You are on page 1of 72

Accepted Manuscript

Polymer Membranes for Acid Gas Removal from Natural Gas


Gigi George, Nidhika Bhoria, Sama AlHallaq, Ahmed Abdala, Vikas Mittal
PII:
DOI:
Reference:

S1383-5866(15)30402-0
http://dx.doi.org/10.1016/j.seppur.2015.12.033
SEPPUR 12758

To appear in:

Separation and Purification Technology

Received Date:
Revised Date:
Accepted Date:

28 September 2015
19 December 2015
21 December 2015

Please cite this article as: G. George, N. Bhoria, S. AlHallaq, A. Abdala, V. Mittal, Polymer Membranes for Acid
Gas Removal from Natural Gas, Separation and Purification Technology (2015), doi: http://dx.doi.org/10.1016/
j.seppur.2015.12.033

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Polymer Membranes for Acid Gas Removal from Natural Gas


Gigi George, Nidhika Bhoria, Sama AlHallaq, Ahmed Abdala, Vikas Mittal *
The Petroleum Institute, Abu Dhabi, United Arab Emirates.
Abstract
The use of polymeric membrane technology is an exciting approach towards the removal of acid
gases, namely, carbon dioxide and hydrogen
sulfide, from natural gas streams. Polymer
membranes exhibit good mechanical, thermal
and chemical stabilities. These membrane
materials can also possess desirable transport
properties such as high permeability and
selectivity. A number of studies have attempted
to
improve
these
properties
without
compromising the advantages of existing
technologies.
Various
preparations
and
structures of polymer membranes were
reviewed. The structure-property analyses of
these polymer membranes used for acid gas removal followed by their selectivity-permeability
relationship and economic aspects are considered here.
Keywords: Acid gas removal, polymer membrane, hydrogen sulfide, carbon dioxide, natural
gas

1. Introduction
Natural gas has been a popular energy source for many decades, and its demand as a fuel is
continuously increasing worldwide.[1] Many of the gas reserves worldwide contain low-quality
natural gas with high levels of impurities or contaminants. Although some of these gas reserves
were discovered decades ago, they were not developed due to the lack of economically feasible
purification technologies. However, with the increase in natural gas prices and natural gas
demands, many countries have directed their focus towards those low-quality gas reserves. The
development of the new low-quality gas fields requires a more complicated series of processes
to produce sales gas as per the markets specifications. This in turn demands the development
of new technologies that can cope with the increase in impurities while maintaining the
economic feasibility of developing the gas plant. Commercial natural gas, mostly methane, has
the highest heating value per unit mass (approximately 21,520 BTU/lbm = 50.1 MJ/kg, LHV)
when compared to other hydrocarbon fuels such as butane, diesel fuel and gasoline.
Additionally, it has the lowest carbon content per unit mass; thus, upon combustion, methane
releases approximately 30% less carbon dioxide than oil and 43% less carbon dioxide than
coal.[2]
*Author to whom correspondence may be addressed:
Vikas Mittal; E-mail: vmittal@pi.ac.ae

