http://www.chemistrymag.org/cji/2000/028036re.htm

  Aug.14, 2000  Vol.2 No.8 P.36 Copyright cij17logo.gif (917 bytes)


Polymeric membrane materials for CO2-selective separation (I)

Hao Jihua
(Department of Chemical Engineering, Tsinghua University, Beijing 100084, China )

Received Apr. 13,  2000.

Abstract Polymeric membrane materials for CO2-selective separation have been reviewed. On the basis of the introduction to solution-diffusion transport mechanism of gases in solid polymeric membranes, transport characteristics of CO2 were summarized. CO2-selective polymeric membrane can be divided into four types, including the chemical and physical modification of the traditional membrane materials, polymer blending, the design and synthesis of new membrane materials and organic-inorganic hybrid membrane. Finally, the research topics in future were also suggested.
Keywords membrane; carbon dioxide;

CO2-selective separation is a very important problem in many fields, such as environment, petrochemical engineering, agriculture and other relative industry.[1-16] One of the promising techniques is membrane processes. Because of the potential economy of the membrane technology, many governments have initiated and extensively focused in the projects on the CO2-selective polymeric membranes and membrane processes. Membranes and membrane processes for the separation of CO2 and H2S from natural gases have been a project under the "BRITE"(Basic Research in Industrial Technology for Europe) formed as a European Community R&D program in the middle of 1980s.[5] In Japan, membrane processes for CO2-selective separation has been an important project under C1 Chemistry "Sunshine" plan.[6] In China, membrane processes for CO2/CH4 separation have been studied because of the existence of several large sour natural gas fields and application in tertiary oil recovery.[7] A number of large American companies first applied the membrane technology for separation of CO2/CH4 and continued efforts on new membrane materials and membrane process design for CO2 separation.[8] The membranes for CO2-selective separation have made much progress during this period.
    Because earlier cellulose acetate membrane for reverse osmosis could be easily adapted to form dry asymmetric structures using solvent-exchange dehydration for gas separation, especially Monsanto invented "caulked membrane" successfully for recovery of hydrogen from ammonia purge gas, solid polymer-based membranes for gas separation have been paid attention to in the membrane field over the past twenty years. New polymeric membrane formation with integrally-skinned dense thin asymmetric structure have been the main research topics.
    There have been two kinds of polymeric membrane for CO2-selective separation processes. One is the solid polymeric membrane based on the solution-diffusion mechanism, another is the facilitated transport membrane based on the solution-reaction-diffusion mechanism.[17]
     In this review we limited only to membrane materials based on the solution-diffusion mechanism.

1 TRANSPORT MECHANISM OF GASES IN SOLID POLYMERIC MEMBRANES AND CHARACTERISTICS OF CO2
The permeation of a gas molecule through a dense polymer membrane proceeds by a solution-diffusion mechanism. In this model, gas molecules first dissolve into the high pressure face of the membrane, diffuse across the membrane to the low pressure side, and desorb from the face.[18, 19]
    The permeability coefficient P of a membrane in solution-diffusion processes can be represented as the product of a diffusivity [D] and solubility [S] coefficient as shown in the following equation.
P=[D][S]
The ideal selectivity can be expressed as

αAB=PA/PB=[DA/DB][SA/SB]
Where, [DA/DB]---------the diffusivity selectivity
             [SA/SB]----------the solubility selectivity
    In addition to operating conditions (i.e. temperature, pressure and composition), gas solubility also depends on factors such as penetrant interactions, and polymer morphology(i.e. crystallinity, orientation, etc.).[20] Gas diffusivity depends on the ability of small gas molecules to undergo diffusive jumps which occur when thermally driven, random, cooperative polymer segmental dynamics from transient gaps, large enough to accommodate penetrants, in the immediate vicinity of the gas molecules.[21] Like solubility, diffusivity depends on membrane operating conditions; moreover, gas diffusivity is also sensitive to properties such as penetrant size, polymer morphology and polymer segmental dynamics.

2 CHARACTERISTICS OF CO2 AND CH4
2.1 Gas condensibility

In most polymers, CO2(Tc=31°C) is more soluble than CH4 (Tc=-82.1°C). Gas critical temperature, Tc, normal boiling point, Tb, and Lennard-Jones force constant are all measures of condensibility which correlate well with the solubility coefficients of a range of penetrants in polymers. Gas solubility in polymers generally increases with increasing gas condensibility.[22, 23]
2.2 Penetrant size and shape 
Diffusion coefficients in polymers are sensitive to penetrant size and shape. The diffusivity of linear or oblong penetrant molecules such as CO2 is higher than the diffusivities of spherical molecules of equivalent molecular volume. The van der Waals volumes of CO2 and CH4 are estimated to be 17.5 and 17.2 cm3/mole.[22, 24] These molecular volumes yield equivalent spherical diameters of 3.33 and 3.31Å for CO2 and CH4, respectively. But the diffusion coefficient of CO2 is typically greater than that of CH4, despite the larger van der Waals volume of CO2. The kinetic diameter of the asymmetric molecule CO2 (3.30Å) is smaller than that of methane, a symmetric molecule (3.80Å). Based on such results, the transport of small, asymmetric molecules is understood to proceed with diffusion jumps of the penetrant through a polymer occurring principally parallel to the long axis of the penetrant.
02803601.gif (4544 bytes)
Fig. 1. The effect of polar carbonyl and sulfone group concentration on CO2/CH4 solubility[25]

    Most glassy gas separation membrane materials achieve high permselectivity as a result of high diffusivity selectivity (i.e. by sieving penetrant molecules based on differences in molecular size).

