/
selective layer tested in this study.
HMA-PEO TFC blend membranes for CO2/N2 separation via
spin-coating the mixture onto a PAN microporous substrate that
had been pre-coated with a cross-linked PDMS gutter layer. By
with a selectivity of much more than 100 are likely to be too
synthesizing a multi-block HMA-PEO copolymer with a tereph-
selective and require a large membrane area because of the
thalic spacer and using it as an additive, the CO2 permeance was
pressure-ratio-issues.
signicantly increased while maintaining the CO2/N2 selectivity
As shown in Fig. 10, the P1-50SL and P2-50SL selectivity layers
in the Pebax 2533 based TFC blend membranes. Their ability
have modest selectivity but higher CO2 permeance than the
to separate CO2 from N2 was studied under variable conditions:
Polaris membrane (patent protected, no details released),
at temperatures of 35–80  C and pressures of 340–1000 kPa. The
which in turn is 10–50 times more permeable than conventional
highest CO2 permeance with reasonable CO2/N2 selectivity was
cellulose acetate membranes used for CO2 removal from natural
found in Pebax 2533/HMA-PEO TFC blend membranes with a
gas. In comparison with such conventional cellulose acetate
50 wt% HMA-PEO mass fraction. The increased CO2 permeance
membranes, the Pebax 2533 TFC membrane and its selective
was mainly attributed to the increase in the CO2-philic charac-
layer (Pebax 2533SL) presented a higher CO2 permeance with
teristic as a result of the introduction of more EO moieties
slightly lower CO2/N2 selectivity. The P1-50 and P2-50 TFC blend
and the increase in FFV as a result of an increase in the Pebax
membranes showed higher CO2 permeance again (ca. 1000
inter-chain distance. In addition, the CO2 permeability of the
GPU) with a CO2/N2 selectivity of $20. Their gas separation
selective layer was increased ve-fold and the CO2/N2 selectivity
was increased to 40. This performance of P1-50SL and P2-50SL
successfully surpassed the most recent upper bounds.
Furthermore, P1-50SL and P2-50SL fell within the CO2 separation
performance targets (permeance $1000 GPU and CO2/N2
selectivity $20) with a CO2 permeance of 1800–2200 GPU and a
modest CO2/N2 selectivity, which identies them as promising
materials for efficient CO2 separation.

Acknowledgements
The authors acknowledge the funding provided by the Austra-
lian Government through the CRC program to support this
research. The authors also acknowledge the Australian
Research Council under the Future Fellowship (FT110100411,
G.G.Q.). Q. Fu is the recipient of an Australian Research Council
Super Science Fellowship (FS110200025) and the Australia-
Fig. 10 CO2/N2 selectivity versus CO2 permeance plot comparing the perfor- China Emerging Future Leaders in Low Emissions Coal Tech-
mance of Pebax 2533, P1-50 and P2-50 TFC blend membranes and their
nology Fellowship. Q. Fu also wishes to acknowledge the
selective layers Pebax 2533SL, P1-50SL and P2-50SL with commercial natural gas
membranes, the Polaris membrane and developmental membranes reported in
University of Melbourne's 2013 Early Career Researcher Grant
the literature. The target area is that proposed by Merkel et al.8 for post (Q. Fu).
combustion capture of carbon dioxide. Open circles (B) are data points for Kai
et al.42 and Duan et al.;43 the filled triangle (=) represents data from Bao and
Trachtenberg;44 the filled diamond (A) from Deng et al.;45 the open square (,) References
from Cai et al.;46 the open diamond (>) from Hanioka et al.;47 the filled star (+)
from Zhao et al.;48 and the inverted triangles (;) from Hendriks et al.49 1 C. E. Powell and G. G. Qiao, J. Membr. Sci., 2006, 279, 1–49.

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 13769–13778 | 13777
View Article Online

Journal of Materials Chemistry A Paper

2 M. G. Buonomenna, W. Yave and G. Golemme, RSC Adv., 27 D. Husken, J. Feijen and R. J. Gaymans, J. Polym. Sci., Part A:
2012, 2, 10745–10773. Polym. Chem., 2007, 45, 4522–4535.
3 W. Yave, A. Car, J. Wind and K.-V. Peinemann, 28 R. W. Baker, Ind. Eng. Chem. Res., 2002, 41, 1393–1411.
Nanotechnology, 2010, 21, 395301–395307. 29 J. Liang, L. He, X. Dong and T. Zhou, J. Colloid Interface Sci.,
4 S. Li, Z. Wang, X. Yu, J. Wang and S. Wang, Adv. Mater., 2012, 2012, 369, 435–441.
24, 3196–3200. 30 T. C. Merkel, V. I. Bondar, K. Nagai, B. D. Freeman and
Published on 16 September 2013. Downloaded by The University of Melbourne Libraries on 03/02/2015 05:16:22.

