Anda di halaman 1dari 13

Journal of Membrane Science 453 (2014) 7183

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Contributions of diffusion and solubility selectivity to the upper bound


analysis for glassy gas separation membranes
Lloyd M. Robeson a,n, Zachary P. Smith b, Benny D. Freeman b, Donald R. Paul b
a
Lehigh University, 1801 Mill Creek Road, Macungie, PA 18062, USA
b
Department of Chemical Engineering, Texas Materials Institute, Center for Energy and Environmental Research, The University of Texas at Austin,
10100 Burnet Road, Bldg. 133, Austin, TX 78758, USA

art ic l e i nf o a b s t r a c t

Article history: Prior analyses of the upper bound of permselectivity versus permeability, both theoretical and empirical,
Received 26 August 2013 have assumed that this relationship is a consequence of the dependence of gas diffusion coefcients on
Received in revised form the molecular diameter of the gases of interest. The solubility selectivity has been assumed to be
23 October 2013
invariant with permeability (and free volume). However, a few literature sources note that the solubility
Accepted 28 October 2013
coefcient for specic families of glassy polymers correlate with free volume. A large database of
Available online 4 November 2013
permeability, diffusivity and solubility coefcients for glassy polymers was compiled to investigate this
Keywords: hypothesis. A critical analysis of the data demonstrates a modest solubility selectivity contribution to
Upper bound permselectivity as a function of free volume and, thus, permeability. The solubility selectivity (Si/Sj)
Diffusion selectivity
generally decreases with increasing permeability (and free volume) when the diameter of gas j is larger
Solubility selectivity
than that of gas i. This empirical trend is likely a consequence of larger gas molecules having less access
Permselectivity
than smaller molecules to sorption sites as the polymer packing density increases and free volume
decreases. The diffusion data permit determination of a diffusivity upper bound, which is modestly
different from the permeability-based upper bound relationship. The diffusion data analysis allows a
determination of a new set of gas diameters more appropriate for gas diffusion in polymers than prior
correlations.
& 2013 Elsevier B.V. All rights reserved.

1. Introduction having the following mathematical form [12]:


P i kup iju where iju P i =P j
n up
1
The separation of gases by polymeric membranes has been an
active area of research and industrial use for several decades [1]. On a loglog plot of ij(u) versus Pi, polymers with the highest
In 1979, Monsanto (now Permea, owned by Air Products and separation efciency and highest gas throughput are bounded by a
Chemicals, Inc.) developed the rst commercial asymmetric hollow linear relationship given by Eq. (1) that is commonly referred to as
ber membrane module for gas separation; the ber was based on the upper bound. There are virtually no data above this line. This
polysulfone, and the rst application was for hydrogen recovery empirical upper bound has been predicted from fundamental
(H2/N2) in the ammonia synthesis process [26]. Polymer mem- principles as will be discussed later [13]. Between 1991 and
branes have since been commercialized for separating a number of 2008, as more permeability data became available, modest shifts
other common gas pairs such as O2/N2, CO2/CH4, H2/CH4, He/air, in the empirical upper bound became apparent for several gas
and for the dehydration of air and natural gas [2]. In general, pairs [14]. Correlating permeability data of Pi versus Pj allowed the
polymer membrane research focuses on identifying materials with determination of the kinetic diameters of He, H2, O2, N2, CO2, and
high combinations of permeability and permselectivity. Permselec- CH4 [15]. Correlations of Pi versus Pj with free volume [16,17] and
tivity is dened as the ratio of the permeability coefcient of the activation energies of permeation and diffusion [18] have been
faster permeating gas, i, to that of the slower permeating gas, j noted in the literature. Furthermore, the effect of temperature on
(i.e., Pi/Pj). However, it was recognized that there is a trade-off the upper bound was recently predicted and showed good agree-
between Pi/Pj and Pi [711]. A large body of experimental evidence ment with the limited data available [19]. Group contribution
empirically suggests an upper bound to this trade-off relationship approaches have also been employed for predicting gas perme-
ability and permselectivity from information on polymer structure
[2022].
n
Corresponding author. Tel.: 1 610 481 0117. As noted, two somewhat different fundamental approaches
E-mail address: lesrob2@verizon.net (L.M. Robeson). have been employed to describe permeability and permselectivity

0376-7388/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.memsci.2013.10.066
72 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

