Sunteți pe pagina 1din 13

Journal of Membrane Science 239 (2004) 105–117

Gas solubility, diffusivity and permeability in poly(ethylene oxide)


H. Lin, B.D. Freeman∗
Department of Chemical Engineering, Center for Energy and Environmental Resources, The University of Texas at Austin,
10100 Burnet Road, Building 133, Austin, TX 78758, USA

Received 14 March 2003; received in revised form 7 August 2003; accepted 25 August 2003

Available online 6 May 2004

Abstract

The effect of pressure on the solubility, diffusivity, and permeability of He, H2 , O2 , N2 , CO2 , CH4 , C2 H4 , C2 H6 , C3 H6 and C3 H8 in
poly(ethylene oxide) (PEO) is reported at 35 ◦ C. Additionally, the temperature dependence of permeability is reported. The effect of polar
ether linkages in PEO on gas transport is illustrated by comparing transport properties in PEO with those in polyethylene (PE). For example,
at 35 ◦ C and infinite dilution, semi-crystalline PEO exhibits CO2 permeability coefficient of 12 Barrers, and CO2 /H2 and CO2 /N2 pure gas
selectivities of 6.7 and 48, respectively. In contrast, at similar conditions, the permeability of PE to CO2 is 13 Barrers, but the CO2 /N2 selectivity
is only 13. In addition to good separation properties for quadrupolar–nonpolar gas pairs, PEO also shows interestingly high selectivity for
olefins over paraffins, which is ascribed to favorable interaction between the polar ether groups in PEO and olefins. For example, the infinite
dilution permeability of PEO to propylene is 3.8 Barrers and pure gas propylene–propane selectivity is 2.7 at 35 ◦ C. At similar conditions, PE
exhibits propylene–propane selectivity of only 1.4.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Poly(ethylene oxide); Carbon dioxide; Separation; Solubility; Permeability

1. Introduction tions [5]. This affinity may be harnessed to prepare materials


that are more permeable to CO2 than to H2 . Such membranes
The use of membranes to selectively remove CO2 from could be used to selectively remove CO2 and other acid gases
mixtures with H2 , CO, N2 and CH4 is of interest for a wide from mixtures with H2 . Using such membranes, H2 would
variety of applications such as syngas processing, flue gas be produced in the high-pressure residue stream, possibly
rectification, and natural gas separations [1,2]. One particu- reducing or eliminating expensive recompression steps.
lar example is hydrocarbon reforming, which is the dominant Because of the interest in improving CO2 separation
technology for H2 production [3]. This process produces performance, there have been many recent studies of poly-
mixtures of H2 and CO2 that must be separated. Unfortu- mer membranes containing poly(ethylene oxide) (PEO),
nately, conventional membranes are usually hydrogen selec- poly(ethylene glycol) (PEG) or other groups bearing polar
tive [4]. In such membranes, the desired hydrogen product ether segments that can interact favorably with acid gases
is produced at low pressure in the permeate stream of the such as CO2 [6–21]. For example, blending PEG with cel-
membrane, but it is typically utilized at pressures equal to lulose nitrate or cellulose acetate increased both CO2 per-
or higher than feed pressure. The cost of repressurizing the meability and CO2 /N2 selectivity [6,8]. Block copolymers
hydrogen to feed pressure can be prohibitive and contributes containing PEO segments exhibit high CO2 permeability
to the use of alternative processes, such as liquid absorption, and high CO2 /N2 and CO2 /H2 selectivity [9–13,15–18,20].
to remove acid gases such as CO2 from such streams. For example, Bondar et al. studied poly(ether-b-amide)
Polymers containing polar moieties, such as ether groups, copolymers; they observed a CO2 permeability of 120 Bar-
have an affinity for CO2 due to dipole–quadrupole interac- rers and a CO2 /H2 pure gas selectivity of 9.8 at 35 ◦ C and
a feed pressure of 1.0 MPa (10 atm) [17]. Table 1 presents
∗ Corresponding author. Tel.: +1-512-232-2803; CO2 permeability and CO2 /N2 selectivity in some block
fax: +1-512-232-2807. copolymers containing roughly the same weight percent
E-mail address: freeman@che.utexas.edu (B.D. Freeman). of PEO segments and different hard segments. While the

0376-7388/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2003.08.031
106 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

Table 1 2. Background
Gas transport properties in block copolymers containing similar amounts
of poly(ethylene oxide) segments and a variety of hard segments
The permeability of a polymer to a gas A, PA , is [4]:
Hard segmenta PEO (wt.%) PCO2 [Barrer] PCO2 /PN2
NA l
PA ≡ (1)
Polyimide p2 − p 1
PMDA-pDDS [10] 68.6 238 49
BPDA-ODA [10] 62.3 117 51 where NA is the steady-state flux of gas through the film, l
Polyamide is the film thickness, and p2 and p1 are the upstream (i.e.,
IPA-ODA [10] 68.3 58 53 high) and downstream (i.e., low) partial pressures of gas
PA12 [17] 55 120 51 A, respectively. If the diffusion process obeys Fick’s law,
PA6 [17] 57 66 56 and the downstream pressure is much less than upstream
Polyurethane pressure, the permeability is given by [4]:
MDI-BPA [10] 60.5 48 47
a
PA = DA SA (2)
The detailed chemical structures are available in the corresponding
references. The gas permeability was measured at 35 ◦ C and 200 kPa where DA is the average effective diffusivity through the
(2 atm) in [10], and at 35 ◦ C and 1.0 MPa (10 atm) in [17].
film, and SA is the apparent sorption coefficient [4]:
C2
SA = (3)
gas selectivity is similar among all these materials, gas p2
permeability varies significantly. These results suggest that where C2 is the concentration of gas dissolved in the polymer
the PEO phase is the continuous path for gas permeation when the gas pressure in contact with the polymer is p2 .
[10,17], and gas permeability depends strongly on the de- The ideal selectivity of a membrane for gas A over gas B
tailed morphology, such as domain shape and interspatial is the ratio of their pure gas permeabilities [4]:
arrangement, which could be influenced by the hard seg-   
PA DA SA
ment composition and the length of PEO and hard segment αA/B = = (4)
blocks. PB DB SB
Despite this interest in PEO-based materials for acid gas where DA /DB is the diffusivity selectivity, the ratio of the
selective separations, a systematic study of the solubility, dif- diffusion coefficients of gases A and B. The ratio of the solu-
fusivity, and permeability properties of pure poly(ethylene bilities of gases A and B, SA /SB , is the solubility selectivity.
oxide) has not been reported, due in large measure to its high Diffusivity selectivity is strongly influenced by the size dif-
crystallinity and, consequently, low transport property val- ference between the penetrant molecules and the size-sieving
ues. In this study we present physical characterization and ability of the polymer matrix, whereas solubility selectivity
gas transport properties of solution cast PEO films. The ef- is controlled by the relative condensability of the penetrants
fect of pressure on gas permeability, diffusivity and solubil- and the relative affinity between the penetrants and the poly-
ity is reported for a variety of nonpolar gases (He, H2 , O2 , mer matrix [4].
N2 ), a series of olefinic and aliphatic hydrocarbons (CH4 , Gas transport properties in a semi-crystalline polymer
C2 H6 , C2 H4 , C3 H8 , C3 H6 ), and CO2 . The effect of tem- such as PEO are usually modeled by assuming that the crys-
perature on permeability is also reported. Relevant physical tals act as an impermeable dispersed phase imbedded in an
properties of these penetrants, which will be used later in amorphous phase and by then developing models for the in-
the discussion, are recorded in Table 2. fluence of crystallinity on solubility and diffusivity. In a rub-
bery polymer, the effect of crystallinity on penetrant sorp-
tion is typically represented as follows [22]:
Table 2
SA = SA,a φa (5)
Penetrant parameters characterizing size (critical volume) and condens-
ability (critical temperature) where SA is the observed solubility coefficient, SA,a is the
Size (critical volume Condensability (critical solubility coefficient in the amorphous polymer, and φa is
[39] (cm3 /mole)) temperature [39] (K)) the amorphous phase volume fraction.
He 57.4 5.19 The influence of crystallinity on diffusivity is traditionally
H2 65.1 33.24 described as follows [23]:
O2 73.4 154.6
N2 89.8 126.2 DA,a
DA = (6)
CO2 93.9 304.21 τβ
CH4 99.2 191.05
C 2 H4 130.4 282.40 where DA,a is the diffusion coefficient in the amorphous
C 2 H6 148.3 305.35 polymer, τ is a tortuosity factor, and β is a chain im-
C 3 H6 181.0 364.9
mobilization factor. τ characterizes the tortuosity of the
C 3 H8 203.0 369.8
amorphous phase caused by the presence of impermeable
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 107