2.3 Polarity   
Gas solubility is sensitive to specific interactions between gas and polymer molecules. Gases such as CO2, which has a quadrupole moment, are in general, more soluble in polar polymers. Many polymeric materials for CO2-selective membrane have been designed by use of this property. Van Amerongen determined that the solubility of quadrupolar CO2 increased substantially as acrylonitrile content increased.[25] Pilate suggested that CO2 interact favorably with the sulfone group in polysulfone via induced dipole-dipole interaction.[26] In this regard, Koros found a correlation between CO2/CH4 solubility selectivity and the concentration of polar carbonyl and sulfone groups in the medium in which the gases were dissolved.[27] These data are reproduced in Fig. 1.[25] In a related work, Chern and coworkers found that aryl nitration of polysulfone, which adds polar nitro groups to the polymer backbone, increases CO2/CH4.[28]
2.4 Pressure    
Permeability of CO2 in polymeric membrane is more sensitive to pressure. The selectivity of CO2/CH4 mixture in the most of the current polymeric membranes decreased at high pressure because of the membrane plasticization. This is one of the key problems to be solved for CO2-selective polymeric membrane.
    At very high gas pressure, permeability may increase with increasing pressure. This increase is related to penetrant plasticization of the polymer matrix and is primarily due to an increase in the penetrant diffusion coefficient. The data in Fig. 2 present plasticization in the presence of CO2.[22, 29] Other condensable gases and hydrocarbon vapors can also plasticize polymers. At sufficiently high pressure the penetrant acts as an efficient plasticizer in hydrophilic polymers such as polyvinyl alcohol. The permeability decreases with pressure initially, consistent with the prediction of dual mode model, and then increases when the gas begins to plasticize the polymer. Plasticization occurs when penetrant molecules dissolve in the polymer chain segmental motion. The increased segmental mobility may be observed by a depression of the glass transition temperature of the polymer-penetrant mixture. The increased segmental mobility provides more opportunities for penetrant molecules to execute diffusive jumps, and, in turn, the penetrant diffusion coefficient increases. Plasticization is undesirable for most gas separation membranes since the polymers used in these applications depend largely on diffusivity selectivity to achieve high overall permselectivity. The relatively large, uncontrolled polymer segmental motions associated with the plasticized or rubbery polymers decrease the diffusivity selectivity and, therefore, overall permselectivity of polymer.

02803602.gif (3030 bytes)
Fig. 2 CO2 induced plasticization of poly(tetrabromophenolphthalein) at 35°C[27]

3 POLYMERIC MEMBRANE MATERIALS FOR SEPARATION OF CO2/CH4     
Much effort has been spent to optimize CO2-selective polymeric membranes: including the chemical and physical modification of current membranes,[30, 31, 32] blending,[33, 34] the design and synthesis of new membrane material[35, 36] and organic-inorganic membrane.[37]
02803603.gif (6000 bytes)
Fig. 3
"Trade-off" curves for the CO2/CH4 separation system

    A typical "trade-off curve" between CO2/CH4 permselectivity and CO2 permeability is pressured in Fig. 3.[38] A crosshatched, undesirable area, below the solid line is associated with "low free volume glasses" and "rubbery polymers". The poor mobility selectivities and high diffusivities in rubbery media noted earlier produce high permeabilities and low overall permselectivities. On the other hand, the well-packed, low free volume glasses, give equally unattractively low permeabilities and high selectivities. The solids line (-.-) represents "optimistic" tradeoff behavior typical of commercially available engineering resins.
    During the past decade many investigators attempted to synthesize entirely new polymers that exhibit both a higher permeability and a higher selectivity than presently available polymers. Or they attempted to enhance the selectivity and/or permeability of available polymers by chemical or physical modification.
02803604.gif (4316 bytes)
Fig. 4
Influence of chemical modification

   Quite a lot of modification have been carried out to those polymers that are proven gas separation materials. [30, 31, 32] Stam et al. modified on-chain of polymers, like polysulfone (PSF), polyethersulfone (PES) and poly(phenylene oxide) (PPO) by substitution of aromatic protons, i.e. by means of nitration, silylation, bromination of alkyl groups and acrylation.[30] Some results are shown in Fig. 4. Even modifications with more complex chemicals have not shown the desired result.
    Kesting et al. improved the flux of polysulfone gas separation membranes by using Lewis acid-base pairs in the aqueous phase inversion bath. The resulting membranes possessed greater free volume and hence, higher flux, than those made conventionally.31 In addition, Stam put additives to the polymer membrane by physical modification. The most successful was the addition of molecular sieves(Na-Y). Addition of 1% by weight of molecular sieve to the polymer membrane has shown a relative improvement of 2% on the permeability without effecting the selectivity. At contents of 15% molecular sieve maximum permeability has been reached. At higher contents selectivity was reducing again, where as permeability was not improving any further.30 Barbari treated polysulfone membrane with molecular bromine. After bromine treatment, the selectivity of a polysulfone membrane for CO2 over CH4 was increased over 100% at 10 atm upstream pressure with only a 36% reduction in CO2 permeability.
    Although all types of chemical and physical modification showed that selectivity or permeability could be improved, even at the expense of permeability or selectivity, the selectivities or permeabilities themselves have not exhibited remarkably attractive.
    However, the tailor-made polymers have shown that it is possible to move away from the traditional trade-off curve connecting the conventional membrane materials (-
D- in Fig. 3).[38] Although considerable data and general correlations relating structure and permeability exist, there are no truly quantitative relationships to guide detailed structure-permeability optimization. Fortunately, qualitative rules have emerged that date back to the pioneering work by Hoeln and his colleagues at DuPont in the 1970's.[39, 40] Simply put, when changing the structure within a family of polymers, inhibiting intersegmental packing, while simultaneously hindering the backbone mobility tends to produce a desirable tradeoff between productivity and permselectivity changes. Currently this is the most reliable guide for understanding structure-mobility studies of a given family of polymers.
    Table 1 lists permeability coefficients for CO2 and CH4 in a variety of rubbery and glassy polymers, as well as the selectivities (ideal separation factors) of these polymers for CO2/CH4. The high selectivities of these tailor-made polymers, such as 6FDA-polyimides, polytriazoles, and polyoxadiazoles are assumed to originate from their rigid structure. Their high permeability is mostly explained in terms of their high free volume, which is due to their bulky groups preventing efficient packing. The solubility is mostly relative to the chemical composition of the polymers.[38]
    In the following section the excellent polymer materials with extra high selectivity and/or permeability will be introduced in detail.