behavior. The approach of Freeman [13] combined fundamental polymers. The vast majority of the data from the literature deter-
principles involving transition state theory for diffusion and mined D from time-lag measurements and calculated S from the
thermodynamic relationships for solubility to predict the upper relationship P DS. In a few cases, S was determined directly from
bound values of nup and kup. The approach of Alentiev and sorption measurements, and D was calculated from PDS. For a
Yampolskii [16,17] involved free volume models applied to diffu- given pressure, the effective diffusion coefcients for glassy poly-
sion and permeability for gas/polymer systems. It is well estab- mers will have slightly different values depending on whether they
lished that diffusion coefcients are related to free volume, and to were determined from time-lag measurements or from DP/S [36].
a lesser extent solubility coefcients are affected by free volume; A detailed discussion of these differences for polysulfone and a
as a result, permeability coefcients are strongly related to free polymer of intrinsic microporosity (PIM-1) is included in Appendix A
volume [20,2330]. Consequently, although the derived para- of this paper to illustrate the range of possibilities. The database
meters from these two approaches have different physical mean- includes over 800 different polymers from over 200 references. The
ings, the general form of the equations is equivalent. database is not as large as previously noted for the permeability
Correlations showing the separate effects of diffusion and correlations since many papers only report permeability results.
solubility coefcients on the relationship between permselectivity Data were compiled for 15 gas pairs involving He, H2, O2, N2, CO2
and permeability are more limited. The most signicant data in the and CH4.
literature have been summarized by Yampolskii and Alentiev The data are analyzed using the following basic equation:
[16,17], who employed an extensive database of P, D, and S values
Dj kd Dni d 3
in various polymers for the gases of interest.
In the expression where i and j are chosen such that nd 41. The data are presented
P i Di S i on loglog plots where Di (y-axis) and Dj (x-axis) yield the
2 following equation:
P j Dj S j
log kd log Dj
for a specic gas pair, the permselectivity (i.e., Pi/Pj) has diffusion log Di  4
nd nd
(i.e., Di/Dj) and solubility (i.e., Si/Sj) selectivity components. It is
generally assumed that as Pi (and free volume) decreases, the When plotted as Di (x-axis) and Dj (y-axis), the relevant equation is
increase in permselectivity is due to increases in diffusion selec-
log Dj log kd nd log Di 5
tivity, and solubility selectivity remains invariant with Pi (and free
volume); thus, the permeability upper bound relationship is Standard least squares ts to both Eq. (4) and to Eq. (5) were made.
controlled by diffusion selectivity [1214]. This assumption is If there is limited scatter in the data, similar results would be
inherent in the upper bound theory, which embodies solubility obtained from both equations; however, these results will differ
coefcients in the constant kup in Eq. (1). The determination of gas more signicantly when there is more scatter in the data. The
kinetic diameters from upper bound data [31] and from perme- previously published permeability correlation [15] had limited
ability data [15] assumed no changes in solubility selectivity as Pi scatter in the data, and a visual t was employed.
varied. A review of the visual t to the permeability data showed that
The assumption that solubility is independent of free volume in the average of the log k and n values from Eqs. (4) and (5) gave
the upper bound analysis needs to be critically addressed. In most virtually the same results as the reported visual t to the data.
cases, literature data show the diffusion coefcient of CO2 to be Since there is signicantly more scatter in the diffusion data for
lower than that of O2 even though the kinetic diameter of CO2 and some gas pairs, a visual t was not realistic; thus, the analysis
O2 are quite similar from upper bound [31] and permeability protocol employed the average of the values obtained by the two
correlations [15]. Additionally, the kinetic diameters determined methods.
from zeolites [32] show CO2 to have a smaller diameter than O2. If Glassy polymers exhibit dual-mode sorption behavior, which is
the solubility coefcient is a function of free volume, which is characterized by sorption isotherms that are concave to the
consistent with several correlations noted in the literature [27,33 pressure axis; this curvature is more pronounced for gases with
35], an accurate description of kup must be included in this high condensability [37,38]. Therefore, a pressure dependency of D
dependency. This paper demonstrates that solubility selectivity (i. and S arises when the sorption and permeation behavior exhibits
e., Si/Sj) and the solubility coefcients (i.e., Si) both vary with dual-mode behavior. The literature values of D and S in our
permeability (and, thus, free volume) leading to a solubility database involve some variation in the pressures used for mea-
selectivity contribution to permselectivity. The extent of this con- surement of P, D or S. While these variations will be expected to
tribution changes for each gas pair considered. Correlations in the yield quantitative differences in our analysis, the qualitative trends
literature for permeability, diffusivity and solubility often employ are not believed to be signicantly affected by dual-mode effects.
fractional free volume. In this paper, we refer to free volume and However, to illustrate the magnitude of the possible effects, an
fractional free volume trends as qualitatively equivalent. analysis of the pressure dependency on gas sorption and diffusion
for polysulfone and PIM-1 is included in Appendix A.
Examples of plots of diffusion coefcient data for several gas
2. Correlations of diffusion coefcients pairs are illustrated in Figs. 1,2 and 3. For gas pairs involving He or
H2, the data show considerable scatter. This is partly due to limited
The database for diffusion coefcients used in this analysis only data relative to the other gas pairs but primarily due to the larger
includes glassy polymers and comprises many of the literature experimental error involved with accurately determining very short
references employed for prior permeability [15] and permselectivity time-lag values [39]. Table 1 includes values of nd from Eqs. (4) and
[14] correlations, as well as including more recent references. The (5), the average nd values, the np values from the permeability
selection of data follows a similar protocol previously employed; correlation using this database, which employs P instead of D (i.e.,
n
namely, data were selected in the range of 2535 1C from polymers P j kp P i p ), and the np values from the previously employed
tested in the same laboratory and for gases tested on the same database [15]. The correlation of np values from this database agree
lm. The database also included solubility coefcients along with well with the much larger database used previously. Fig. 4 presents
permeability coefcients. Peruorinated glassy polymers were not the correlation between nd from diffusion data and np from
included due to their unique gas solubility relative to other permeability data. The correlation for gas pairs not containing CO2
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 73

Fig. 3. Correlation of H2 and N2 diffusion coefcient data.


Fig. 1. Correlation of O2 and N2 diffusion coefcient data. Line shown is the least
squares t to Eq. (4).

Table 1
Values of nd and np determined from the permeability and diffusion coefcient
databases.

Gas pair nd nd Average Permeability Permeability


from from nd correlation, correlation,
Eq. (4) Eq. (5) np (previous np (this
database)a,b database)b

O2/N2 1.16 1.14 1.15 1.16 1.12


H2/N2 1.66 1.41 1.55 1.52 1.48
He/N2 2.09 1.64 1.87 1.83 1.77
H2/CH4 2.01 1.71 1.86 1.75 1.68
He/H2 1.21 1.10 1.16 1.16 1.19
He/CH4 2.83 1.83 2.33 2.13 2.03
N2/CH4 1.18 1.14 1.16 1.14 1.12
He/O2 1.71 1.43 1.57 1.59 1.62
H2/O2 1.41 1.24 1.33 1.33 1.29
O2/CH4 1.39 1.31 1.35 1.31 1.26
CO2/CH4 1.26 1.17 1.21 1.30 1.23
CO2/N2 1.09 1.02 1.05 1.12 1.10
He/CO2 2.08 1.48 1.78 1.60 1.71
H2/CO2 1.76 1.48 1.62 1.35 1.37
O2/CO2 1.14 1.07 1.10 0.983 1.03

a
Ref. [15].
b n
Permeability correlation from P j kp P i p .

employed in this correlation. Shieh and Chung [40] noted that CO2
Fig. 2. Correlation of CO2 and CH4 diffusion coefcient data. Line shown is the least showed poor correlation with these values and proposed an
squares t to Eq. (4). p
effective gas diameter dened as sef f sk sc , which yielded an
excellent correlation for all gases including CO2. More recently,
is quite good except that the correlation line shows values of nd Dal-Cin et al. calculated gas diameters based on upper bound data,
from diffusion data to be uniformly higher than np values from sDC [31], and Robeson et al. calculated gas diameters based on
permeability data, which is not expected. Note that the permeability permeability correlations using a large permeability database, sPC
database [15] primarily consisted of glassy polymers although it did [15]. Analogous to our previous permeability correlation [15], gas
include several crystalline polymers, liquid crystalline polymers, and diameters have also been determined from the diffusion database
some rubbery polymers as well as peruorinated polymers. It is for this study, sD, and a discussion of these calculations will follow.
interesting to note that when CO2 is the j gas in the Di/Dj pair, the nd The values for these diameters are listed in Table 2.
values are all above the np data. Conversely, when CO2 is the i gas in Using the data in Table 2, values of dj =di 2 1 were calculated
the Di/Dj pair, the nd values are all below the np data. and compared with values of nd  1 in Table 3. These terms are
The theory and the experimental data indicate a relationship predicted to be equal to one another according to the Freeman
between diffusion coefcients and the square of the gas diameters. theory [13]. The results are shown in Fig. 5 for the three best
The kinetic diameter, sk, determined from zeolite measurements correlations. Both the Dal-Cin and effective diameter predictions
[32] and the Lennard-Jones collision diameter, sc, [40] are often yield slopes close to unity although there is considerable scatter in
74 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

Table 3
Values of dj =di 2  1 from Table 2.