crystallites dispersed in it. Simple models from compos- PEO is susceptible to oxidative degradation, due to the pres-
ites theory, such as the one given below, are often used to ence of weak C–O bonds in the backbone [24], the anneal-
describe the influence of crystallinity on tortuosity [23]: ing time needed to be as short as possible and the annealing
1 temperature needed to be as low as possible. Therefore, to
τ= (7) minimize the degradation and prepare a nonporous film, we
φa
decided to anneal the sample under vacuum just above its
The chain immobilization factor, β, accounts for the re- melting temperature for 2 h.
stricted segmental mobility in the amorphous phase by
crystallites. In the simplest case, when β = 1 (i.e., no chain 3.2. Density
immobilization), gas permeability is given by [23]:
PA = SA DA = PA,a φa2 (8) Polymer film density was determined by hydrostatic
weighing using a Mettler Toledo balance (Model AG204,
where PA,a is the estimated permeability of penetrant A in Switzerland) and a density determination kit [25]. In this
amorphous phase polymer. method, a liquid with a known density (ρo ) (the so-called
Generally, crystallites decrease both gas diffusivity and auxiliary liquid) is needed, and the film density (ρ) is
solubility. By reducing amorphous phase polymer chain mo- calculated as follows:
bility, crystallites can also increase activation energy of dif- MA
fusion, especially for larger penetrants [23]. These effects ρ= ρo (9)
MA − M L
become more pronounced as the crystalline volume fraction
increases or as crystallite size decreases [23]. where MA is the film weight in air, and ML is the film
weight in the auxiliary liquid. Iso-octane was used as the
auxiliary liquid because PEO is reported to be insoluble in
3. Experimental iso-octane [26]. The film weight determination in iso-octane
was performed as quickly as possible to reduce any swelling
3.1. Materials and sample preparation of PEO due to iso-octane sorption.

Poly(ethylene oxide) with a weight-average molecular 3.3. Thermal characterization


weight of approximately 1,000,000 was purchased from
Aldrich Chemical Company, Milwaukee, WI and used as- Thermal transitions were determined using a Perkin
received. The gases and vapors used in pure gas permeation Elmer (Shelton, CT) DSC 7 differential scanning calorime-
and sorption measurement had a purity of at least 99% and ter (DSC). Experiments were performed at a heating rate of
were obtained from National Specialty Gases (Raleigh, NC). 20 ◦ C/min over temperatures ranging from 0 to 120 ◦ C. The
Gas permeation and sorption experiments were conducted degree of crystallinity was evaluated from the area under
using dense PEO films of uniform thickness, prepared by the melting endotherm in the first scan [27].
dissolving solid PEO in ultrapure water, which is produced
by a Milli-Q water purification system (Millipore Corpora- 3.4. Permeation measurements
tion, Bedford, MA). Films were cast from 3 wt.% water so-
lutions into plastic or Teflon flat-bottomed Petri dishes. Af- The pure gas permeation properties in PEO were deter-
ter air-drying at ambient conditions for 3–7 days, films were mined using a constant volume/variable pressure apparatus
removed from the dish and placed on a Teflon plate. PEO [17]. The permeation cell is a stainless steel filter holder from
films were stored under vacuum in an oven at room temper- Millipore Corporation (47 mm disc filter, Bedford, MA) with
ature for 24 h to assist in removal of residual water and then an area of 13.8 cm2 . PEO samples were partially masked us-
annealed at 70–80 ◦ C for 2 h. After annealing, samples were ing impermeable aluminum tape on the upstream face, or on
removed from the oven and cooled at ambient conditions the upstream and downstream faces as described by Mogri
before being placed in a desiccator, where they were stored and Paul [28]. Both masking methods gave very similar per-
until use. Film thicknesses were in the range of 100–300 ␮m. meability values, consistent with the observations of Mogri
Initially, it was difficult to prepare defect free PEO films. and Paul [28]. The o-ring in the permeation cell was in direct
We arrived at the film preparation protocol described in the contact with the aluminum tape so that the soft rubbery PEO
previous paragraph after trying several solvents (acetoni- film would not be damaged by the o-ring. After aluminum
trile, chloroform, and water), concentrations and evaporation tape masking, the surface area of the sample available for
rates. However, these variables did not have as much impact gas transport was 5.1 cm2 .
on the ability to prepare defect free samples as annealing. After a film was mounted in the system, both upstream
Generally, samples prepared without annealing were defec- and downstream volumes were exposed to vacuum overnight
tive (i.e., they showed similar permeabilities to all gases). to degas the film. The leak rate in the system was always
In contrast, annealing solvent cast films above the polymer measured before starting the permeation experiments, and
melting temperature resulted in defect free films. Because then the pressure increase in the downstream volume was
108 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

recorded. Gas permeability (cm3 (STP) cm/(cm2 s cm Hg))


was calculated from the steady-state rate of pressure in-
crease in a fixed downstream volume:
    
Vd l dp1 dp1
PA = − (10)
p2 ART dt ss dt leak

where Vd is the downstream volume (cm3 ), l is the film thick-


ness (cm), p2 is the upstream absolute pressure (cm Hg), A
is the film area available for gas transport (cm2 ), the gas
constant, R the 0.278 cm Hg cm3 /(cm3 (STP) K), T is the ab-
solute temperature (K) and (dp1 /dt)ss and (dp1 /dt)leak are the
steady-state rates of pressure rise (cm Hg/s) in the down-
stream volume at a fixed upstream pressure and under vac-
uum, respectively. (dp1 /dt)leak was usually less than 10% of
(dp1 /dt)ss . The downstream pressure was always less than
2 cm Hg, which was very low compared with lowest up-
Fig. 1. First scan differential scanning calorimetry thermogram of poly
stream pressure considered (400 kPa (4 atm)). (ethylene oxide).