Thus, methane is the cleanest hydrocarbon fuel source. Raw natural gas, however, is
composed of methane with a variety of other components, including higher hydrocarbons,
water, acidic gases (carbon dioxide and hydrogen sulfide), and other impurities such as
mercaptans (R-SH), helium, and nitrogen. For natural gas to reach sales gas specifications, the
typical conditioning and processing chain includes inlet separation, sweetening, mercury
removal, dehydration, natural gas liquids recovery, and, finally, compression for transportation.
In addition, a condensate stabilization section is needed to recover the light gases from liquid
hydrocarbons, and a sulfur recovery plant is needed to recover sulfur from the hydrogen sulfide
(H2S) removed from the raw gas. There is a prediction of considerable growth in the natural gas
sector over the next two decades; natural gas is also occasionally predicted to even overtake
other conventional fossil fuels (e.g., oil) as the main fuel between 2020 and 2030.[3-5] At the
end of 2013, the known worldwide natural gas reserves were reported to be approximately
185.7 trillion cubic meters, a volume that is estimated to be sufficient for 55.1 years of global
demand.[4] Natural gas is therefore expected to have a significant impact on the industrial,
transport, power, residential and commercial sectors.[6, 7]
As H2S is a highly toxic gas[8], there were serious concerns about the safety tests to analyze
the amount present in natural gas. Most of the literature reportedly used gas permeation test
rigs under high levels of operation safety. We would like to highlight the standard American
Society for Testing and Materials (ASTM) test method for the analysis of hydrogen sulfide in
gaseous fuels (lead acetate reaction rate method), which is ASTM D4084-07(2012). This
method is used in the industry to determine the concentration of hydrogen sulfide and to verify
compliance with operational and environmental needs. The sales gas criteria as per the US
pipeline specifications are as follows: (1) minimum gross calorific value of 950 BTU/SCF, (2)
maximum water content of 7.0 lbs/mmcf, (3) maximum H2S content of 4 ppm vol., (4) maximum
total inert gases content of 4% (maximum CO2 content of 2%), and (5) maximum hydrocarbon
dew point of -10C at operating pressure. The calorific value is the main parameter to represent
the sales gas because it quantifies the energy that can be obtained from the gas as a fuel.
Additionally, it determines the price of the produced gas in the market. To maintain the calorific
value higher than the limit, inert gases such as nitrogen and carbon dioxide must be removed to
a maximum of 4 mol%, of which only a maximum of 2 mol% carbon dioxide is allowed. Carbon
dioxide has to be removed not only for the sake of the calorific value but also due to its
corrosiveness in the presence of water, where it forms a weak acid. Similarly, hydrogen sulfide
forms a weak acid in the presence of water and results in a highly corrosive environment.
However, hydrogen sulfide creates a more serious problem; it is toxic to humans at ppm levels,
and it causes instant death at 1000 parts per million (ppm). The water content should also be
decreased to avoid water condensation, hydrate formation and blockage in the pipelines.
Finally, the hydrocarbon dew point must be maintained to avoid selling heavier hydrocarbons
with the gas because they can make higher revenue/profit if recovered and sold as natural gas
liquids.
All of the processes in the natural gas conditioning chain are important to achieve those four
main criteria of pipeline specification. The removal of acid gases, including CO2 and H2S, is
challenging, and the existence of these gases in the natural gas increases the risk associated
with the gas plant and requires the usage of special materials that can withstand the corrosive
environment. Various technologies have been identified for acid gas removal from natural gas;
including using a liquid desiccant to absorb the acidic gases, using a solid desiccant to adsorb
the acidic gases, cryogenic distillation, direct conversion by chemical reactions, and membrane
2