Gas pair sk sc seff sDC sPC sD nd  1

O2/N2 0.107 0.151 0.132 0.131 0.136 0.167 0.151


H2/N2 0.586 0.588 0.593 0.581 0.540 0.587 0.551
He/N2 0.960 1.03 0.997 0.972 0.821 0.873 0.866
H2/CH4 0.729 0.711 0.726 0.850 0.763 0.899 0.861
He/H2 0.236 0.281 0.254 0.245 0.182 0.180 0.158
He/CH4 1.14 1.19 1.16 1.31 1.08 1.24 1.33
N2/CH4 0.0898 0.0775 0.0836 0.171 0.144 0.196 0.162
He/O2 0.771 0.768 0.764 0.744 0.603 0.604 0.570
H2/O2 0.433 0.378 0.407 0.398 0.355 0.360 0.326
O2/CH4 0.206 0.240 0.227 0.324 0.301 0.396 0.349
CO2/CH4 0.326  0.0880 0.102 0.283 0.318 0.231 0.214
CO2/N2 0.217  0.154 0.0166 0.0962 0.152 0.0293 0.0531
He/CO2 0.611 1.40 0.964 0.799 0.582 0.820 0.781
H2/CO2 0.304 0.877 0.567 0.442 0.338 0.542 0.621
O2/CO2  0.0903 0.360 0.114 0.0317  0.0131 0.134 0.102

Fig. 4. Data from Table 1 comparing nd and np values. See Eq. (3) and Table 1 for
denitions of these terms. The coefcients labeled np (previous database), which
were taken from the permeability database in reference [15], are based primarily
on glassy polymers, but the database also contains several crystalline polymers,
liquid crystalline polymers, some rubbery polymers, and peruoropolymers.
As shown by the equal value line, the coefcients labeled np (previous database)
match closely with the np coefcients from this study. The coefcients labeled nd
(this database) are consistently higher than the coefcients labeled np.

Table 2
Gas diameters determined from various techniques.

Gas sk () sc () seff () sDC () sPC () sD ()

He 2.60 2.58 2.59 2.555 2.644 2.55 70.10


H2 2.89 2.92 2.90 2.854 2.875 2.777 0.08
O2 3.46 3.43 3.44 3.374 3.347 3.23 70.02
N2 3.64 3.68 3.66 3.588 3.568 3.497 0.02
CO2 3.30 4.00 3.63 3.427 3.325 3.447 0.06
CH4 3.80 3.82 3.81 3.882 3.817 3.817

sk is the kinetic diameter [32], sc the collision diameter [40], seff the effective
diameter [40], sDC the Dal-Cin et al. correlation diameter [31], sPC the permeability
correlation diameter [15], and sD the diffusion correlation diameter.

the data. Diameters from the permeability correlation give a value


for the slope that is less than unity, as expected from the results Fig. 5. Correlation of nd  1 with the values of dj =di 2  1 from Table 3.

shown in Fig. 4.
The values of the gas diameters of CO2 in Table 2 appear to be technique.
inconsistent with diffusion data except in the case of the effective
 2
gas diameter, seff, the Dal-Cin diameter, sDC, and the diffusion dj
database for this study, sD. The comparison of D(O2) versus D(CO2) nd  1 1 6
di
shown in Fig. 6 illustrates that virtually all the diffusion coef-
cients for O2 are higher than those for CO2. Thus, the effective CO2 For this analysis, the diameter of CH4 was xed at 3.817 , and
gas diameter is clearly larger than that of O2. The permeability the diameters of He, H2, O2, CO2, and N2 were calculated. Our
correlations from this paper and the previous correlation [15] reason for xing the diameter of CH4 is that this molecule is nearly
show the gas diameter values to be virtually the same. This result spherical, and there is little variation in reported diameters for CH4
is a consequence of a modest solubility selectivity contribution to in the literature [15].
permeability as the free volume changes. Fig. 7 shows the correlation between nd  1 and dj =di 2  1.
Gas diameters were determined from diffusion correlations The line of best t, which was forced through the origin, is
analogous to the method used for permeability correlations [15]. nd  1 0:997dj =di 2  1. Single prediction error bands, which
An average value of nd was determined by tting Eqs. (4) and (5) are shown in gray, are used to visually demonstrate the goodness
to the database of diffusion coefcients. By assuming that one of t. The bands shown by the long, dashed lines represent a
gas diameter is known, the diameters of the other gases can condence interval of 68%, and bands shown by the short dashes
be determined from Eq. (6) using a least squares minimization indicate a condence interval of 95%.
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 75