3.5. Sorption measurements where H is the melting enthalpy, obtained from the area
under the melting endotherm, and Hc is the melting en-
Gas solubility was determined using a dual-volume, dual- thalpy of 100% crystalline PEO. From the DSC experiment,
transducer apparatus based on the barometric, pressure- H was 121 J/g, which is close to the literature value of
decay method [16]. Polymer samples were degassed by 125 J/g for PEO with a molecular weight of 4,000,000 [27].
exposure to vacuum in the sorption cell overnight before A variety of values of Hc exist in the literature ranging
beginning sorption measurements and between each gas. from 166.4 to 265 J/g [26]. In this study, we used the value
Because PEO is highly crystalline and, therefore, exhibits of 166.4 J/g suggested by Simon and Rutherford [30]. Us-
low solubility, the uncertainty in the measured solubility co- ing this value in Eq. (11) yields a crystallinity value of
efficients of the low sorbing permanent gases was high. For 72.7 wt.%.
example, the uncertainty in solubility was estimated to be Volume fraction crystallinity, φc , was calculated as fol-
±50% for N2 , O2 and CH4 , and less than 10% for the other lows:
penetrants. Solubilities of H2 and He at 35 ◦ C were too low  
ρ
to measure. Uncertainty was estimated by a standard propa- φc = χc (12)
ρc
gation of errors analysis [29], where uncertainties of all rel-
evant measured parameters propagate and contribute to the where ρ is the measured PEO density and ρc is the crystal
uncertainty of the solubility. The dominant contribution to density calculated from the unit cell dimensions of a PEO
the uncertainty comes from the uncertainty in measuring the crystallite, which is 1.234 g/cm3 at room temperature [30].
pressure in the sorption cell. For this measurement, we use PEO film density, ρ, was 1.209 g/cm3 , which is consistent
the two Sensotec pressure transducers (Model Super THE), with values (1.20–1.22 g/cm3 ) reported in the literature for
and the uncertainty in the measured pressure is ±1.7 kPa PEO [25]. Using Eq. (12), the crystallinity was 71 vol.%.
(±0.017 atm). The full pressure range of the transducer is The volume percent crystallinity can also be estimated
0–3.4 MPa (0–34 atm). from the density using the following relation:
ρ − ρa
φc = (13)
ρc − ρ a
4. Results and discussion
where ρa is the density of amorphous PEO (1.124 g/cm3 )
4.1. Thermal and physical characterization [30]. The φc value calculated from Eq. (13) was 0.77, which
is consistent with that obtained from Eq. (12). In the remain-
Fig. 1 presents a first scan DSC thermogram of PEO. The der of this report, the crystallinity calculated from the DSC
peak of the melting endotherm is 68 ◦ C, which is near the result will be used to estimate amorphous phase permeabil-
value of 66 ◦ C reported for PEO with a molecular weight of ity and solubility values.
1,000,000 [30]. The weight fraction crystallinity in the film, Transport properties in polymers are sensitive to the
χc , was estimated as follows [27]: amount of free volume in the polymer matrix [4]. In semi-
crystalline polymers such as PE and PEO, transport is
H presumed to occur only in the amorphous regions of the
χc = (11)
Hc polymers, so it is of interest to compute the fractional free
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 109

Table 3
Physical property comparison between poly(ethylene oxide) and
poly(ethylene)
Polymer Tg a (◦ C) FFVb Densityc Crystallinity Solubility
(g/cm3 ) (vol.%) parameter
(MPa1/2 )
[26]
PEO −50 0.139 1.124 71 19.8
PE −44 0.188 0.854 43d 16.6
a In these polymers, Tg depends strongly on the degree of crystallinity
[50]. The value reported for PEO corresponds to 74 vol.% crystallinity
[50]. The value reported for PE corresponds to a sample with 43 vol.%
crystallinity; in a PE sample having 71 vol.% crystallinity, the Tg would be
−10 ◦ C [50]. If one extrapolates to hypothetical wholly amorphous PEO
and PE, the estimated Tg values are −85 and −80 ◦ C, respectively [50].
b FFV values based on estimated density of amorphous PE and PEO.
c Estimated density of wholly amorphous polymer at room temperature.
d Alathon 14 polyethylene was chosen from the reference by Michaels

and Bixler [22]. Later in this report, gas transport properties of Alathon
14 are compared with those of PEO.

volume in the amorphous regions of such polymers. The


amorphous density data were used to estimate the frac-
tional free volume (FFV) of the amorphous phase using the
following group contribution method [31]:
V − Vo
FFV = (14)
V
where V is the specific volume of the amorphous polymer at
the temperature of interest and Vo is the specific occupied
volume at 0 K, which was estimated as 1.3 times the van
der Waals volume [32]. The densities of amorphous PEO
and PE were taken to be 1.124 g/cm3 at 20 ◦ C [30] and
0.854 g/cm3 at 25 ◦ C [22], respectively. Using these values,
the FFV of the amorphous phase of PEO is 0.139. This
Fig. 2. (a) Permeability coefficients of penetrants in semi-crystalline PEO
value is significantly lower than that of PE, which has an at 35 ◦ C as a function of upstream pressure. The scale for gas permeability
amorphous phase FFV of 0.188. These values are recorded values is shown on the left y-axis for all gases except CO2 , whose scale is
in Table 3. in the right y-axis. (b) Permeability coefficients of hydrogen and several
hydrocarbons in semi-crystalline PEO at 35 ◦ C as a function of upstream
pressure.
4.2. Permeability

Fig. 2a and b present permeability coefficients for gases simple empirical model [33]:
at 35 ◦ C as a function of upstream pressure. For the perma- PA = PA,o (1 + m p) = PA,o (1 + mp2 ) (15)
nent gases (He, H2 , O2 , and N2 ), permeability coefficients
are essentially independent of pressure, while the permeabil- where PA,o is the permeability coefficient at an upstream
ity coefficients of CO2 and hydrocarbons such as C3 H8 in- pressure, p2 , of 0 (i.e., infinite dilution permeability), m the
crease with increasing pressure. This behavior is consistent adjustable constant, and p is the difference between the
with gas permeation properties in rubbery polymers [33]. upstream and downstream pressure, p = p2 − p1 . Since
For strongly sorbing penetrants, such as CO2 and C3 H8 , high downstream pressure, p1 is much less than upstream pres-
penetrant concentrations can plasticize the polymer matrix sure p2 , p can be replaced by p2 . Infinite dilution perme-
by increasing polymer local segmental motion, thus enhanc- ability coefficients for different penetrants were calculated
ing penetrant diffusion coefficients and, in turn, permeabil- using Eq. (15). Estimated amorphous phase permeability
ity [34]. Additionally, the solubility of condensable compo- coefficients were estimated using Eq. (8). This calculation
nents typically increases with increasing pressure, and this presumes that tortuosity obeys Eq. (7) and that the chain
factor also acts to increase permeability coefficients [35]. immobilization factor, β, is unity. Because we did not have
For the data in Fig. 2a,b, the pressure dependence of the PEO samples with varying levels of crystallinity, neither of
permeability coefficients may be described by the following these assumptions can be tested. As a result, the estimated
110 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

Table 4
Penetrant permeation parameters in semi-crystalline and amorphous poly(ethylene oxide) and in amorphous polyethylene [22,23]
Gas PEO PE
a
PA,o (Barrer) m× 103 (atm−1 ) b
PA,a (Barrer) Ep (kJ/mole) PA,o (Barrer) PA,a b (Barrer)

He 1.7 ± 0.1 −4 ± 8 20 84 ± 3 4.9 34


H2 1.8 ± 0.1 −2 ± 7 21 76 ± 5
O2 0.68 ± 0.05 −0.4 ± 7 8.1 89 ± 6 2.9 18
N2 0.25 ± 0.02 −10 ± 7 3.0 95 ± 5 0.97 6.7
CO2 12 ± 1.0 34 ± 10 143 70 ± 7 13 78
CH4 0.60 ± 0.05 12 ± 8 7.1 133 ± 6 2.9 23
C2 H 4 2.9 ± 0.5 8 ± 22 34 50 ± 10
C 2 H6 1.6 ± 0.3 23 ± 21 19 75 ± 9 6.8 66
C3 H 6 3.8 ± 1.6 250 ± 90 45 58 ± 11 14 147
C 3 H8 1.4 ± 0.6 180 ± 70 17 70 ± 16 9.5 108
aInfinite dilution permeability in semi-crystalline PEO.
bEstimated infinite dilution amorphous phase permeability coefficients at 35 ◦ C in PEO and 25 ◦ C in PE. Permeability coefficients in amorphous phase
PE are calculated as the product of solubility coefficients in amorphous PE [22] and diffusion coefficients in amorphous PE estimated with Eq. (6) [23].