5 S. A. Lee, G. W. Stevens and S. E. Kentish, J. Membr. Sci., 2013, I. Pinnau, J. Polym. Sci., Part B: Polym. Phys., 2000, 38, 415–
429, 349–354. 434.
6 S. Li, Z. Wang, C. Zhang, M. Wang, F. Yuan, J. Wang and 31 Q. Wang, Z. Dong, Y. Dua and J. F. Kennedy, Carbohydr.
S. Wang, J. Membr. Sci., 2013, 436, 121–131. Polym., 2007, 69, 366.
7 W. Yave, A. Car, S. S. Funari, S. P. Nunes and 32 S. Kelman, H. Lin, E. S. Sanders and B. D. Freeman, J. Membr.
K.-V. Peinemann, Macromolecules, 2010, 43, 326–333. Sci., 2007, 305, 57–68.
8 T. C. Merkel, H. Lin, X. Wei and R. Baker, J. Membr. Sci., 2010, 33 L. Liu, A. Chakma and X. Feng, Ind. Eng. Chem. Res., 2005, 44,
359, 126–139. 6874–6882.
9 S. Zhao, Z. Wang, Z. Qiao, X. Wei, C. Zhang, J. Wang and 34 A. Bos, I. G. M. Pünt, M. Wessling and H. Strathmann,
S. Wang, J. Mater. Chem. A, 2013, 1, 246–249. J. Membr. Sci., 199, 155, 67–78.
10 F. d. Clippel, A. L. Khan, A. Cano-Odena, M. Dusselier, 35 H. Lin, T. Kai, B. D. Freeman, S. Kalakkunnath and
K. Vanherck, L. Peng, S. Oswald, L. Giebeler, S. Corthals, D. S. Kalika, Macromolecules, 2005, 38, 8381–8393.
B. Kenens, J. F. M. Denayer, P. A. Jacobs, 36 C. A. Scholes, J. Bacus, G. Q. Chen, W. X. Tao, G. Li, A. Qader,
I. F. J. Vankelecom and B. F. Sels, J. Mater. Chem. A, 2013, G. W. Stevens and S. E. Kentish, J. Membr. Sci., 2012, 389,
1, 945–953. 470–477.
11 K. A. Lundy and I. Cabasso, Ind. Eng. Chem. Res., 1989, 28, 37 N. Du, H. B. Park, G. P. Robertson, M. M. Dal-Cin,
742–756. T. Visser, L. Scoles and M. D. Guiver, Nat. Mater., 2011,
12 F. Yuan, Z. Wang, S. Li, J. Wang and S. Wang, J. Membr. Sci., 10, 372–375.
2012, 421–422, 327–341. 38 H. B. Park, S. H. Han, C. H. Jung, Y. M. Lee and A. J. Hill,
13 Y. Hirayama, Y. Kase, N. Tanihara, Y. Sumiyama, Y. Kusuki J. Membr. Sci., 2010, 359, 11–24.
and K. Haraya, J. Membr. Sci., 1999, 160, 87–99. 39 L. Zhao, E. Riensche, R. Menzer, L. Blum and D. Stolten,
14 H. Lin and B. D. Freeman, J. Mol. Struct., 2005, 739, 57–74. J. Membr. Sci., 2008, 325, 284–294.
15 H. Lin, E. V. Wagner, B. D. Freeman, L. G. Toy and 40 L. M. Robeson, J. Membr. Sci., 2008, 320, 390–400.
R. P. Gupta, Science, 2006, 311, 639–642. 41 H. B. Park, C. H. Jung, Y. M. Lee, A. J. Hill, S. J. Pas,
16 H. Cong and B. Yu, Ind. Eng. Chem. Res., 2010, 49, 9363–9369. S. T. Mudie, E. V. Wagner, B. D. Freeman and
17 B. Xue, X. Li, L. Gao, M. Gao, Y. Wang and L. Jiang, J. Mater. D. J. Cookson, Science, 2007, 318, 254–258.
Chem., 2012, 22, 10981. 42 T. Kai, T. Kouketsu, S. Duan, S. Kazama and K. Yamada, Sep.
18 J. Xia, S. Liu and T.-S. Chung, Macromolecules, 2011, 44, Purif. Technol., 2008, 63, 524–530.
7727–7736. 43 S. Duan, F. A. Chowdhury, T. Kai, S. Kazama and Y. Fujioka,
19 H. Lin and B. D. Freeman, J. Membr. Sci., 2004, 239, 105–117. Desalination, 2008, 234, 278–285.
20 A. Car, C. Stropnik, W. Yave and K.-V. Peinemann, Sep. Purif. 44 L. Bao and M. C. Trachtenberg, J. Membr. Sci., 2006, 280,
Technol., 2008, 62, 110–117. 330–334.
21 A. Car, C. Stropnik, W. Yave and K.-V. Peinemann, J. Membr. 45 T.-J. Kim, B. Li and M.-B. Hägg, J. Polym. Sci., Part B: Polym.
Sci., 2008, 307, 88–95. Phys., 2004, 42, 4326–4336.
22 J. H. Kim and Y. M. Lee, J. Membr. Sci., 2001, 193, 209–225. 46 Y. Cai, Z. Wang, C. Yi, Y. Bai, J. Wang and S. Wang, J. Membr.
23 L. Liu, A. Chakma and X. Feng, J. Membr. Sci., 2004, 235, 43– Sci., 2008, 310, 184–196.
52. 47 S. Hanioka, T. Maruyama, T. Sotani, M. Teramoto,
24 E. Toccia, A. Gugliuzzaa, L. D. Lorenzoa, M. Macchione, H. Matsuyama, K. Nakashima, M. Hanaki, F. Kubota and
G. D. Lucaa and E. Drioli, J. Membr. Sci., 2008, 323, 316–327. M. Goto, J. Membr. Sci., 2008, 314, 1–4.
25 S. Sridhar, R. Suryamurali, B. Smitha and T. M. Aminabhavi, 48 L. Zhao, R. Menzer, E. Riensche, L. Blum and D. Stolten,
Colloids Surf., A, 2007, 297, 267–274. Energy Proc., 2009, 1, 269–278.
26 S. R. Reijerkerk, M. H. Knoef, K. Nijmeijer and M. Wessling, 49 J. P. V. D. Sluus, C. A. Hendriks and K. Blok, Energy Convers.
J. Membr. Sci., 2010, 352, 126–135. Manage., 1992, 33, 429–436.

13778 | J. Mater. Chem. A, 2013, 1, 13769–13778 This journal is ª The Royal Society of Chemistry 2013

View publication stats

You might also like