Fig. 6. Correlation of O2 and CO2 diffusion coefcient data. Line noted is least Fig. 7. Determining gas diameters by minimizing error in Eq. (6). Data for nd  1 are
squares t of Eq. (4). from Table 3, dCH4 is xed at 3.817 , and values of di and dj are allowed to vary to
minimize the sum of squares. The solid, red line shows the best t to the data
(y 0.997x), the long, dashed, gray lines show a single prediction band with
A list of calculated gas diameters is presented in Table 2. These condence intervals of 68%, and the short, dashed, gray lines show the single
diameters represent the best t to dj and di as calculated from Eq. prediction band with condence intervals of 95%. Uncertainties represent one
(6). Uncertainties were determined by varying chi squared values standard deviation. (For interpretation of the references to color in this gure
near their minimum for each gas diameter [41]. Each uncertainty legend, the reader is referred to the web version of this article.)
listed in Table 2 represents one standard deviation.
The largest uncertainties in gas diameter are for He and H2.
These gases have higher diffusion coefcients and lower gas
solubility coefcients than the other gases listed in Table 2; of kd from the following relationship:
therefore, it is reasonable to expect that there is more inherent   
D S S 1a
error associated with determining diffusion coefcients for these ln ij ln i ln i  ij ln Di ln i  ij b  f 7
Dj Sj Sj RT
gases. The larger error for He and H2 gas diameters can be
observed in Fig. 7, where the largest error bars are always where b11.5 cm2/s for glassy polymers, a has a universal value of
associated with gas pairs containing He, H2, or both gases. 0.64, T is chosen to be 308 K and f is considered an adjustable
Additionally, the relative size of gas diameters can be compared parameter. Values of f from the upper bound theory and the
to those reported from other correlations. For this analysis, gas permeability correlation study were both found to be 12,600 cal/
diameters increase in the following order: Heo H2 oO2 oCO2 o mol. Furthermore, ij dj =di 2 1, where d is the gas diameter of
N2 o CH4. This order is identical to that reported for Dal-Cin and gas i or gas j.
effective gas diameters [31,40], but differs slightly from the order Rearranging Eq. (7) in terms of ln Dj yields
reported for collision, kinetic, and permeability correlation dia-  
1a
meters [15,32,40]. Breck diameters and permeability correlated ln Dj 1 ij ln Di ij b  f 8
RT
diameters report O2 to have a slightly larger gas diameter than
CO2. For this analysis, the diameter of CO2 was found to be This equation can be recast in the form of Eq. (3) to obtain
between that of O2 and N2.   
1a 1
The effective gas diameters calculated in this study represent Dj exp ij b  f Di ij 9
RT
the average molecular cross-section for gases in glassy polymers.
For non-spherical molecules (e.g., H2, O2, N2, CO2, etc.), effective gas where the coefcients from Eq. (3), nd and kd, are written as
diameters would likely be inuenced by the free volume distribu- follows:
tion of their polymer media; as this distribution narrows, the   
1a
smallest cross-section of a non-spherical molecule becomes the nd 1 ij and kd exp nd  1 b  f 10
RT
dominant dimension that relates to gas transport. The extreme case
for this effect would come from molecular sieves such as zeolites Although current theory predicts that nd np, empirical evidence
where the smallest cross-section of the molecule is dominant in suggests that this equality is not completely valid (i.e., nd anp).
transport correlations. Comparisons of predicted and experimental values are listed in
Table 4 and shown in Fig. 8. The predicted values were calculated
from Eq. (10) using parameters from Freeman [13], and experi-
3. Correlations of the parameter kd mental values of nd and np were determined from the database
used in this study. There is scatter in the data, but the correlation
The values of the parameter kd, dened by Eq. (3), were of nd is quite good judging from the R value, and the correlation
determined using the averaged values from Eqs. (4) and (5) that slope is less than one. Using a lower value of f would allow for a
are listed in Table 1. The Freeman theory [13] allows for a prediction correlation slope of unity.
76 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

Table 4
Comparison of the front factor kd: theory versus experiment. For the predicted
values, f was held at a value of 12,600 cal/mol.

Gas pair kd (experimental) kd (predicted) using nd kd (predicted) using np

O2/N2 2.84 1.85 1.65


H2/N2 8.76 9.50 7.11
He/N2 114 34.5 22.9
H2/CH4 138 33.7 16.1
He/H2 3.05 1.91 2.20
He/CH4 7100 226 66.2
N2/CH4 4.93 1.94 1.61
He/O2 12.0 10.3 12.4
H2/O2 1.81 3.79 3.25
O2/CH4 24.8 4.17 2.88
CO2/CH4 6.73 2.39 2.52
CO2/N2 2.52 1.24 1.50
He/CO2 61.2 24.3 18.0
H2/CO2 29.8 12.6 4.46
O2/CO2 2.39 1.52 1.11

Fig. 9. Diffusivity upper bound correlation for He/CH4.

Fig. 8. Comparison between predicted values of kd from nd and np data with


experimental values of kd.

4. Upper bound for diffusivity selectivity


Fig. 10. Diffusivity upper bound correlation for He/N2.

The permeability upper bound data discussed in prior papers


[12,14] used the following equation:
However, for some gas pairs, there is considerable difference
P i kup iju
n
up
11
between these terms. To illustrate this difference, Figs. 9 and 10
An analogous equation can be written for diffusion coefcients show plots in the form of Eq. (13) both in terms of nud and nup. For
the case of nup, the term nud in Eq. (13) is replaced by the term nup.
Di kud niju
ud
12
For these gures, nup is taken from reference [14].
Plotting ij (y-axis) versus Di (x-axis) on a loglog plot yields the
following relationship:
5. Correlations of solubility selectivity
log kud log Di
log iju  13
nud nud
A database of solubility coefcients was compiled from the
The slope of the upper bound tradeoff relationship is dened by same references as the database of diffusion coefcients. As with
either nup or nud. The term nup denes the slope of the upper the database of diffusion coefcients, there are signicantly fewer
bound when plotted as the log of the permselectivity versus the values for He and H2 than for the other gases, but there are
log of the permeability, and the term nud denes the slope of the still sufcient data to permit a reasonable analysis. Solubility
upper bound when plotted as the log of the diffusion selectivity coefcients vary signicantly between different polymeric struc-
versus the log of the diffusion coefcient. Therefore, if solubility is tures. In general, polyimides have higher sorption coefcients than
independent of permeability, nup and nud should be identical. polycarbonates, polysulfones, and poly(aryl ethers) for all gases.
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 77

Table 5
Solubility data and condensability correlation parameters.

Gas Number of data entries Solubility coefcienta(average)  103 Solubility coefcienta(median)  103 T/kb (K) Tbc (K) Tcb (K)

He 125 0.883 0.697 10.22 4.222 5.19


H2 206 2.17 1.60 59.7 20.39 32.98
N2 722 6.63 4.70 71.4 73.36 126.20
O2 693 8.60 6.44 106.7 90.20 154.58
CH4 670 22.1 14.1 148.6 111.67 190.56
CO2 766 101 57.0 195.2 304.12

a
Units are cm3(STP)/(cc cmHg).
b
Data taken from Ref. [44].
c
Data taken from Ref. [45].

Table 6
Average and median coefcients from Eq. (14).

Ma N R

Lennard-Jones (average)  7.30 0.0249 0.994


Lennard-Jones (median)  7.49 0.0232 0.994
Boiling point (average)  6.96 0.0269 0.981
Boiling point (median)  7.21 0.0255 0.990
Critical temperature (average)  6.91 0.0153 0.999
Critical temperature (median)  7.14 0.0142 0.999

a
Units are cm3(STP)/(cc cmHg).

As the polymer type is varied, the sorption coefcients for a given


gas span a range of one to two orders of magnitude.
The sorption coefcients in the database used here can be
correlated with normal boiling point (Tb), critical temperature (Tc),
or the Lennard-Jones parameters (/k) according to Eq. (14) [42,43]
ln Si M NT x 14

where Tx is Tb, Tc, or /k in units of K.