Table 1. Permeabilities and selectivities of CO2/CH4 in rubbery and glassy polymers

 

Permeability coefficient, P×1010
[cm3(STP). cm/s. cm2. cmHg]
CO2                            CH4

Overall selectivity

CO2/CH4

Poly[1-(trimethylsilyl)-1-propyne
Poly(dimethylsiloxane)
Poly(dimethyl silmethylene)
Poly(vinyltrimethylsilane)
Poly(cis-isoprene)
Poly(butadiene-styrene)
Ethyl cellulose
Poly(phenylene oxide)
Polystyrene
Poly(ethyl methacrylate)
Bisphenol-A polycarbonate
Cellulose acetate
Bisphenol-A polysulfone
Polyetherimide(Ultem)(PEI)
PMDA-4,4'-ODA polyimide
Kapton polyimide
Poly(methyl methacrylate)
6FDA-m-DDA polyimide
poly(1,3,4-oxadiazole)(POD)
Polyaniline

28000
4550
520
200
191
171
75.0
61.0
12.4
7.01
6.8
4.75
4.6
1.33
2.7
0.22
0.62
1.8
0.3
0.32

13000
1430
150
14
47.4
34.2
6.8
4.3
0.78
0.35
0.36
0.15
0.18
0.036
0.059
0.0032
0.0052
0.013
0.0018
<0.002

2.2
3.2
3.5
14
4.0
5.0
11
14.2
16
20
19
32
26
36.9
46
68.8
119
136
168
336

3.1 Improvement of the selectivity of polymers with high permeability
3.1.1 Poly(organosiloxane)
Generally, rubbery polymers exhibit high permeabilities and low selectivity, in which silicone polymers, and in particular poly(dimethylsiloxane) (PDMS), has received considerable attention. The high permeability of PDMS, [-(CH3)2SiO-]n, has been attributed to its large free volume, which may be due to the flexibility of the siloxane (-SiO-) linkages of this polymer. Although the gas selectivity of PDMS is very low, the structure/permeability relationships of other silicone polymers have been systematically investigated in the last few years. These studies had the objective of finding membrane materials that exhibit higher gas selectivity than PDMS as well as high gas permeability.
    The substitution of increasingly bulkier functional groups in the backbone or side chains of PDMS, or the replacement of its flexible -SiO- linkages with stiffer -SiCH2- linkages, raises the glass-transition temperature, Tg, of the polymers.[21] This means the increase of chain-packing density. The effects of such substitution on the relationships between the gas permeability and selectivity of silicone polymers were listed in Table 2.

Table 2. Permeabilities and selectivities of CO2/CH4 in rubbery and glassy polymers

 

Permeability coefficient, P×1010
[cm3(STP). cm/s. cm2. cmHg]
CO2                      CH4

Overall selectivity

CO2/CH4

Side-chain modification
[(CH3)2SiO]n
(CH3C3H7SiO)n
(CH3C8H17SiO)n
[(CF3C2H4)CH3SiO]n
(C6H5CH3SiO)n
Backbone modifications
[(CH3)2Si(CH2)6Si(CH3)2O]n
[(CH3)2SiCH2]n
[(CH3)2Si-p-C6H4Si(CH3)2O]n


4570  
1520  
917  
1210  
226

1310  
542  
52.3  


1450  
531  
314  
201  
36.3

395  
130  
10.4


3.2  
3.0  
3.0  
6.9  
6.5

3.3  
4.2  
5.2

As shown in Table 2, the high CO2/CH4 selectivity of poly(trifluoropropyl methyl siloxane) [(CF3C2H4)CH3SiO]n is due to an anomalously high CO2 solubility, which results from specific interactions between this gas, whose molecules have a strong quadrupole moment, and the polar fluorine-containing groups in the polymer.[41, 42]
02803605.gif (4210 bytes)
Fig.5 Effect of differential pressure on the CO2/CH4 relative pure gas permeabilities in the diester functionalised membranes: diester functionality (a) 30 mol%; (b) 20mol%; (c) 10mol%;  (d) 0 mol%.
02803606.gif (2708 bytes)
Fig.6
CO2 permeabilities coefficient at 35C and atmospheric pressure versus composition of the CO2:CH4 permeant gas mixture in diester functionalised membranes: diester functionality (a) 0 mol%; (b) 10mol%; (c) 30mol%.