separation. The widely applied sweetening method is amine absorption, where alkaline amine
solution is used to absorb the acidic gases in a high-pressure column. However, the focus of
this review is gas sweetening using membranes.
Gas separation by membranes initially emerged as an industrial application in the 1980s, and
the first major membrane production was for hydrogen separation. A few years later, membrane
technology was introduced for use in nitrogen separation from air and carbon dioxide removal
from natural gas. Since then, many studies have been conducted to utilize membranes for gas
separation in various applications, including acid gas removal from natural gas. As per the
industrial expectations and due to the increase in demand for natural gas, membrane
separation technology is expected to flourish more in the coming decade when low-quality gas
reserves are expected to be developed. This is because membrane technology is an excellent
candidate for removing high concentrations of acid gas, and it competes strongly with other
technologies for bulk acid gas removal. Additionally, with regard to economics, the operating
costs for the current absorption-based methods are directly proportional to the amount of acid
gases in the feed gas.[9] However, for membrane systems, the acid gas concentration in feed
affects only the capital cost of membrane modules, while the operating cost is minimal because
the plant runs almost unmanned.
Until recently, membranes have been limited to the removal of carbon dioxide from natural gas.
Membranes are now becoming competitive for other applications (e.g., separation of nitrogen,
hydrogen sulfide, and natural gas liquids) of natural gas processing.[10] New membrane
materials and configurations can exhibit better efficiency and offer more stability towards the
contaminants found in natural gas. When selecting a membrane material for a specific
separation, a number of factors must be considered, including a favorable combination of the
required permeability and selectivity and the mechanical and chemical properties of the
membrane. Inorganic membranes can be categorized as porous or dense depending on the
structure of the membrane material.[11, 12] In porous inorganic membranes, a thin layer of
porous material is laid on top of a porous metal or ceramic support, which provides mechanical
strength while offering minimum resistance to mass transfer. Carbon, glass, alumina, zeolite,
silicon carbide, titania and zirconia membranes are the main candidates for use as porous
inorganic membranes supported by substrates such as zirconia, zeolite, -alumina, -alumina
and porous stainless steel. There are various advantages and disadvantages of inorganic
membranes compared with polymeric membranes.[13] Inorganic membranes are highly stable
at high temperatures and can be resistant to corrosive and environmentally harsh conditions.
For instance, zeolite membranes are microporous in nature, with crystalline alumina silicate
membrane pores having uniform sizes. They separate gaseous components based on a
molecular sieving mechanism. The molecular sieving principle requires pinhole and crack-free
zeolite membranes. They have relatively lower gas fluxes in comparison to other inorganic
membranes. Thus, thicker zeolite layers are required to obtain pinhole-free and crack-free
zeolite layers. Thermal effects also present a disadvantage. At higher temperatures, the zeolite
layer shrinks while the support which keeps expanding, usually resulting in thermal stress
problems. An ideal zeolite membrane should be thermally stable, solvent resistant and have
perfect shape selectivity. [14]
For the polymeric membranes, main mechanism of gas transport is based on solution-diffusion
mechanism. The model works in the way that the dissolution of membrane material happens
followed by the diffusion. Different amounts of material get dissolved across the polymer
3

membrane and thus the separation occurs through a three-step process such as dissociation,
diffusion and desorption[15, 16].
An important focus in the last decade was on utilizing membranes for carbon dioxide removal
from flue gases to reduce its effect as a greenhouse gas and from natural gas to achieve
pipeline specifications. According to the technical report from UOP Honeywell, the membranes
for CO2 removal from natural gas are a fully established technology in the oil and gas industry.
One of the worlds largest membrane plants for CO2 removal is designed to decrease the CO2
content in natural gas from 45% to 6% for 680 MMSCFD feed in Malaysia, and it began
operation in 2007.[17] In addition, new designs are approaching one BCFD of natural gas as a
feed to membrane plants.[18] This success in CO2 removal at high concentrations shows the
potential of utilizing membranes for both H2S and CO2 separation. Many studies have been
performed to identify the transport properties of CO2 through different polymeric membrane
materials. Those polymeric materials usually had the same tendency to allow permeation of H2S
along with CO2. However, limited research was carried out to identify the transport properties of
H2S due to its toxicity. This forms an obstacle for H2S testing and necessitates special testing
procedures and precautions in laboratories. Despite that and due to the increased interest in
utilizing membranes for H2S gas removal, intensive safety precautions are taken, and research
is being conducted to determine the transport properties of H2S with respect to promising
membrane materials. Indeed, considerable focus has been applied recently to developing new
membrane materials for H2S removal. Not only the membrane transport properties are
considered but also the material processability, chemical and mechanical stability and the
fabrication cost. The key required properties for a membrane to be suitable for gas separation
are as follows:[19, 20] high gas permeation rate for the most permeable (the target of the
separation) gas, high selectivity of the target gas against other compounds in the feed,
manufacturing reproducibility, cost effectiveness and ability to be cheaply manufactured into
different membranes modules, mechanical stability under high operation pressures (aging
resistance), tolerance to contaminants and moderate temperature excursion (thermal and
chemical stability) and plasticization resistance. The first two criteria are the most important
ones that drive the rest; a high permeation flux requires a lower membrane area and leads to a
lower-cost membrane system, and a high selectivity requires a lower driving force to achieve the
desired separation and corresponds to a lower operating cost of the membrane system.[19, 21]
The article by R.W. Baker represents the milestones in the development of membranes for gas
separation, starting from defining Grahams law of diffusion until the installation of membrane
modules in operating plants. This article was limited to the generalized concepts of membrane
structure and fabrication to propose the potential applications of membranes in natural gas
treatment. However, the article left a gap with regards to material design, which we are
addressing in this review, specifically for acid gas removal using polymer materials. [22] Bakers
book provides a detailed overview of the history of membranes for various applications.[23] A
recent review published in 2015 by Jeon et al. focused on polyimide and polysulfone
compounds for the separation of carbon dioxide/methane (CO2/CH4). [24] Our review examines
the simultaneous separation of carbon dioxide, hydrogen disulfide and methane, which is
important from a practical perspective for natural gas purification, using polymer membrane
technology. Upon examining the literature published on membrane gas separation, four different
fields can be identified: development of new membrane materials and the characterization
thereof, membrane area calculation, process simulation and optimization, and membrane
economics. Most of the research has been done on developing new membrane materials; as it
4