Gas solubility coefcients and associated metrics of condensa-
bility are listed in Table 5, and the average and median values for Fig. 11. Solubility coefcient correlation with Lennard-Jones potential well.
the parameters in Eq. (14) are presented in Table 6. The gas
solubility coefcients and parameters from Tables 5 and 6 repre-
sent the average and median values for each gas in the database.
The correlations for Lennard-Jones potential well depth and
critical temperature are shown in Figs. 11 and 12, respectively.
For a given polymer, the sorption isotherms of the more con-
densable gases show stronger dual-mode curvature in their iso-
therms. Therefore, the solubility coefcients and the parameters in
Eq. (14) are strongly affected by the pressure used in the experi-
ment. To illustrate the effect of pressure on gas solubility, an
analysis is shown for polysulfone and PIM-1 in Appendix A. In
general, secant sorption coefcients decrease with pressure, and
this effect is greater for more condensable gases. For example, the
sorption coefcient for N2 in polysulfone is approximately 0.0035
in the limit of zero pressure but is approximately 0.0028 at 10 atm
in units of cm3(STP)/(cc cmHg). For CO2 in polysulfone, the secant
sorption coefcient is approximately 0.077 in the limit of zero
pressure but is approximately 0.028 at 10 atm again in units of
cm3(STP)/(cc cmHg). While the summary of data shown in
Tables 5 and 6 provide general trends to multiple sets of data
found in the literature, in reality, the parameters in Eq. (14) are
susceptible to variations between polymers that exhibit strong
pressure dependence of gas sorption.
The solubility selectivity can be assessed several different ways
Fig. 12. Solubility coefcient correlation with critical temperature.
in this analysis. For example, Fig. 13 shows Si/Sj plotted versus Sj
for S(He)/S(CO2) versus S(CO2). For the purpose of this paper, a plot
of Si/Sj versus Pi correlates more directly with free volume and is
shown in Fig. 14 for S(He)/S(CO2) versus P(He). Alternatively, increasing S(CO2). Fig. 14 demonstrates that the S(He)/S(CO2) ratio
absolute values of S(He) and S(CO2) are plotted versus P(He) in decreases with increasing P(He). Fig. 15 illustrates that both S(He)
Fig. 15. Fig. 13 shows that the S(He)/S(CO2) ratio decreases with and S(CO2) increase with increasing P(He). As the permeability of
78 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

Fig. 13. He/CO2 solubility selectivity versus CO2 solubility.

Fig. 15. He and CO2 solubility versus helium permeability.

Fig. 14. He/CO2 solubility selectivity versus helium permeability. Fig. 16. H2/N2 solubility selectivity versus hydrogen permeability.

He (as well as the permeability of other gases) increase with increasing on the y-axis. However, the trends noted with the gas pairs shown are
free volume; it is apparent from Fig. 15 that the solubility coefcient is consistent with the other gas pairs regardless of plotting convention.
a function of free volume as noted in various literature references The solubility selectivity, Si/Sj, generally decreases with increas-
[27,3335]. Interestingly, when compared to increasing helium per- ing permeability, Pi, when the diameter of gas j is larger than that
meability (and thus increasing free volume), the solubility of CO2 of gas i. Larger gases generally become more soluble as free
increases faster than that of helium. Data for H2/N2 solubility selectiv- volume increases than is the case for smaller gases. This trend
ity versus H2 permeability (Fig. 16) and H2 and N2 solubility versus H2 suggests that sorption sites become less available for the larger gas
permeability (Fig. 17) show similar trends as those observed for the He molecules as free volume decreases.
and CO2 gas pair. From this data and the other gas pair data, there is a The effect of free volume on gas solubility has implications for
modest effect of free volume (or permeability) on solubility selectivity. the Freeman theory, which describes the fundamental basis for the
There is considerable scatter in the data for the solubility correlations upper bound tradeoff empirically noted by Robeson [1214].
in Figs. 13 through 17, and the least squares t will depend upon the According to this theory, sorption coefcients are assumed to be
choice of which gas is plotted on the x-axis and which gas is plotted invariant with free volume. The tradeoff relationships employed in
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 79

Fig. 17. H2 and N2 solubility as a function of hydrogen permeability.

Fig. 19. Comparison of permeability selectivity, diffusivity selectivity, and solubility


selectivity as a function of the more permeable gas in the H2/N2 gas pair.

1  n
Traditionally, the term Sj =Si Si p
is assumed to be constant
and, therefore, independent of permeability (or free volume) for
different gas pairs. However, our analysis from the solubility
coefcient database shows a slight dependency on permeability.
The absolute values of Si increase slightly (for all gases) with
increasing permeability (free volume), and Si/Sj decreases mod-
estly for gases where the diameter of gas i is less than that of gas j.
Comparisons of D(CO2) and S(CO2) versus P(CO2) are shown in
Fig. 18, and the least squares t to these data are given by the
following:

SCO2 2:40PCO2 0:182 ; R 0:441 18

DCO2 0:416PCO2 0:818 ; R 0:946 19

The bold solid line in Fig. 18 is P(CO2) plotted versus itself as


an equal value line. Because P DS from the solution-diffusion
model, the product of the prefactors and the sum of the exponents
Fig. 18. Comparison of D(CO2) and S(CO2) with changes in P(CO2). in Eqs. (18) and (19) should equal unity, which is indeed the case.
This comparison conrms the expected self-consistency of the
solution-diffusion model and shows the effect of free volume on
this work for permeability and diffusion coefcients are described diffusion and solubility coefcients.
by the following equations: The permeability selectivity, diffusivity selectivity, and solubi-
n
n P j kp P i p lity selectivity for H2/N2 are plotted versus hydrogen permeability
P j kp P i p Dj kd Dni d then 15 in Fig. 19, and the least squares t of the data are noted. When
Dj kd Dni d
comparing correlations of P DS, the tted parameters from the
The Freeman theory denes kp and kd by the following equations: standard least squares regression correspond to the equations
      noted below with excellent agreement:
Sj 1  np 1a 1a
kp Si exp np  1 b  f ; kd exp nd  1 b  f
Si RT RT
Pi y Di y Si y
16 xp P i p xd P i d xs P i s 20
Pj Dj Sj
If nd np, as is generally assumed, then Eqs. (15) and (16) can be
rewritten as follows: From Eq. (2), xp (xd)(xs) (xp 0.00747 and (xd)(xs) 0.00740);
n yp yd ys (yp  0.446 and yd ys  0.446); similar agreement
Pj Sj 1  np P i p Pj Sj Dj Pj Dj S j
Si ; and 17 is observed when O2/N2 selectivity values are plotted versus P(O2)
Dj Si Dni d n
Pi p
n n
S i p Di d
n
Pi p Di Si np as shown in Fig. 20.
80 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