amorphous phase permeability coefficients should be re- CO2 permeability in amorphous phase PEO is within the
garded as rather crude estimates of the actual permeation range of values reported for block copolymers containing
properties of the amorphous material. These data are re- PEO (cf. Table 1). The CO2 /H2 pure gas selectivity in PEO
ported in Table 4. The permeability coefficients decrease in is 6.7 at infinite dilution and increases to 9.9 at 1.5 MPa
the following order: (14.7 atm), which is similar to values reported for polyether-
b-polyamide segmented block copolymers [17]. To empha-
CO2 > C3 H6 > C2 H4 > H2 ≈ He > C3 H8 size the effect that this difference in properties for CO2 has
≈ C2 H6 > CH4 ≈ O2 > N2 on separation, one can also consider the CO2 /N2 selectivity.
In PE, the pure gas CO2 /N2 selectivity at infinite dilution is
Except for CO2 and the olefins, this order of permeability approximately 12. However, in polar PEO, the CO2 /N2 pure
coefficients is typical for rubbery polymers [33]. Generally, gas selectivity at infinite dilution is 48, which is essentially
penetrants with higher critical temperature are more con- equivalent to the selectivity in block copolymers containing
densable and, therefore, more soluble; however, more con- PEO, as shown in Table 1. These results are consistent with
densable penetrants often have larger critical volume values enhanced solubility of quadrupolar CO2 in PEO.
(cf. Table 2), which reduces diffusion coefficients. Perme- It is also of interest to explore the difference between
ability reflects the tradeoff between these often conflicting olefin and paraffin transport in these two polymers because
contributions from solubility and diffusivity. of the importance of this separation to the petrochemical
The gas transport properties of polyethylene (PE) have industry [36]. In amorphous PE, propane and propylene have
been well studied [22,23]. Polyethylene provides an inter- very similar permeability coefficients, and the C3 H6 /C3 H8
esting material for comparison with PEO because PE is selectivity is only 1.4 (cf. Table 5). However, in amorphous
essentially PEO without the polar ether linkages. Several PEO, olefins exhibit significantly higher permeability than
physical properties of these two polymers are compared in their paraffin analogues. The pure gas C3 H6 /C3 H8 selectivity
Table 3. The amorphous phase FFV of PEO is significantly at infinite dilution and 35 ◦ C is 2.7, which is almost 100%
lower than that of PE, and the solubility parameter of PEO
is higher than that of PE. These trends are qualitatively
consistent with the much more polar nature of PEO. Table 5
Table 4 presents a comparison of infinite dilution perme- Estimated amorphous phase propylene permeability, diffusivity and sol-
ubility, and C3 H6 /C3 H8 selectivity in PE at 25 ◦ C and PEO at 35 ◦ C at
ability coefficients in amorphous PEO and PE. Unlike PEO, infinite dilution
literature values for tortuosity and chain immobilization fac-
Propylene sorption and Propylene–propane
tors are available for PE [23], and these data were used to transport properties selectivity
estimate amorphous phase permeability coefficients in PE.
PEb PEO PE PEO
With the exception of CO2 , PE exhibits higher permeability
values than PEO, even though the PE data were determined PA a 147 45 P= /P c 1.4 2.7
Da a × 107 D= /D c
at lower temperature (25 ◦ C, rather than 35 ◦ C for PEO). 3.2 1.7 1.5 1.6
SA a 3.5 2.0 S= /S c 0.89 1.7
CO2 is the only gas exhibiting higher permeability in PEO
a The units for permeability (P), diffusivity (D) and solubility (S) are
than in PE. Interestingly, CO2 is much more permeable than
Barrer, cm2 /s, and cm3 (STP)/(cm3 polymer atm), respectively.
all of the other gases considered, including H2 , which, in b Michaels and Bixler [22,23].
conventional glassy gas separation polymers, always has a c The subscripts “ ‘’ and “ ” represent propylene and propane, respec-
=
higher permeability coefficient than CO2 [4]. The estimated tively.
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 111

Table 6 At each temperature, with the exception of N2 at 25 ◦ C,


Effect of pressure and temperature on pure gas permeability coefficients the permeability coefficients were extrapolated to infinite
in semi-crystalline PEO
dilution by fitting the data to Eq. (15). Since there is only one
Pressure (atm) Permeability (Barrer) data point for N2 at 25 ◦ C, this value was used as the effective
25 ◦ C 35 ◦ C 45 ◦ C infinite dilution permeability coefficient for this case. This
should be a reasonable approximation since N2 permeability
N2
4.4 0.24 0.99
is essentially insensitive to pressure. Using these infinite
7.8 0.07 0.24 1.0 dilution permeability coefficients, the activation energy of
14.6 0.22 1.0 permeation was estimated as follows [37]:
 
O2 −Ep
4.4 0.26 0.68 2.5 PA = PAo exp (16)
RT
7.8 0.26 0.68 2.6
14.6 0.26 0.68 2.5 where PAo is a pre-exponential factor, R is the gas constant,
CH4 T is the absolute temperature, and Ep is the activation energy
4.4 0.63 3.8 of permeation. The results of this calculation are presented in
7.8 0.16 0.65 3.9 Table 4. Based on the solution–diffusion model, Ep is [37]:
14.6 0.19 0.70 3.8
H2
Ep = Ed + HS (17)
4.4 1.8 6.4
where Ed is the activation energy of diffusion and HS is
7.8 0.85 1.8 6.6
14.6 0.81 1.8 6.7 the enthalpy change of sorption. Sorption is typically con-
sidered to be comprised of two hypothetical thermodynamic
CO2
4.4 13 40
steps, i.e., condensation of the pure penetrant and mixing
7.8 8.1 15 46 of the hypothetical pure condensed penetrant with polymer
14.6 9.5 17 52 segments. Within the context of this picture, the above equa-
C2 H 4 tion can be rewritten as follows [37]:
4.4 1.4 3.0 5.3 Ep = Ed + Hcond + Hmix (18)
7.8 1.6 3.2 5.7
10 1.7 3.2 5.9 Activation energy of permeation values in PEO were posi-
C2 H 6 tive, so permeability coefficients increased with increasing
4.4 0.62 1.7 3.3 temperature. Generally, penetrant activation energy of dif-
7.8 0.78 1.9 3.5 fusion (Ed ) is positive and increases with increasing pene-
11.2 0.88 1.9 3.5
trant size [4]. Condensation is exothermic (i.e., Hcond <
C3 H6 0), and the magnitude of the enthalpy change on conden-
3.0 3.2 6.6 12
sation increases with increasing penetrant condensability,
5.0 5.6 8.6 14
7.8 8.3 11 17 as characterized by properties such as critical temperature.
The enthalpy of mixing (Hm ) can be negative or positive,
C3 H8
4.0 1.1 2.4 5.1
depending on the energetic interactions between the con-
6.4 1.6 3.0 5.9 densed penetrant and the polymer. If penetrant molecules
7.8 1.9 3.5 6.5 have favorable interactions with polymer segments, such as
dipole–quadrupole interaction, the mixing enthalpy can be
negative. As shown in Table 4, there is no monotonic trend
larger than that in PE. This result suggests the possibility of Ep values with penetrant size, which suggests that each
of specific interactions between the polar ether linkages and term in Eq. (18) may have an important influence on Ep .
olefins. Sorption data will be presented later to support this Based on the data presented in Table 6, pure gas selectiv-
hypothesis. ity values were calculated as a function of temperature and
The effect of temperature on permeation properties was pressure, and these data are recorded in Table 7. In general,
determined. Permeability coefficients were measured at 25, selectivity increases as temperature decreases. For example,
35 and 45 ◦ C. Table 6 presents representative permeability the pure gas selectivity of CO2 /H2 at 790 kPa (7.8 atm) is
coefficients at different temperatures and pressures. All pen- 6.9 at 45 ◦ C and increases to 9.5 at 25 ◦ C. These results sug-
etrants exhibit higher permeability at higher temperature. Af- gest that operation of a PEO membrane at lower temperature
ter experiments at 45 ◦ C, which was the highest temperature should result in better CO2 /H2 separation performance.
explored and was, therefore, closest to the melting point of
PEO, H2 permeability at 35 ◦ C was re-measured to confirm 4.3. Solubility
that PEO transport properties had not changed as a result of
performing experiments at 45 ◦ C. No significant difference Fig. 3a,b present sorption isotherms of various penetrants
in H2 permeability coefficients was observed. at 35 ◦ C in semi-crystalline PEO. For less soluble penetrants,
112 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