    In order to improve the permselectivity of rubbery polymers without unacceptable loss in overall permeability via improved solubility selectivity Ashworth et al. introduce a side-chain organofuntional groups, such as mono- or di-ester, into a PDMS matrix.[43, 44] The increase of selectivity results from the interaction of CO2 with the ester groups in the ester side-chains. The results have shown in Fig. 5, 6.
    In addition, the surface of poly(dimethylsiloxane) membrane was modified by plasma treatments using Ar, N2, O2 and NH3.[45] As a result, the selectivity of membrane has been remarkably improved by the plasma treatment. The selectivity and the permeation rate of CO2/CH4 obtained were 53 and 3.3×10-4 cm3/
s.cm2. cmHg.

3.1.2 Polyacetylenes

Table 3. Permeability coefficients and selectivities for CO2/CH4 in substituted polyacetylenes

No. (-CR=CR'-)n
          R         R'

Permeability coefficient, P×1010
[cm3(STP).cm/s.cm2. cmHg]
CO2                   CH4

Selectivity

CO2/CH4

With bulky substitute
1 Me      SiMe3
2 H         t-Bu
3 Me      SiMe2CH2SiMe3
4 Me      SiMe2CH2CH2SiMe3
5 H         o-C6H4SiMe3
6 H         o-C6H4CF3

With long n-alkyl group
7 Me       n-C7H15
8 Cl         n-C8H17
9 Cl         n-C6H13
10 Cl       n-C4H9
11 Me     SiMe2-n-C6H13
12 H        CH(n-C5H11)SiMe3
13 H        CH(n-C3H7)SiMe2-n-C6H13

With phenyl group
14 Ph        Me
15 Ph        Et
16 Ph        n-C6H13
17 Ph        Cl
18 H         o-C6H4Me
19 H         CH(n-C3H7)SiMe2Ph


19000 
560
310
150 
290 
130 


130 
170 
130 
180 
71 
120 
70

25
40
48 
23
15 
54


4300 
85
45 
28 
38 
6.6  


40 
46 
33 
30 
13 
21 
17

2.8
4.4
14 
1.3
3.0 
7.0


4.4 
6.6
6.9 
5.3 
7.6 
19.6 


3.3 
3.7 
3.9 
6.0 
5.5 
5.7 
4.1

8.9
9.1
3.4 
17.1
5 
7.7

Poly[1-(trimethylsilyl)-1-propane], [-CH3C=CSi(CH3)3-]n, designated hereafter PTMSP, has the highest intrinsic gas permeability of all synthetic polymers known at present.[46] Although it is a glassy polymer (Tg>200°C), was found to have an approximately ten-fold higher permeability for light gases than rubbery PDMS, but a somewhat lower selectivity.
    The exceptionally high permeability of PTMSP has been attributed to its large free volume, which was believe to arise from the PTMSP matrix consists of relatively stiff backbone chains separate by bulky trimethylsilyl side groups, Si(CH3)3. The stiffness of the backbone chains is due in part to the alternating double bonds of the polymer and in part to the fact that the bulky trimethylsilyl groups hinder intrasegmental rotation. These groups also prevent chain packing by serving as "intersegmental spacers".

    In order to study the structure/permeability relationships of these polymers and discover polymer structure that exhibit a substantially higher gas selectivity than PTMSP without a significant decline in the permeability, a large number of substituted polyacetylenes have been synthesized.[47, 48] Some of the results are presented in Table 3.[49]
     Because the rates of gas permeation through PTMSP decrease markedly with time or on thermal cycling, and their gas selectivities low, no practical uses of the polyacetylenes as membrane materials have been found so far.

3.1.3 Poly(vinyltrimethylsilane)
Poly(vinyltrimethylsilane), designated hereafter PVTMS, [-CH2CHSi(CH3)3-]n, is another glassy polymer which exhibits a high gas permeability (as shown in Table 1) due to its bulky trimethylsilyl pendant groups, as is the case also for PTMSP.[50] The main difference between these two polymers lies in that PVTMS has flexible vinyl backbone chains whereas PTMSP has stiff backbone chains with alternating doublebonds. The gas permeability of PVTMS is 1-3 orders of magnitude lower than that of PTMSP, but the gas selectivity of PVTMS is markedly higher than that of PTMSP. Now PVTMS membranes have been used in Russia to produce oxygen-enriched air for oxygen therapy. But this material for gas separation appears to have received relatively little attention in the west.
3.2 Improvement of the permeability of polymers with high selectivity
3.2.1 Polyimide
 
Polyimides have been identified as materials with high selectivities for CO2/CH4 (over 60-80). In addition to high selectivity, polyimides have high glass transition temperature(Tg>300°C) and can, therefore, be used at high service temperatures. A substantial insight has been gained in the last ten years into the effects of variations in the chemical structure of polyimides on gas solution and transport in these polymers. As a result, polyimides exhibiting increasingly higher gas selectivities and permeabilities are being developed in academic and industrial laboratories. In particular, those polyimides with a hexafluoro substitute carbon -C(CF3)2 in the backbone have been found to be considerably more CO2-selective than other glassy polymers with comparable permabilities.[51-60] A family of polyimides and related materials are used by DuPont and Ube for commercial gas separation membranes.Table 4 lists the permeability coefficients and selectivities of CO2/CH4 in the several excellent polyimides known as far.