Appendix A

A.1. Effect of pressure and experimental measurement method on


experimentally determined diffusion and solubility coefcients

The solution-diffusion model relates gas permeability, P, to the


sorption, S, and diffusion, D, coefcients [46]
P DS 21
The parameter P can be determined from the steady-state ux
normalized by lm thickness and pressure driving force [47]. The
parameters D and S can be determined by transient permeation or
equilibrium sorption experiments [47]. With transient permeation
experiments, diffusion coefcients are determined from time-lag
measurements, and sorption coefcients are calculated by dividing
permeability coefcients by these time-lag diffusion coefcients.
Alternatively, equilibrium sorption experiments can be used to
directly determine a secant sorption coefcient (S C/p where C is
the concentration of gas in the polymer at pressure p), and the
diffusion coefcient can be calculated by dividing the permeability
coefcient by the equilibrium secant sorption coefcient. The
relationship between permeability, diffusion, and sorption coef-
cients determined from these two methods is given by
Fig. 20. Permeability, diffusivity and solubility selectivities for O2/N2 versus O2
permeability. P DS D S Dsec Ssec 22
where D is the diffusion coefcient calculated from time-lag
experiments, S is the sorption coefcient determined from P and
6. Conclusions D , Ssec is the secant sorption coefcient determined from equili-
brium sorption experiments, and Dsec is the diffusion coefcient
The role of solubility in the tradeoff between permselectivity determined from P and Ssec . The dual sorption-mobility model well
and permeability in polymer lms has been evaluated. The large describes sorption and permeation in glassy polymers; this model
database compiled for this analysis, which was limited to glassy predicts that D a Dsec and S aSsec [36]. To evaluate the differ-
polymers, shows the existence of a modest contribution of ences in these sorption and diffusion coefcients, we consider N2,
solubility selectivity to changes in permeability (and thus free CH4, and CO2 transport in polysulfone and PIM-1.
volume). All gases chosen for this analysis (He, H2, N2, O2, CO2, and The diffusion coefcient, as determined by transient permea-
CH4) show an increase in solubility with increasing permeability or tion experiments, can be written as follows [38,48]:
free volume. For specic classes of polymers, there is support in l
2
the literature for this trend; however, we have more broadly D 23
6
extended the correlation between increasing solubility and
free volume for many classes of polymers here. The solubility where is the time-lag of gas diffusing through a polymer of
selectivity (Si/Sj) for the ij gas pair is also observed to be a function thickness, l. The term can be written in terms of Henry's law
of permeability and free volume. When the j gas is larger in diffusion coefcient, DD , and dual-mode parameters
size than the i gas, the value of Si/Sj generally decreases with l
2

increasing permeability of the i gas (Pi). Thus, the solubility sites in f F; K; bp2  24
6DD
the polymer are less available to the larger gas as the polymer
packing density increases and free volume decreases. An upper where p2 is the gas pressure upstream of the polymer lm, and the
bound relationship also exists for Di/Dj versus Di similar to that function in the brackets has been tabulated by Paul and Koros [36].
noted for permeability. The slope of the diffusivity upper bound The terms K and F are dened as follows:
can be different than that of the permeability upper bound due to C H0 b
a contribution from the effect of free volume on solubility K 25
kD
selectivity.
The solubility database was used to determine the relationship DH
F 26
of the equation employed for relating solubility coefcients with DD
critical temperature, boiling point or Lennard-Jones parameters,
where the terms C H0 , b, and kD are the dual-mode sorption
i.e., ln Si M NT x . Correlations of the average and median S
parameters, and DH and DD are the diffusion coefcients of gas
values were observed with M and N values in the range of those
associated with the Langmuir (H for holes) and Henry's law
typically noted in the literature.
(D for dissolved) modes of sorption.
The effective diffusing gas diameter for CO2 is larger than that
Permeability can be described in terms of dual-mode para-
of O2 and slightly less than that of N2. These results differ from
meters as follows:
zeolite data, which show that CO2 has a smaller kinetic diameter  
than O2 and N2. The permeability correlation, which inherently FK
P kD DD 1 27
assumed the solubility selectivity was invariant with permeability, 1 bp2
yielded CO2 and O2 diameters that were approximately the same. Therefore, we can dene S in terms of Eqs. (23)(27)
This analysis allowed for a calculation of a new set of gas  
diameters with improved prediction, relative to prior determina- P FK
S kD 1 f F; K; bp2  28
tions, for diffusion in glassy polymers. D 1 bp2
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 81

Equilibrium sorption isotherms can be described by the dual- The relationship between Dsec and D for polysulfone and
mode sorption model PIM-1 can be described using the dual-mode parameters for N2,
C H0 b CH4, and CO2 listed in Table 7 [33,49].
Ssec kD 29 To employ Eqs. (31) and (32) in this analysis, the parameter F
1 bp2
was determined by tting permeability data from McHattie et al.
The diffusion coefcient calculated from Eqs. (27) and (29) is [50], and Li et al. [49] to Eq. (27) assuming that DD is constant. The
dened as follows: parameters F and K for polysulfone and PIM-1 are presented in
Table 8. The value of K increases with increasing penetrant
P kD DD 1 FK=1 bp2
Dsec 30 condensability, as is found for other polymers such as polycarbo-
Ssec kD C H0 b=1 bp2 nate [51]. However, there is some scatter in the data for the
Therefore, after some basic algebraic simplication, we can parameter F, which typically increases with decreasing penetrant
describe a ratio of diffusion coefcients and solubility coefcients condensability [51]. The scatter in F may be an indication of the
according to Eqs. (31) and (32); Eq. (31) is the inverse of Eq. (32). difculty in determining the permeability dependence on pressure
for N2, CH4, and CO2 in these polymers.
Dsec f F; K; bp2 kD 1 FK bp2 As shown in Fig. 21, the ratio of Dsec =D from Eq. (31) can be
31
D kD bC H0 kD p2 plotted as a function of pressure for each gas. Permeation data t
for this analysis was taken at pressures up to approximately
Ssec kD bC H0 kD p2 10 atm for PIM-1 and 20 atm for polysulfone. Sorption data were
32
S f F; K; bp2 kD 1 FK bp2 taken at pressures up to approximately 28 atm for PIM-1 and
20 atm for polysulfone. To clearly show the correlation between
Eq. (31) and pressure, we have extrapolated the tted parameters
to permeability and sorption data (i.e., F, K, kD , C H0 , b) to 50 atm. For
Table 7
Dual-mode parameters for diffusion correlation calculations. polymers that are far from their glass transition temperature, the
difference in Dsec and D is more pronounced. For example, Dsec for
   