Table 7
Pure gas selectivity in semi-crystalline poly(ethylene oxide) as a function
of pressure and temperature
Pressure (atm) Pure gas selectivity

25 ◦ C 35 ◦ C 45 ◦ C

O2 /N2
4.4 2.8 2.5
7.8 3.7 2.8 2.6
14.6 3.1 2.5
CO2 /N2
4.4 55 40
7.8 140 63 48
14.6 79 52
CO2 /CH4
4.4 21 10
7.8 51 23 12
14.6 50 25 13
CO2 /H2
4.4 7.4 6.2
7.8 9.5 8.4 6.9
14.6 12 9.9 7.7
C2 H4 /C2 H6
4.4 2.3 1.8 1.6
7.8 2.4 1.7 1.6
11.2 1.9 1.7 1.7
C3 H6 /C3 H8
4.4 3.9 3.0 2.6
7.8 4.4 3.1 2.7

such as O2 , N2 , CH4 , C2 H6 and C2 H4 , the isotherms are lin-


ear and solubility, SA , is constant. We represent the average
solubility for these gases as S∞ . The more soluble pene-
trants, such as C3 H6 , C3 H8 and CO2 , exhibit slightly convex
isotherms which may be described by the Flory–Huggins
theory [38]: Fig. 3. (a) Sorption isotherms in semi-crystalline PEO at 35 ◦ C. The
scale for gas sorption values is shown on the left y-axis for all gases
p
ln = lnΦ + (1 − Φ) + χ(1 − Φ)2 (19) except CO2 , whose scale is on the right y-axis. (b) Sorption isotherms in
p0 semi-crystalline PEO at 35 ◦ C.

where p0 (atm) is the penetrant saturation vapor pressure


at the temperature of the sorption experiment; it is esti-
mated from the Antoine or Wagner equation [39]. χ is the Using Eq. (21), infinite dilution solubility coefficients, S∞ ,
Flory–Huggins interaction parameter, and Φ is the volume of CO2 , C3 H6 and C3 H8 in semi-crystalline PEO were esti-
fraction of dissolved gas in the amorphous phase of the poly- mated. These values, along with infinite dilution solubility
mer, which is given by: coefficients for the other gases are recorded in Table 8. Gen-
erally, solubility coefficients increase with increasing critical
CV̄ temperature. CO2 and the olefins (ethylene and propylene)
Φ= (20)
φa + CV̄ show higher solubility coefficients than expected based on
their condensability (i.e., critical temperature). The solubil-
where φa is amorphous phase volume fraction in the poly- ity data were corrected to an amorphous basis using Eq. (5).
mer, and V̄ is the partial molar volume of the penetrant Fig. 4 presents the estimated solubilities at infinite dilution
(cm3 /cm3 (STP)), which is approximated as the mean value in amorphous PEO at 35 ◦ C as a function of penetrant critical
of the partial molar volume data reported by Kamiya et al. temperature. These data are compared with solubility values
[38]. χ values were obtained by fitting experimental mea- in amorphous PE at 25 ◦ C [22]. For permanent gases (N2 ,
surements of C as a function of p to Eqs. (19) and (20). At O2 , and CH4 ), PEO has similar solubility values to those in
infinite dilution, Eqs. (19) and (20) can be simplified to: PE if the uncertainty in solubility is considered. Therefore,
φa the higher permeability coefficients of permanent gases in
S∞ = exp(−1 − χ) (21)
p0 V̄ PE are due to higher diffusion coefficients.
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 113

Table 8 Fig. 4 further demonstrates the extraordinarily high level


Solubility coefficients of gases in semi-crystalline poly(ethylene oxide) at of CO2 sorption in PEO relative to solubilities of penetrants,
35 ◦ C
such as C2 H6 , that have similar critical temperatures but
Gas p0 (atm) V̄ c × 103 χ S∞ cannot participate in specific interactions with the polymer
(cm3 /cm3 (STP)) (cm3 (STP)/(cm3 atm))
matrix. In PE, for example, the CO2 /C2 H6 solubility selec-
N2 0.036 ± 0.018
tivity is 0.35. However, in more polar PEO, the CO2 /C2 H6
O2 0.041 ± 0.017
CH4 0.078 ± 0.037 solubility selectivity is 3.1, which is nearly an order of
C2 H 4 0.17 ± 0.02 magnitude higher than that observed in PE. This result sug-
C2 H 6 0.12 ± 0.02 gests that CO2 experiences favorable interactions with the
CO2 81.9a 2.01 0.56 0.37 ± 0.02 ether groups in PEO. The interaction between quadrupolar
C3 H6 14.68b 3.26 1.4 0.58 ± 0.02
CO2 and polar groups in polymers is well known [41]. Van
C3 H8 12.1b 3.57 2.0 0.34 ± 0.01
Amerongen studied the effect of polar acrylonitrile content
aThis value is hypothetical, estimated from the Wagner equation [33].
b
in butadiene/acrylonitrile copolymers on CO2 solubility and
Calculated using the Wagner equation for C3 H8 , and the Antoine
found that polymers with higher concentrations of polar
equation for C3 H6 [39].
c Kamiya et al. [38]. acrylonitrile had higher CO2 sorption [42]. Koros suggested
that the solubility selectivity for the CO2 /CH4 system in-
creases as the mass density of polar carbonyl or sulfone
The infinite dilution CO2 solubility coefficient is groups in the polymer system increases [43]. Bondar et al.
1.3 cm3 (STP)/(cm3 atm) at 35 ◦ C in amorphous PEO. This studied gas sorption in a series of copolymer (Pebax)
result is compared with the result of Bondar et al. [16], who containing polyether segments; they found that solubility
studied gas sorption in a series of phase separated block selectivity of carbon dioxide over nonpolar gases increases
copolymers containing PEO segments. Using solubility data as ether linkage concentration in the polymer increases
in copolymers, they estimated, by extrapolation, a value of [16].
1.4 cm3 (STP)/(cm3 atm) for the solubility of CO2 in wholly Interestingly, olefins exhibit higher solubility than their
amorphous PEO, which is in excellent agreement with the paraffin analogs in PEO. Often, an olefin (e.g., propy-
value determined in this study. High pressure phase equilib- lene) exhibits lower solubility than its paraffin analog
ria data for CO2 and poly(ethylene glycol) (PEG) with an av- (e.g., propane) in polymers (e.g., PE) due to the slightly
erage molecular weight of 200 have also been reported [40]. lower critical temperature of the olefin. However, PEO
At 40 ◦ C and CO2 pressure of 5.71 MPa, CO2 solubility in exhibits the opposite behavior. As shown in Table 5, the
PEG was 1.3 cm3 (STP)/(cm3 atm), which is also consistent propylene–propane solubility selectivity at infinite dilution
with our estimate of CO2 solubility in amorphous PEO. is 1.7, which, together with a propylene–propane diffusivity
selectivity of 1.6, gives an overall selectivity of 2.7. On
the other hand, in amorphous PE, the propylene–propane
solubility selectivity is approximately a factor of two lower
than that in PEO, which diminishes the propylene–propane
permeability selectivity to 1.4. Both polymers have sim-
ilar diffusivity selectivity values (1.5 in PE and 1.6 in
PEO). Additionally, the C2 H4 /C2 H6 solubility selectivity
in PEO is 1.4, even though C2 H4 has a lower Tc value
and, on this basis, would be expected to be less soluble
than C2 H6 . Together, these results indicate that olefins un-
dergo favorable interaction with the polar ether oxygen
in PEO.
Olefin solubility is also enhanced in liquids such as
tetrahydrofuran, a cyclic ether. Polar organic solvents, such
as tetrahydrofuran (THF), are Lewis acids and, as such, they
can interact with Lewis bases, such as the π electrons in
aromatic hydrocarbons or olefins [44,45]. Therefore, such
polar solvents form solubility-enhancing complexes with
olefins but not with paraffins [44,45].
To better illustrate the interesting behavior of olefin sorp-
tion in polar and nonpolar polymers, olefin and paraffin sol-
ubilities in a polar solvent were compared with those in a
Fig. 4. Estimated amorphous, infinite dilution solubility coefficients in
poly(ethylene oxide) (35 ◦ C) (䊉) and poly(ethylene) (25 ◦ C) (䊊) [22]. The nonpolar solvent. THF was selected as the polar solvent since
line represents the best fit of the poly(ethylene) data to a model in which it has an ether oxygen, and its interactions with aromatic
the logarithm of solubility increases linearly with critical temperature [4]. hydrocarbons have been documented [44]. Hexane, a well-
114 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