Table 4. Permeabilities and selectivities of CO2/CH4 in polyimides* [51-60]

 

Permeability coefficient, P×1010
[cm3(STP). cm/s. cm2. cmHg]
CO2                       CH4

Overall selectivity

CO2/CH4

6FDA-ODA
6FDA-MDA
6FDA-IPDA
6FDA-6FpDA
6FDA-m-DDS       dense
                      asymmetric
6FDA-DMMDA
HQDPA-DMMDA

23.0
24.2
30.0
63.9
3.2
69.0
17.0
1.7

0.38
0.43
0.70
1.60
0.028
0.51
0.29
0.021

60.5
56.3
42.9
39.9
116
136
59.4
79.4

*6FDA:2,2-bis(3,4-dicarboxyphenyl) hexafluoro isopropane dianhydride
ODA:oxydianiline
MDA:methylenedianiline
IPDA:isopropylidenedianiline
PDA:phenylenediamine
DDS:diaminodiphenylsulfone
DMMDA:3,3'-dimethyl-4,4'-methylene dianiline
HQDPA:1,4-bis(3,4-dicarboxyphenyl) benzene dianhydride

    To reduce undesirable effects caused by CO2 induced swelling in CO2/CH4 separation processes, the polyimide structure was stabilized with crosslinks. One method for crosslinking membrane materials is the UV-irradiation of benzophenone-containing polyimides or bisphenol A based polyacrylates.[61, 62] This leads, especially in the separation of gas pairs with a large difference in molecular size, to a significant improvement of permselectivity. But simultaneously the permeability is decreased by crosslinking due to the strongly reduced chain mobility is decreased packing density of the polymer chains. A disadvantage of photochemical crosslinked membrane materials is that the degree of crosslinking depends strongly on experimental conditions, i.e. irradiation time or the type of mercury lamp. Moreover, under the experimental conditions of UV-irradiation crosslinking reaction as well as photo-fries rearrangements are possible.[62]
    Claudia et al. synthesized polyimides with strong polar associating functional groups as well as chemical crosslinked polyimides and copolyimides with simultaneous inhibition of chain mobility and intrasegmental packing by changes in backbone structure.[63] In the polymerization reaction 6FDA was used as dianhydride monomer and mPD(m-phenylene diamine) and DABA (diamine benzoic acid) were used as diamine monomers. With copolyimides containing strong polar carboxylic acid groups (i.e. 6FDA-mPD/DABA 9:1) reduced plasticization was seen up to a pure CO2 can be reduced at least up to a pure CO2 feed pressure of 35atm. With increasing degree of crosslinking, increasing CO2/CH4 selectivity was found because of reduced swelling and polymer chain mobility. By using ethylene glycol as a crosslingking agent, CO2 permeability was not significantly lower because the reduced chain mobility was compensated by the additional free volume caused by the crosslinks.
3.2.2 Polyaniline   
Polyaniline, (C6H4NH)n, has attracted considerable attention because of the surprisingly high gas selectivity.[64-66] Anderson et.al reported that a remarkable selectivity of CO2/CH4 in polyaniline, 336, was achieved by a complex process involving doping, undoping, and redoping membranes made from this polymer with select counterions, like acid solution of HF, HCl, HBr and HI. This report shows that conjugated polymer is one of the most exciting materials for gas separation membranes.[64]
    Hachisuka et al. used an oxidized polyaniline as undoped film, use various protonic acids, e.g., methane sulphonic acid, polyvinyl sulphonic acid, polyisoprene sulphonic acid, and polyvinyl sulfate as dopants. It has shown that CO2 permeability was increased by the formation of a quinonediimine unit in polyaniline with the oxidation. The selectivity of polyaniline membranes was found to be remarkably improved by doping. In particular, the selectivity of the polyaniline membrane for CO2/CH4 doped with polyvinyl sulphonic acid as a polymer dopant went up to over 2000. Table 5 show the selectivity of CO2/CH4 in polyaniline membranes doped with various dopants.[65]

Table 5. Permeabilities and selectivities of CO2/CH4 in polyaniline

Dopant

Permeability coefficient, P×1010
[cm3(STP). cm/s. cm2. cmHg]
CO2

Overall selectivity

CO2/CH4

Polyvinyl sulphonic acid
Polyvinyl sulfate
Polyisoprene sulphonic acid
HCl
Methane sulphonic acid
no

0.029
0.012
0.023
0.0027
0.015
0.079-0.14

2200
250
400
430
200
59-37

3.2.3 Polytriazoles and polyoxadiazoles    
These polymers have been stated to be chemically resistant as well as thermally stable with glass transition temperature between 260 and 360°C. Gebben et al. have measured the permeability of poly[1,3-phenyl-1,4-phenyl]-4-phenyl-1,3,4-triazole] (TIPT) to CO2 and CH4 at the temperatures up to 200°C.[35] Its selectivity with a CO2/CH4 mixture containing 20mol% CO2 at 25°C is over 60, a CO2 permeability is 10-20Barrer. The Tg of 270°C of TIPT is lower than that of a number of polyimides with comparable selectivities to CO2/CH4 gas pair. This suggests that the intrasegmental mobility of TIPT is higher than that of the latter polymers, perhaps due to the mobility of the phenyl rings in the para positions. The high selectivity of TIPT to CO2/CH4 gas pairs is unexpected in view of its relatively large fractional free volume (FFV) and moderate Tg. A sharper free-volume distribution in TIPT any conceivably account for the gas selectivity of TIPT.
    An unusually high CO2/CH4 selectivity of 168, at an unspecified temperature, has been reported for a poly(1,3,4-oxadiazole)(POD) which incorporates a diphenyl ether (DPE) moiety in its backbone chains.[64] However, the permeability of this polymer to CO2 is very low, 0.3 Barrer. The high Tg of DPE-POD of 333°C indicates a low intrasegmental mobility, i.e., a high backbone-chain stiffness, in spite of the presence of potentially mobile -O- linkages in the DPE moiety. The very high CO2/CH4 selectivity and low permeability to CO2 suggest a high chain packing density and low FFV, and possible a narrow free-volume distribution. Rotation around -O- linkages is probably inhibited by the high chain packing.
    By contrast, substitution in POD of either 1,1,3-trimethyl-3-phenylindane (PIDA) or 4,4',2,2'-diphenyl hexafluoropropane (HF) yielded polymers with much higher permeabilities to CO2, 78-93 Barrer, but relatively low CO2/CH4 selectivities of 26-28.
3.3 Polymer blending