Polymer Gas kD cm3 STP
C H0 cm3 STP b atm  1 Ref. CO2 is approximately 1.4 times higher than D for polysulfone at
cm3 Polymeratm cm3 Polymer
about 20 atm (cf., Fig. 21A), whereas Dsec for CO2 diffusion is
Polysulfonea N2 0.166 0.957 0.104 [33] approximately 1.7 times higher than D for PIM-1 at the same
CH4 0.257 6.58 0.0901 pressure (cf., Fig. 21B). The Tg of polysulfone is 186 1C [33], and the
CO2 0.728 19.6 0.26
Tg of PIM-1 is undetectable before the onset of polymer degrada-
PIM-1b N2 0.493 31.012 0.076 [49] tion around 350 1C [52].
CH4 0.592 64.966 0.15 Using the data from Fig. 21, diffusion selectivity ratios between
CO2 2.351 106.796 0.421
Dsec and D (i.e., Dsec =D ) can be determined for different gas
a
Data taken at 35 1C. pairs. This comparison is shown for polysulfone and PIM-1 in
b
Data taken at 25 1C. Fig. 22. Because more highly sorbing gases show more pronounced
dual-mode effects at lower pressures, a maximum in Dsec =D
Table 8 results for each gas pair, and the pressure location of this
Parameters F and K for polysulfone and PIM-1. maximum will vary from gas to gas. For polysulfone, the largest
value of Dsec =D is almost 1.3 for CO2/N2 at around 30 atm. For
Polymer Gas F K
PIM-1 and CH4/N2, Dsec =D varies between approximately 0.9 and
Polysulfonea N2 0.14 0.60
1.5 between vacuum and 50 atm. Therefore, for the samples
CH4 0.21 2.31 considered here, diffusion selectivities calculated from equilibrium
CO2 0.12 7.00 sorption experiments could differ by as much as a factor of
PIM-1b N2 0.026 4.8 1.5 when compared to calculating diffusion selectivities from
CH4 0.009 16.5 time-lag techniques.
CO2 0.026 19.1 Analogous to the diffusion analysis in Figs. 21 and 22,
a
comparisons can be made between sorption coefcients calcu-
Data taken at 35 1C.
b
Data taken at 25 1C. lated directly from equilibrium sorption experiments (i.e., Ssec ),

Fig. 21. Comparison of Dsec =D for CO2, CH4, and N2 as a function of pressure for (A) polysulfone and (B) PIM-1.
82 L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183

Fig. 22. Comparison of Dsec =D for CO2/N2, CO2/CH4, and CH4/N2 for (A) polysulfone and (B) PIM-1.

Fig. 23. Comparison of Ssec =S for CO2, CH4, and N2 as a function of pressure for (A) polysulfone and (B) PIM-1.

Fig. 24. Comparison of Ssec =S for CO2/N2, CO2/CH4, and CH4/N2 for (A) polysulfone and (B) PIM-1.

sorption coefcients calculated from permeability coefcients and 24. Similar to the ndings for diffusion coefcients, solubi-
and D (i.e., S ), and the sorption selectivities from each lity coefcients and solubility selectivities will differ if calcu-
measurement (i.e., Ssec =S ). The relationship between Ssec , S , lated from sorption isotherms or if derived from time-lag
and Ssec =S can be determined by taking the inverse of the experiments. Furthermore, solubility will vary depending on
plots in Figs. 21 and 22, and these plots are shown in Figs. 23 experimental pressure.
L.M. Robeson et al. / Journal of Membrane Science 453 (2014) 7183 83