Table 9
Diffusion coefficient comparison in semi-crystalline (DA ) and amorphous
(DA,a ) poly(ethylene oxide) at 35 ◦ C and poly(ethylene) at 25 ◦ C

Gas PEO PE

DA × 107 DA,a × 107 DA × 107 DA,a × 107


(cm2 /s) (cm2 /s) (cm2 /s) (cm2 /s)
O2 1.3 ± 0.5 4.5 4.6 18
N2 0.53 ± 0.27 1.8 3.2 12
CO2 2.5 ± 0.2 8.6 3.72 13
CH4 0.58 ± 0.28 2.0 1.93 8.6
C 2 H4 1.3 ± 0.2 4.5
C2 H6 1.0 ± 0.2 3.4 0.68 3.9
C3 H6 0.50 ± 0.20 1.7 0.58 3.2
C 3 H8 0.31 ± 0.10 1.1 0.32 2.1

4.4. Diffusivity

Based on Eqs. (2), (6) and (7), amorphous phase diffu-


sion coefficients for various penetrants at infinite dilution
in PEO at 35 ◦ C were estimated and are given in Table 9,
where the gases are arranged in order of increasing criti-
cal volume. The diffusivity in PEO was also compared with
that in amorphous phase PE at 25 ◦ C [23]. Generally, diffu-
sion coefficients decrease with increasing critical volume ex-
cept for CO2 which, due to its strongly non-spherical shape,
presents a significantly smaller effective cross-section for
diffusion than indicated by its critical volume. Ethylene also
shows similar behavior; its diffusivity is higher than that of
methane, despite ethylene’s larger critical volume. Gener-
ally speaking, diffusion coefficients are higher in PE than in
PEO. However, there are several complicating factors that
make a direct comparison difficult. First, the data for PE
and PEO are at different temperatures and different crys-
tallinities. Additionally, diffusivity is very sensitive to the
Fig. 5. (a) Effect of temperature on C2 H4 and C2 H6 solubility in liquid
distribution of inter-crystalline spacings in semi-crystalline
tetrahydrofuran (THF) [47]. (b) Effect of temperature on C2 H4 and C2 H6 polymers [23], which is dependent on the solid film prepa-
solubility in liquid n-hexane [48,49]. ration process. The crystallite structure in the PE sample
was not discussed in the paper in which the transport data
were reported [23]. Nevertheless, it is unlikely that the PEO
studied in this paper would necessarily have similar macro-
studied paraffin, was chosen as a typical nonpolar hydrocar- crystalline structure. For example, polyethylene crystallites
bon solvent. Fig. 5a presents the solubilities of C2 H4 and have dimensions in the 4–30 nm range [46], while PEO may
C2 H6 in THF as a function of temperature. The solubilities contain macrospherulites with diameters up to 2 cm [24].
of C2 H4 and C2 H6 in n-hexane as a function of temperature Therefore, tortuosity and chain immobilization factors char-
are illustrated in Fig. 5b. In THF, C2 H6 has somewhat higher acterizing the effect of crystallites on penetrant diffusion
solubility at higher temperature. However, C2 H4 solubility could well be different in PE and PEO. Since, in this study,
increases more rapidly than that of C2 H6 as temperature we do not have tortuosity and chain immobilization factors
decreases. As a result, at 0 ◦ C, C2 H4 is significantly more for different penetrants in PEO, it is impossible to directly
soluble than C2 H6 in this polar, ether-containing solvent. In compare amorphous phase diffusivity in these two polymers.
contrast, over the entire range of temperatures investigated, Local effective diffusion coefficients, Deff , characterizing
C2 H6 is more soluble than C2 H4 in n-hexane, consistent with the penetrant diffusivity in the polymer at a penetrant con-
the higher critical temperature of C2 H6 . These results sug- centration of C2 , can be evaluated using the following stan-
gest that olefins are more soluble than paraffins in polar liq- dard equation [33]:
uids such as THF as long as the temperature is low enough to    
allow specific interactions to play a dominant role in the ther- dPA dp
Deff (C2 ) = PA + p (22)
modynamic interactions between the olefin and the liquid. dp p2 dC2 p2
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 115

5. Conclusion

Permeation and sorption properties of a variety of gases


were determined in poly(ethylene oxide), a polar rubbery
polymer. Relative to its nonpolar analog, polyethylene, PEO
exhibits much higher CO2 /H2 pure gas selectivity due to
high CO2 /H2 solubility selectivity, which is qualitatively
consistent with the notion that polar, basic ether linkages
interact favorably with acidic penetrants such as CO2 , thus
increasing CO2 sorption. Interestingly, PEO is more perme-
able to olefins than to paraffins due, in part, to enhanced
olefin–paraffin solubility selectivity. Presumably, the double
bonds in the olefins interact favorably with the polar ether
linkages, and paraffins cannot access this interaction.