02803607.gif (15226 bytes)
Fig. 7
Permeability of CO2 for various PSF/PI blends at 40°C

Although polyimide show excellent mechanical properties, high temperature and chemical resistance, and separation performances for CO2, the polyimides are seriously affected by highly soluble penetrants CO2, which above a given partial pressure can plasticize the polymer matrix.[68] The critical partial pressure of plasticization for polyimide is relatively low, varying between 10-20 bar.[69] For this reason polyimide membranes have found little application in the separation of CO2 from light hydrocarbons in enhanced oil recovery or in landfill gas upgrading programs.
    Polymer blending has been taken as optimization of membrane properties. If one could achieve equivalent segmental composition by either blending or copolymerization, it is currently thought that equivalent solution/diffusion separation properties would result. Although earlier extensive work on polymer blend membranes was attempted to do, it is very difficult to find miscible polymer blends with satisfied separation performances.[33, 34]
    However, Sakellaropoulos et al. prepared the blend membrane from the mixtures of the polysulfone and the aromatic polyimide.[69, 70] Because the critical pressure of plasticization for polysulfone exceed 55Bar, blend membranes were considerably more resistant to the plasticization phenomenon compared with those of pure polyimide, shown as in Fig. 7.[71] In addition, the membranes showed significant permeability improvements, compared to pure polyimide, with minor change in their selectivity. Polysulfone and polyimide proved to be completely miscible polymers as confirmed from optical microscopy, glass transition temperatures and spectroscopy analyses of the prepared mixtures. The complete miscibility permits the preparation of symmetric and asymmetric blend membranes in any proportion (1-99%) of polysulfone and polyimide.
    Blending of suitably selected polymers is generally considered as a viable method to modify the properties of polymers used for preparation of gas separation membranes. But the candidate polymer pairs for preparation of miscible blends are few.

3.4 Organic-inorganic hybrid membranes

Table 6. Permeabilities and selectivities of CO2/CH4 in the 6FPAI and 6FPAI/TiO2

 

Permeability coefficient, P×1010
[cm3(STP). cm/s. cm2 . cmHg]
CO2                        CH4

Overall selectivity

CO2/CH4

6FPAI
6FPAI/TiO2

52.7
44.7

1.80
1.3

29.3
34.4

Because of the limitations of purely organic polymers, especially to plasticization, organic-inorganic hybrid materials may offer new possibilities to gas separation application.[35, 72] Hybrid organic-inorganic materials can be formed utilizing a sol-gel process via a number of approaches. One method involves treating an orthosilicate directly with the organic polymer or an oligomer, which contains functional groups capable of a cross-reaction with the inorganic oxide, thus provide connectivity between the organic phase and inorganic network.[73] Alternatively, alkoxysilane monomers with a polymerizable organic moiety covalently attached to the silicon can be precipitated, followed by polymerization of the organic moiety.[72] Finally, in situ polymerization of the alkoxide within a swollen polymer network can be used to form nano- or micro composite materials without covalent crosslinks.[72]
    Hu et al. have successfully fabricated a fluorinated poly(amide-imide)/TiO2 nano-composite membranes by a sol-gel method. The nano-composite membrane has a denser and a more rigid structure compared to the corresponding pure poly(amide-imide) membrane. Furthermore, there appear to be specific intermolecular interactions between CO2 and the TiO2 domains. Finally, the composite membrane has shown higher permselectivities for CO2/CH4, as shown in Table 6, even at very low volume concentration of the TiO2 component. Such results are encouraging, because they suggest that possibly higher selectivities could be achieved at increasingly higher concentrations of the ceramic component.