is the key element in the success of membrane separation. We have summarized this work in
table 1, where various polymer membranes and membrane materials used for the acid gas
removal from natural gas are tabulated with their respective permeability and selectivity data.
2.

Challenges of membrane technology for natural gas sweetening

Despite of the numerous advantages of current commercial membranes, these systems are
blamed to perform at lower efficiency than amine systems for acid gas removal for a number of
reasons such as the presence of contaminants, concentration polarization
permeability/selectivity trade-off, physical aging and plasticization.[22]
2.1 Contaminants and pre treatment
It is important for all membrane systems to have a proper pre-treatment of the natural gas feed
stream entering to ensure better performance and efficiency. There are a number of
contaminants present in natural gas which can lead to a decline in performance. One of them is
the moisture content which causes the swelling and henceforth the destruction of the whole
membrane integrity. If BETEX and heavy hydrocarbons (C6+) happen to present in membrane
system, will lead to the formation of a film around the membrane surface and thus leading to
drastic drop in permeation rates. There are several corrosion inhibitors and additives used for
various offshore activities that are destructive for the membrane integrity. Particulate material, if
present can block the membrane flow area which depends upon module to module. The spiral
wound module is having fewer blockages than hollow fibre module. But it is expected that the
long term particle flow into any membrane module will block it eventually. While considering the
pre-treatment system for membrane gas separation, contaminants must be removed and must
ensure that liquids will not be formed within the membranes themselves. Condensation can be
prevented by achieving a predetermination of dew point before the membrane and then heating
the gas enough to provide sufficient margin of superheat.[25] Uddin et al reported the influence
on different impurities such as TEG, MEG, HHCs, and H2S on membrane performance [26]and
their recently developed Hybrid FSC membranes shows great potential in natural gas
sweetening. [27]
2.2 Concentration polarisation
Concentration polarization is building up of concentration gradient in the boundary layer of
membrane system. This phenomenon can be seen in all membrane separation processes
because of the selective permeability of membrane. It leads to an inverse relation between the
available driving force and the permeable species leading to adverse effects. This means that
there is decrease in available driving force for species with more permeability and vice versa.
This in turn reduces the overall efficiency of separation and raises the total cost including the
capital and operation cost. The major factor which affects the concentration polarization is the
permeation rate. With more permeation rate, there will be more serious concentration
polarization. Also, for a given permeation rate, the membrane with higher separation factor will
experience more concentration polarization. By increasing the gas feed velocity we can lower its
effect but cannot be eliminated. For a fixed pressure across membrane, the effect is relatively
small on actual operation pressure. But with change in gas feed pressure for a fixed pressure
difference across membrane, there will be change in both the permeate flux as well as
concentration polarization. Generally the gas composition has no effect on the overall
concentration polarization. Thus the accountability of this phenomenon is important while doing
membrane module designing.[28, 29]
5