Fig. 6. Local effective diffusion coefficient in semi-crystalline poly(ethy-


Acknowledgements
lene oxide) at 35 ◦ C as a function of penetrant concentration. The lines
were calculated using Eqs. (22) and (24) for C2 H4 and C2 H6 , and using
Eqs. (22) and (25) for CO2 , C3 H6 and C3 H8 . The authors gratefully acknowledge partial support of
this project by the United States Department of Energy
under grant number DE-FG02-99ER14991. This research
Substituting Eq. (15) in Eq. (22) results in the following work was also partially supported with the funding from
expression for the local diffusivity: the United States Department of Energy’s National Energy
  Technology Laboratory under a subcontract from Research
dp
Deff (C2 ) = PA,o (1 + 2mp2 ) (23) Triangle Institute through their Prime Contract No.: DE-
dC2 p2
AC26-99FT40675.
The solubility of less soluble penetrants is independent of
pressure, which leads to:
  Appendix A. List of symbols
dp 1
= ∞ (24)
dC2 p2 S A film area available for gas transport (cm2 )
C2 concentration of gas A dissolved in the
For CO2 , C3 H6 and C3 H6 , the following equation can be
polymer (cm3 (STP)/cm3 )
derived from Eqs. (19) and (20):
(dp1 /dt)leak rate of pressure rise in the downstream
    volume during leak test (cm Hg/s)
eχΦ −2χΦ−Φ
2
dp
= 2χΦ2 − 2χΦ − Φ + 1 (dp1 /dt)ss steady-state rate of pressure rise in the
dC2 p2 S∞ downstream volume (cm Hg/s)
 2 DA average effective diffusivity of gas A (cm2 /s)
1
× (25) DA,a diffusion coefficient in the amorphous
1 + C2 V̄ /φa
polymer (cm2 /s)
Fig. 6 presents calculated effective diffusion coefficients in DA /DB diffusivity selectivity of gas A to gas B
semi-crystalline PEO as a function of local concentration Deff local effective diffusion coefficient (cm2 /s)
for several hydrocarbons and CO2 . For all of the penetrants Ed activation energy of diffusion (kJ/mole)
shown, Deff increases with increasing local concentration, Ep activation energy of permeation (kJ/mole)
suggesting that these penetrants plasticize the polymer ma- H melting enthalpy, calorimetry data (J/g)
trix. This behavior is consistent with our previous argument Hc melting enthalpy of 100% crystalline
that higher penetrant contacting pressure can increase per- polymer (J/g)
meability by enhancing penetrant average diffusivity through Hcond enthalpy of condensation (kJ/mole)
the film. Interestingly, poly(dimethylsiloxane) (PDMS), Hm enthalpy of mixing (kJ/mole)
which does not exhibit a particular affinity for CO2 , exhibits HS enthalpy change of sorption (kJ/mole)
plasticization to C2 H6 and C3 H8 but not to CO2 , even at l film thickness (cm)
CO2 concentrations as high as 80 cm3 (STP)/cm3 [33,34]; m adjustable constant in Eq. (15) (atm−1 )
on the other hand, PEO was significantly plasticized by CO2 MA polymer film weight in air (g)
even at concentrations of the order of 10 cm3 (STP)/(cm3 MB polymer film weight in the auxiliary liquid (g)
amorphous polymer). NA steady-state flux of gas A (cm3 (STP)/(cm2 s))
116 H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117

p pressure difference across the polymer [2] B.D. Bhide, S.A. Stern, Membrane process for the removal of acid
gases from natural gas. Part I. Process configurations and optimization
film (atm)
of operating conditions, J. Membr. Sci. 81 (1993) 209–237.
p1 downstream partial pressure of gas A [3] Facts and figures for the chemical industry, Chem. Eng. News 74
p2 upstream partial pressure of gas A (1996) 38.
PA permeability of a polymer to gas A (Barrer) [4] B.D. Freeman, I. Pinnau, Polymeric materials for gas separations,
PA,a estimated amorphous phase permeability at in: B.D. Freeman, I. Pinnau (Eds.), Polymer Membranes for Gas
and Vapor Separation, ACS Symposium Series, vol. 733, Washington
infinite dilution (Barrer)
DC, 1999, pp. 1–27.
PAo pre-exponential factor (Barrer) [5] K. Ghosal, R.T. Chen, B.D. Freeman, W.H. Daly, I.I. Negulescu,
PA,o infinite dilution permeability (Barrer) Effects of basic substituents on gas sorption and permeation in
p0 penetrant saturation vapor pressure (atm) polysulfone, Macromolecules 29 (1996) 4360–4369.
R gas constant [6] M. Kawakami, H. Iwanaga, Y. Hara, M. Iwamoto, S. Kagawa, Gas
S∞
permeabilities of cellulose nitrate/poly(ethylene glycol) blend mem-
infinite dilution solubility (cm3 /(cm3 atm)) branes, J. Appl. Polym. Sci. 27 (1982) 2387–2393.
SA observed solubility of gas A in a polymer [7] M. Kawakami, H. Iwanaga, Y. Yamashita, M. Yamasaki, M. Iwamoto,
(cm3 (STP)/(cm3 atm)) S. Kagawa, Enhancement of carbon dioxide permselectivity of im-
SA,a estimated solubility of gas A in wholly mobilized liquid polyethylene glycol membrane by addition of metal
amorphous polymer (cm3 (STP)/(cm3 atm)) salts, Nippon Kagaku Kaishi 6 (1983) 847–853.
[8] J. Li, K. Nagai, T. Nakagawa, S. Wang, Preparation of polyethyleneg-
SA /SB solubility selectivity of gas A to gas B lycol (PEG) and cellulose acetate (CA) blend membranes and their
T absolute temperature (K) gas permeabilities, J. Appl. Polym. Sci. 58 (1995) 1455–1463.
Vd downstream volume (cm3 ) [9] K. Okamoto, M. Fujii, S. Okamyo, H. Suzuki, K. Tanaka, H. Kita,
V specific volume of amorphous polymer Gas permeation properties of poly(ether imide) segmented copoly-
mers, Macromolecules 28 (1995) 6950–6956.
at a temperature (cm3 /g)
[10] M. Yoshino, K. Ito, H. Kita, K. Okamoto, Effects of hard-segment
V̄ penetrant partial molar volume polymers on CO2 /N2 gas separation properties of poly(ethylene
(cm3 /cm3 (STP)) oxide)-segmented copolymers, J. Polym. Sci. Part B: Polym. Phys.
Vo specific volume of amorphous polymer 38 (2000) 1707–1715.
at 0 K (cm3 /g) [11] G. Chatterjee, A.A. Houde, S.A. Stern, Poly(ether urethane) and
poly(ether urethane urea) membranes with high H2 S/CH4 selectivity,
J. Membr. Sci. 135 (1997) 99–106.
Greek letters [12] H. Suzuki, K. Tanaka, H. Kita, K. Okamoto, H. Hoshino, T. Yoshi-
αA/B permeability selectivity of gas A to gas B naga, Y. Kusuki, Preparation of composite hollow fiber membranes
β chain immobilization factor of poly(ethylene oxide)-containing polyimide and their CO2 /N2 sep-
χ Flory–Huggins interaction parameter aration properties, J. Membr. Sci. 146 (1998) 31–37.
[13] Y. Hirayama, Y. Kase, N. Tanihara, Y. Sumiyama, Y. Kusuki, K.
χc weight fraction crystallinity in the polymer
Haraya, Permeation properties to CO2 and N2 of poly(ethylene
φa amorphous phase volume fraction oxide)-containing and crosslinked polymer films, J. Membr. Sci. 160
of a polymer (1999) 87–99.
φc crystalline phase volume fraction [14] K. Tsutsui, H. Yoshimizu, Y. Tsujita, T. Kinoshita, Gas permeation
of a polymer properties of a composite membrane filled with poly(ethylene oxide)
into a porous membrane, J. Appl. Polym. Sci. 73 (1999) 2733–2738.
Φ volume fraction of dissolved gas
[15] I. Blume, I. Pinnau, US Patent 4,963,165 (1990).
in the amorphous polymer [16] V.I. Bondar, B.D. Freeman, I. Pinnau, Gas sorption and characteriza-
ρ observed film density tion of poly(ether-b-amide) segmented block copolymers, J. Polym.
ρa density of amorphous polymer (g/cm3 ) Sci. Part B: Polym. Phys. 37 (1999) 2463–2475.
ρc crystal density calculated from unit [17] V.I. Bondar, B.D. Freeman, I. Pinnau, Gas transport properties of
poly(ether-b-amide) segmented block copolymers, J. Polym. Sci. Part
cell dimensions (g/cm3 ) B: Polym. Phys. 38 (2000) 2051–2062.
τ tortuosity factor [18] J. Kim, S. Ha, Y. Lee, Gas permeation of poly(amide-6-b-ethylene
oxide) copolymer, J. Membr. Sci. 190 (2001) 179–193.
Subscripts [19] Y. Hirayama, N. Tanihara, Y. Kusuki, Y. Kase, K. Haraya, K.
a amorphous phase Okamoto, Permeation properties to hydrocarbons, perfluorocarbons
and chlorofluorocarbons of cross-linked membranes of polymethacry-
A gas A late with poly(ethylene oxide) and perfluorononyl moieties, J. Membr.
B gas B Sci. 163 (1999) 373–381.
c crystalline phase [20] J. Kim, Y. Lee, Gas permeation properties of poly(amide-6-b-ethylene
1 downstream side oxide)—silica hybrid membranes, J. Membr. Sci. 193 (2001) 209–
2 upstream side 225.
[21] J. Sanchez, C. Charmett, P. Gramain, Poly(ethylene oxide-co-
epichlorohydrin) membranes for carbon dioxide separation, J. Membr.
Sci. 205 (2002) 259–263.
References [22] A.S. Michaels, H.J. Bixler, Solubility of gases in polyethylene, J.
Polym. Sci. 50 (1961) 393–412.
[1] J.J. McKetta, W.A. Cunningham, Encyclopedia of Chemical Process- [23] A.S. Michaels, H.J. Bixler, Flow of gases through polyethylene, J.
ing and Design, vol. 6, Marcel Dekker, New York, 1978, pp. 292–310. Polym. Sci. 50 (1961) 413–439.
H. Lin, B.D. Freeman / Journal of Membrane Science 239 (2004) 105–117 117