4 CURRENT PROBLEMS AND RESEARCH TOPICS IN FUTURE  
These "tailor-made" polymers allow both higher selectivities of CO2/CH4 and higher permeabilities of CO2 compared to the traditional polymers. Although a very substantial amount of data on the solution, diffusion, and permeation of CO2/CH4 in a variety of rubbery and glassy polymers was already available, the relationships between the chemical structure of polymers and their gas permeability and selectivity were only poorly understood. The new polymers with good gas selectivity or permeability, like 6FDA-ODA, polyaniline, PTMSP and PDMS etc., was occasionally synthesized or found. But it is largely based on prior experience and considerable trial and error. And these "tailor-made" polymers themselves exist some problems for CO2-selective separation processes. Polyimide is sensitive to plasticize by CO2. Polyaniline with highest selectivity for CO2/CH4 known so far is not sufficiently stable in the presence of humidity, prolonged oxygen exposure and physical aging.
    Systematic primary chemical structure modificatioin of the polymer backbone which simultaneously increase chain rigidity and decrease chain packing density have been shown to lead to membranes with improved permeability without significant losses in gas permselectivity. But as chain rigidity increases, it becomes more difficult to find acceptable solvents in which to dissolve the material. Some of the lower free volume imides must also be cast in amide acid form followed by thermal ring closure. Some such materials, in fact are used in commercial membranes; however, they are clearly more complex and expensive.
    A large increase in solubility might induce plasticization, which would decrease membrane permselectivity. Another deleterious effect of strong polymer-penetrant interactions is the loss of mobility selectivity due to excessive binding between the gas and the polymer. Clearly, the strategy of augmenting permselectivity by increasing solubility selectivity is a delicate balance. An increase in polymer polarity with a simultaneous increase polymer free volume might be expected to increase CO2/CH4 solubility selectivity, permeability and permselectivity. However, studies involving the introduction of substitutents, which simultaneously interact with one component in a gas mixture and frustrate chain packing have not been reported.
    From a practical standpoint of processing advance materials, most current generation commercial membranes have asymmetric forms with an open porous support on top of which is an integral skin formed during the cast of a solution of the polymer into a nonsolvent bath. Besides high permeability and selectivity values, processability of the membrane material into desired membrane morphology is of utmost importance.
    In extreme cases, the gaseous feed stream can be composed of over 50% CO2 at an elevated temperature and up to a feed pressure of 60atm. These extreme operation conditions result in the swelling and plasticization of most membrane materials by the CO2 present in the feed stream. For the application of membrane based gas separation in EOR processes, it is important to develop new membrane materials with reduced plasticization effects due to higher CO2 partial pressure.
    Materials research on membranes for CO2/CH4 will fouse on these two fields: (a) development of new polymer membrane materials with improved selectivities and permeabilities, simultaneously with unplasticization to higher pressure of CO2. (b) fabrication of new membrane morphologies allowing higher fluxes(e.g., asymmetric membranes, hollow fiber membranes, and composite membranes) for a given polymer with high permeability of CO2 and high selectivity of CO2/CH4.