References [26] A. Assogna, G. Perego, A. Roggero, R. Sisto, C. Valentini, Structure and gas
permeability of silylated polyphenylene oxide, J. Membr. Sci. 71 (1992)
97103.
[1] D.F. Sanders, Z.P. Smith, R. Guo, L.M. Robeson, J.E. McGrath, D.R. Paul, [27] S. Miyata, S. Sato, K. Nagai, T. Nakagawa, K. Kudo, Relationship between gas
B.D. Freeman, Energy-efcient polymeric gas separation membranes for a transport properties and fractional free volume determined from dielectric
sustainable future: a review, Polymer 54 (2013) 47294761. constant in polyimide lms containing the hexauoroisopropylidene group, J.
[2] J.M.S. Henis, Commercial and practical aspects of gas separation membranes, Appl. Polym. Sci. 107 (2008) 39333944.
in: D.R. Paul, Y.P. Yampolskii (Eds.), Polymeric Gas Separation Membranes, CRC [28] K.-i. Okamoto, K. Tanaka, H. Kita, M. Ishida, M. Kakimoto, Y. Imai, Gas
Press, Boca Raton, 1994. permeability and permselectivity of polyimides prepared from 4,4-diamino-
[3] P. Bernardo, E. Drioli, G. Golemme, Membrane gas separation: a review/state of triphenylamine, Polym. J. 24 (1992) 451457.
the art, Ind. Eng. Chem. Res. 48 (2009) 46384663. [29] K. Tanaka, M. Okano, H. Toshino, H. Kita, K.-i. Okamoto, Effect of methyl
[4] Y. Yampolskii, Polymeric gas separation membranes, Macromolecules 45 substituents on permeability and permselectivity of gases in polyimides
(2012) 32983311. prepared from methyl-substituted phenylenediamines, J. Polym. Sci. B: Polym.
[5] R.W. Baker, Future directions of membrane gas separation technology, Ind. Phys. 30 (1992) 907914.
Eng. Chem. Res. 41 (2002) 13931411. [30] N. Muruganandam, W.J. Koros, D.R. Paul, Gas sorption and transport in
[6] R. Spillman, Economics of gas separation membrane processes, in: R.D. Noble, substituted polycarbonates, J. Polym. Sci. B: Polym. Phys. 25 (1987)
S.A. Stern (Eds.), Membrane Separation Technology, Principles and Applica- 19992026.
tions, Elsevier Science B.V., Amsterdam, 1995, pp. 589663. [31] M.M. Dal-Cin, A. Kumar, L. Layton, Revisiting the experimental and theoretical
[7] T.-H. Kim, W.J. Koros, G.R. Husk, Advanced gas separation membrane materi- upper bounds of light pure gas selectivitypermeability for polymeric mem-
als: rigid aromatic polyimides, Sep. Sci. Technol. 23 (1988) 16111626. branes, J. Membr. Sci. 323 (2008) 299308.
[8] T.A. Barbari, W.J. Koros, D.R. Paul, Polymeric membranes based on bisphenol-A [32] D.W. Breck, Zeolite Molecular Sieves: Structure, Chemistry, and Use, John
for gas separations, J. Membr. Sci. 42 (1989) 6986. Wiley & Sons, New York, NY (1973) (771 pp).
[9] S.A. Stern, Y. Mi, H. Yamamoto, A.K.S. Clair, Structure/permeability relation- [33] C.L. Aitken, W.J. Koros, D.R. Paul, Effect of structural symmetry on gas
ships of polyimide membranes. Applications to the separation of gas mixtures, transport properties of polysulfones, Macromolecules 25 (1992) 34243434.
J. Polym. Sci. Part B: Polym. Phys. 27 (1989) 18871909. [34] M.R. Pixton, D.R. Paul, Gas transport properties of polyarylates Part II:
[10] V.T. Stannett, W.J. Koros, D.R. Paul, H.K. Lonsdale, R.W. Baker, Recent advances tetrabromination of the bisphenol, J. Polym. Sci. B: Polym. Phys. 33 (1995)
in membrane science and technology, Advances in Polymer Science, Springer, 13531364.
Berlin Heidelberg (1979) 69121. [35] C.T. Wright, D.R. Paul, Gas sorption and transport in poly(tertiary-butyl
[11] N. Muruganandam, D.R. Paul, Evaluation of substituted polycarbonates and a methacrylate), Polymer 38 (1997) 18711878.
blend with polystyrene as gas separation membranes, J. Membr. Sci. 34 (1987) [36] D.R. Paul, W.J. Koros, Effect of partially immobilizing sorption on permeability
185198. and the diffusion time lag, J. Polym. Sci.: Polym. Phys. Ed. 14 (1976) 675685.
[12] L.M. Robeson, Correlation of separation factor versus permeability for poly- [37] K. Okamoto, K. Noborio, J. Hao, K. Tanaka, K. Hidetoshi, Permeation and
meric membranes, J. Membr. Sci. 62 (1991) 165185. separation properties of polyimide membranes to 1, 3-butadiene and n-
[13] B.D. Freeman, Basis of permeability/selectivity tradeoff relations in polymeric butane, J. Membr. Sci. 134 (1997) 171179.
gas separation membranes, Macromolecules 32 (1999) 375380. [38] W.R. Vieth, J.M. Howell, J.H. Hsieh, Dual sorption theory, J. Membr. Sci. 1
[14] L.M. Robeson, The upper bound revisited, J. Membr. Sci. 320 (2008) 390400. (1976) 177220.
[15] L.M. Robeson, B.D. Freeman, D.R. Paul, B.W. Rowe, An empirical correlation of [39] Z.P. Smith, R.R. Tiwari, T.M. Murphy, D.F. Sanders, K.L. Gleason, D.R. Paul, B.
gas permeability and permselectivity in polymers and its theoretical basis, D. Freeman, Hydrogen sorption in polymers for membrane applications,
J. Membr. Sci. 341 (2009) 178185. Polymer 54 (2013) 30263037.
[16] A.Y. Alentiev, Y.P. Yampolskii, Free volume model and tradeoff relations of gas [40] J.-J. Shieh, T.S. Chung, Gas permeability, diffusivity, and solubility of poly(4-
permeability and selectivity in glassy polymers, J. Membr. Sci. 165 (2000) vinylpyridine) lm, J. Polym. Sci. B: Polym. Phys. 37 (1999) 28512861.
201216. [41] P.R. Bevington, K.D. Robinson, Data Reduction and Error Analysis for the
[17] A. Alentiev, Y. Yampolskii, Correlation of gas permeability and diffusivity with Physical Sciences, third edition, McGraw Hill, Boston, 2003.
selectivity: orientations of the clouds of the data points and the effects of [42] G.J. van Amerongen, Diffusion in elastomers, Rubber Chem. Technol. 37 (1964)
temperature, Ind. Eng. Chem. Res. 52 (2013) 88648874. 10651152.
[18] Y. Yampolskii, Correlations with and prediction of activation energies of gas [43] V. Stannett, Simple gases, in: J. Crank, G.S. Park (Eds.), Diffusion in Polymers,
permeation and diffusion in glassy polymers, J. Membr. Sci. 148 (1998) 5969. Academic Press, New York, NY, 1968.
[19] B.W. Rowe, L.M. Robeson, B.D. Freeman, D.R. Paul, Inuence of temperature on [44] B.E. Poling, J.M. Prausnitz, J.P. O'Connell, The Properties of Gases and Liquids,
the upper bound: theoretical considerations and comparison with experi- McGraw Hill Book Co., New York, NY, 2001.
mental results, J. Membr. Sci. 360 (2010) 5869. [45] W.M. Haynes, D.R. Lide, T.J. Bruno, CRC Handbook of Chemistry and Physics
[20] J.Y. Park, D.R. Paul, Correlation and prediction of gas permeability in glassy 20122013, CRC Press, Boca Raton, 2012.
polymer membrane materials via a modied free volume based group [46] J.G. Wijmans, R.W. Baker, The solution-diffusion model: a review, J. Membr.
contribution method, J. Membr. Sci. 125 (1997) 2339. Sci. 107 (1995) 121.
[21] L.M. Robeson, C.D. Smith, M. Langsam, A group contribution approach to [47] H. Lin, B.D. Freeman, in: H. Czichos, T. Saito, L. Smith (Eds.), Springer Hand-
predict permeability and permselectivity of aromatic polymers, J. Membr. Sci. book: Permeation and Diffusion, Springer, New York, 2006, pp. 371387.
132 (1997) 3354. [48] P. Meares, The diffusion of gases through polyvinyl acetate, J. Am. Chem. Soc.
[22] D.V. Laciak, L.M. Robeson, C.D. Smith, Group contribution modeling of gas 76 (1954) 34153422.
transport in polymeric membranes, in: B.D. Freeman, I. Pinnau (Eds.), Polymer [49] P. Li, T.S. Chung, D.R. Paul, Gas sorption and permeation in PIM-1, J. Membr.
Membranes for Gas and Vapor Separation, vol. 733, 1999, American Chemical Sci. 432 (2013) 5057.
Society Symposium Series, Washington, DC, 1999, pp. 151177. [50] J.S. McHattie, W.J. Koros, D.R. Paul, Gas transport properties of polysulphones:
[23] J.M. Mohr, D.R. Paul, G.L. Tullos, P.E. Cassidy, Gas transport properties of a 1. Role of symmetry of methyl group placement on bisphenol rings, Polymer
series of poly(ether ketone) polymers, Polymer 32 (1991) 23872394. 32 (1991) 840850.
[24] F.A. Ruiz-Trevio, D.R. Paul, Gas permselectivity properties of high free volume [51] W.J. Koros, D.R. Paul, A.A. Rocha, Carbon dioxide sorption and transport in
polymers modied by a low molecular weight additive, J. Appl. Polym. Sci. 68 polycarbonate, J. Polym. Sci.: Polym. Phys. Ed. 14 (1976) 687702.
(1998) 403415. [52] N. Du, G.P. Robertson, J. Song, I. Pinnau, S. Thomas, M.D. Guiver, Polymers of
[25] S.L. Liu, R. Wang, T.S. Chung, M.L. Chng, Y. Liu, R.H. Vora, Effect of diamine intrinsic microporosity containing triuoromethyl and phenylsulfone groups
composition on the gas transport properties in 6FDA-durene/3,3-diaminodi- as materials for membrane gas separation, Macromolecules 41 (2008)
phenyl sulfone copolyimides, J. Membr. Sci. 202 (2002) 165176. 96569662.

Anda mungkin juga menyukai