[24] F.E. Bailey, J.V. Koleske, Poly(ethylene oxide), Academic Press, [38] Y. Kamiya, Y. Naito, K. Mizoguchi, K. Terada, G.A. Mortimer, Ther-
New York, 1976. modynamic interactions in rubbery polymer/gas systems, J. Polym.
[25] P. Zoller, D. Walsh, Standard Pressure–Volume–Temperature Data for Sci. Part B: Polym. Phys. 35 (1997) 1049–1053.
Polymers, 1st ed., Technomic Publishing Co. Inc., Lancaster, 1995. [39] R.C. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and
[26] J. Brandrup, E.H. Immergut, E.A. Grulke, Polymer Handbook, 4th Liquids, McGraw-Hill, New York, 1987.
ed., John Wiley & Sons Inc., New York, 1999. [40] D. Gourgouillon, M. Nunes da Ponte, High pressure phase equilibria
[27] S. Sunderrajan, B.D. Freeman, C.K. Hall, I. Pinnau, Propane for poly(ethylene glycol)s + CO2 : experimental results and modeling,
and propylene sorption in solid polymer electrolytes based on Phys. Chem. Chem. Phys. 1 (1999) 5369–5375.
poly(ethylene oxide) and silver salts, J. Membr. Sci. 182 (2001) 1–12. [41] S.G. Kazarian, M.F. Vincent, F.V. Bright, C.L. Liotta, C.A. Eckert,
[28] Z. Mogri, D.R. Paul, Membrane formation techniques for gas perme- Specific intermolecular interaction of carbon dioxide with polymers,
ation measurements for side-chain crystalline polymers, J. Membr. J. Am. Chem. Soc. 118 (1996) 1729–1736.
Sci. 175 (2000) 253–265. [42] G.J. van Amerongen, Diffusion in elastomers, Rubber Chem. Technol.
[29] P.R. Bevington, D.K. Robinson, Data Reduction and Error Analysis 37 (1964) 1065–1152.
for the Physical Sciences, 2nd ed., McGraw-Hill, Inc., New York, [43] W.J. Koros, Simplified analysis of gas/polymer selective solubility
1992. behavior, J. Polym. Sci. Part B: Polym. Phys. Ed. 23 (1985) 1611–
[30] F.T. Simon, J.M. Rutherford, Crystallization and melting behavior of 1628.
polyethylene oxide copolymers, J. Appl. Phys. 35 (1964) 82–86. [44] J.M. Prausnitz, R.N. Lichtenthaler, E.G. de Azevedo, Molecular Ther-
[31] W.M. Lee, Selection of barrier materials from molecular structure, modynamics of Fluid-Phase Equilibria, 3rd ed., Prentice Hall, New
Polym. Eng. Sci. 20 (1980) 65–69. Jersey, 1999.
[32] D.W. VanKrevelen, Properties of Polymers: Their Correlation with [45] R.V. Orye, R.F. Weimer, J.M. Prausnitz, Lewis acidity of polar
Chemical Structure. Their Numerical Estimation and Prediction from organic solvents from thermodynamic measurements, Science 148
Additive Group Contributions, Elsevier, Amsterdam, 1990, p. 75. (1965) 74–75.
[33] T.C. Merkel, V.I. Bondar, K. Nagai, B.D. Freeman, I. Pinnau, Gas [46] B. Wunderlich, Macromolecular Physics: Crystal Structure, Morphol-
sorption, diffusion, and permeation in poly(dimethylsiloxane), J. ogy, Defects, vol. 1, Academic Press, New York, 1973, p. 399.
Polym. Sci. Part B: Polym. Phys. 38 (2000) 415–434. [47] F. Gibanel, M.C. Lopez, F.M. Royo, J. Santafe, J.S. Urieta, Solubility
[34] W.J. Koros, M.W. Hellums, Gas separation membrane material se- of nonpolar gases in tetrahydrofuran at 0–30 ◦ C and 101.33 kPa
lection criteria: differences for weakly and strongly interacting feed partial pressure of gas, J. Solut. Chem. 22 (1993) 211–217.
components, Fluid Phase Equilib. 53 (1989) 339–354. [48] J.A. Waters, G.A. Mortimer, H.E. Clements, Solubility of some light
[35] A. Singh, B.D. Freeman, I. Pinnau, Pure and mixed gas hydrocarbons and hydrogen in some organic solvents, J. Chem. Eng.
acetone/nitrogen permeation properties of polydimethylsiloxane Data 15 (1970) 174–176.
[PDMS], J. Polym. Sci. Part B: Polym. Phys. 36 (1998) 230–289. [49] J.A. Waters, G.A. Mortimer, Some solubility data for ethane in n-
[36] R.B. Eldridge, Olefin/paraffin separation technology: a review, Ind. hexane, J. Chem. Eng. Data 17 (1972) 156–157.
Eng. Chem. Res. 32 (1993) 2208–2212. [50] K.J. Ivin, Transitions and Relaxation, in: H.F. Mark, N.M. Bikales
[37] S.V. Dixon-Garrett, K. Nagai, B.D. Freeman, Ethybenzene solubility, (Eds.), Encyclopedia of Polymer Science and Technology, vol. 2
diffusivity, and permeability in poly(dimethylsiloxane), J. Polym. Sci. (supplement), John Wiley & Sons Inc., New York, 1977, pp. 745–
Part B: Polym. Phys. 38 (2000) 1461–1473. 839.

S-ar putea să vă placă și