REFERENCES   
[1] Hao J H. The preparation of Integrally-skinned asymmetric membrane for separation of CO2/CH4 and studies on its formation mechanism, PhD Dissertation, Tianjin University, 1995.
[2] Schell W J. CEP, 1982, 78: 45.
[3] David P. Oil & Gas Journal, Technology, 1984: 85.
[4] Goddin C S. Hydrocarbon Processing, 1982, 61 (5): 125.
[5] Bhide B D, Stern S A. J. Membrane Sci., 1993, 81: 209.
[6] Cooley T E, Dethlff W L. Chem. Eng. Progress, 1985, 81: 45.
[7] Fournie F J C, Agostini T P. J. Petroleum Tech., 1987, 39: 707.
[8] Mc Kee R L, Changela M K, Reading G J. Hydrocarbon Processing, 1991, 70 (4): 63.
[9] Maddox R N, Morgan D J. Hydrocarbon Processing, 1986, 64 (8): 59.
[10] Babcoek R E, Spillman R W, Goddin G S et al. Energy Progress, 1988, 8 (3): 135.
[11] Schell W J. J. Membrane Sci. 1985, 22: 217.
[12] Vansant E F, Dewolfs R. Gas Separation Technology. Amsterdam: Elsevier Science Publishers B. V., 1990: 457.
[13] Rautenbach R, Welsch K J. membrane Sci., 1994, 87 (1): 107.
[14] Cecille L, Toussaint J C. Future Industrial Prospects of Membrane Processes, New York: Elsevier, 1989.
[15] Koros W J, Fleming G K. J. Membrane Sci., 1993, 83 (1): 1-80.
[16] Bhide B D, Stern S A. J. Membrane Sci., 1993, 81 (3): 238.
[17] Teramoto M, Nakai K, Ohnishi N et al. Ind. Eng. Chem. Res. 1996, 35 (2): 538-45.
[18] Koros W J, Fleming G K, Jordan S M et al. Prog. Polym. Sci., 1988, 13: 339-401.
[19] LeBlanc O H, Ward W J, Matson S L et al. J. Membrane Sci., 1980, 6: 339-343.
[20] Weinkauf D H, Paul D R. in Barrier Polymers and Barrier Structures, W. J. Koros, ed., American chemical Society, 1990.
[21] Koros W J, Hellums M W. Encyclopedia of polymer Science, 2nd edn., New York:Wiley-Interscience Publishers, Supplement Volume, 1989: 724.
[22] Ghosal K, Freeman B D. Polymers for Advanced Technologies, 1994, 5: 673-697.
[23] Chern R T, Koros W J, Hopfenberg H B. Material Science of Synthetic membranes, D. R. Lloyd, ed., Washington, DC: American Chemical Society, 1985: Chapter 2.
[24] Chern R T, Brown N F. Macromolecules, 1990, 23: 2370.
[25] van Amerongen G J. J. Polym. Sci., 1950, 5: 307.
[26] Pilato L, Litz L, Hargitay B et al. Polym. Preprint, Am. Chem. Soc. Div. Polym. Sci., 1950, 5: 307.
[27] Koros W J. J. Polym. Sci.:Polym. Phys. Ed., 1985, 23: 1611.
[28] Ghosal K, Chern R T, Freeman B D et al. J. Polym. Sci., Part B: Polym. Phys. 1995, 33 (4): 657-66.
[29] Chern R T, Provan C N. Macromolecules, 1991, 24: 2203.
[30] Stam H. Future Industrial Prospects of Membrane Processes, Ed. By Lecille L and Toussaint J C, NY: Elsevier, 1989.
[31] Kesting R E, Fritzsche A K, Murphy M K et al. J. Applied Polym. Sci., 1990, 40: 1557.
[32] Barbari T A, Datwani S S. J. Membrane Sci., 1995, 107: 263-266.
[33] Paul D R. Polymer Compatibility and Incompatibility: Principle and Practice, K. Solc Ed., MMI Press Symp. Ser., 2, New York: Harwood Academic Publishers, 1982.
[34] Chiou J S, Paul D R. J. Appl. Polym. Sci., 1987, 33: 2935.
[35] Gebben B, Mulder N H V, Smolders C A. J. Membrane Sci., 1989, 46: 29-41.
[36] Li Y, Ding M, Xu J. Polymer International, 1997, 42 (1): 121-126.
[37] Hu Q, Marand E, Dhingra S et al. J. Membrane Sci., 1997, 135: 5-79.
[38] Hensema E R. Adv. Mater., 1994, 6 (4): 269-279.
[39] Pye D G, Hoehn M H, Panar M. J. Applied Polym. Sci., 1976, 20: 287-301.
[40] Hoehn H H. Material Science of Synthetic membranes, D. R. Lloyd, ed., Washington, DC: American Chemical Society, 1985: 81-98.
[41] Stern S A, Shah V M, Hardy B J. J. Polym. Sci., Part B: Polym. Phys., 1987, 25: 1263-1298.
[42] Shah V M, Hardy B J, Stern S A. J. Polym. Sci., Part B: Polym. Phys., 1986, 24: 2033-2047.
[43] Ashworth A J, Brisdon B J, England R et al. Br. Polym. J., 1989, 21: 491.
[44] Ashworth A J, Brisdon B J, England R et al. J. Membrane Sci., 1991, 56: 217.
[45] Matsuyama H, Teramoto M, Hirai K. J. Membrane Sci., 1995, 99: 139-147.
[46] Masuda T, Isobe E, Higashimura T. J. Am. Chem. Soc., 1983, 105: 7473-7474.
[47] Takada K, Matsuya H, Masuda T et al. J. Applied Polym. Sci., 1985, 30: 1605-1616.
[48] Masuda T, Iguchi T, Tang B Z. Polymer, 1988, 29: 2141-2149.
[49] Odani H, Masuda T. Design of polymer membranes for gas separation, in N. Toshima(Ed.), New York, NY, 1992, 107-144.
[50] Davydova M B, Yampol'skii Yu P, Durgar'jan S G. Poly. Sci. USSR, 1988, 30: 1501-1507.
[51] O'Brien K C, Koros W J, Husk G R. J. Membrane Sci., 1988, 35: 217-230.
[52] Kim T H, Koros W J, Husk G R et al. J. Membrane Sci., 1988, 37: 45-62.
[53] Stern S A, Mi Y, Yamamoto H et al. J. Polym. Sci., Polym. Phys. Ed., 1989, 27: 1887-1909.
[54] Tanaka K, Kita H, Okano M et al. Polymer, 1992, 33: 585-592.
[55] Matsumoto K, Xu P. J. Membrane Sci., 1993, 81: 23-30.
[56] Tanaka K, Kita H, Okamoto K. J. Polym. Sci., Part B: Polym. Phys., 1993, 31: 1127-1133.
[57] Kawakami H. J. Applied Polym. Sci., 1996, 62: 965-971.
[58] Rene A, Canada, CA2133046, 1995.
[59] Hiroyoshi K, Masato M, Shoji N. J. Applied Polm. Sci., 1996, 62 (7): 965-970.
[60] Hiroyoshi K, Masato M, Shoji N. J. Membrane Sci., 1996, 118 (2): 223-230.
[61] Kita H, Inada T, Tanaka K et al. J. Membrane Sci., 1994, 87: 139.
[62] Wright C T, Paul D R. J. Membrane Sci., 1997, 124: 161 .
[63] Staudt-Bickel C, Koros W J. J. Membrane Sci., 1999, 155 (1): 145-154.
[64] Anderson M R, Mattes B R, Reiss H et al. Science, 1991, 252: 1412-1415.
[65] Hachisuka H, Ohara T, Ikeda K I et al. J. Applied Polym. Sci., 1995, 56: 1479-1485.
[66] Chang M J, Liao Y H, Myerson A S et al. J. Applied Polym. Sci., 1996, 62 (9): 1427-1436.
[67] Hensema E R, Ph.D. Dissertation, University of Twenty, Enschede, 1991.
[68] Koros W J, Coleman M R, Walker D R B, Ann. Rev. Sci., 1992, 22: 47.
[69] Sakellaropoulos G P, Kaldis S P, Kapantaidakis G C et al. European patent, EP 778077, 1997.
[70] Kapantaidakis G C, Kaaldis S P, Dabou X S et al. J. Membrane Sci., 1996, 110: 239-247.
[71] Sander E S, Jordan S M, Sabrmanian R. J. Membrane Sci., 1992, 74: 29.
[72] Okazaki I, Ohya H, Semenova S I et al. J. Membrane Sci., 1998, 141 (1): 65-78.
[73] Smaihi M, Jermoumi T, Marigan J et al. J. Membrane Sci., 1996, 116: 211-220.

 


[ Back ] [ Home ] [ Up ] [ Next ] Mirror Site in USA  Europe  China  CSTNet ChinaNet