2.3 Permeability/selectivity trade-off (the upper bound)


The trade-off of permeability and selectivity is the main hurdle faced by polymeric membranes.
An inverse relationship exists between permeability and selectivity, which indicates that the
selectivity of a membrane to different gas pairs can be increased only with a corresponding
decrease in gas permeability.[30-32] This permeability/selectivity trade-off was demonstrated by
Robeson in 1991 and thus proving the major drawback in the commercialisation of polymeric
membranes. It has been observed that there is an upper bound of permeability and selectivity in
polymeric membranes. That study analysed a vast number of amorphous polymers with high
glass transition temperature which were reported over a period of four to five decades of time.
Figure 6 shows the CO2 /CH4 upper bound relationship for glassy and rubbery polymers as
elucidated by Robeson in 1991 and revised in 2008. The study was based on the molecular
diameter of gas molecules and it was concluded that the diffusion coefficient of the constituent
gases governs the membrane separation competencies. The correlation between the polarity of
gases and its relation to permselectivity properties were also discussed in the study. This
inverse relationship has been reported in a number of studies.[33, 34] A graphical
representation of this upper bound has been reported by Robeson, and this representation is
referred to as a benchmark for the development of gas permeation membranes. [16, 35, 36]
Until now, very few membranes have been able to override this benchmark.[37] A very recent
2015 study by Lin aimed at quantitatively interpreting and thereby redefining the upper bound of
the separation performance using polymer membranes and free volume theory. Pure and
mixture of gases were used and it was observed that the CO2 induced plasticization can result
in the reduction in glass transition temperature and increase in free volume in the polymerCO2
systems.[38]
2.4 Physical aging
Physical aging of polymeric membranes reduces gas permeability and alters other polymer
physical properties, including enthalpy, entropy and specific volume.[39-48] Aging slows with
time for two main reasons; firstly when the excess free volume decreases the driving force for
physical aging decreases, and secondly the reduction in free volume decreases the mobility of
the polymer chain, which in turn decreases the segmental motions required to reorganise the
polymer chains. The dependence of physical aging on polymer thickness is widely recognised,
especially when the thickness is very less. The effects of thickness on aging are relevant
concerns for polymeric gas separation membranes because the thicknesses of these
membranes are often on the order of 0.1 mm [49]. The operations of current commercial
membranes are limited to near-ambient temperatures. Additionally, the performance of
commercial membranes degrades over time and corrosive and high temperature environments
are not at all suitable.[50, 51] The decrease in the efficiency of polymeric membranes with time
is dependent on a number of factors, including thermal instability, compaction, fouling and
chemical degradation etc.
2.5 Plasticization
Both CO2 and H2S are strong plasticization agents. A quantitative understanding of plasticization
was studied by Zhang et al. The CO2 induced plasticization of polyimide membranes was
studied where it performs in such a way that the permeability increases with time. It was shown
that the preferential sorption sites are the imide groups. The ether and CF 3 groups are the other
sorption sites. The plasticisation effects at different loadings were also determined and the
effect of ether group in supressing the plasticization was also established. [52] Highly
permeable cPIM-1, Torlon and cPIM-1/Torlon membranes where prepared by Yong where the
improvements that are closer to Robeson upper bound were attributed to hydrogen bonding and
6

charge transfer complexes. These membranes exhibited high plasticization resistance up to 30


atm pressure. [53] On a recent study by Wang, the plasticization characteristics of thermally
reduced (TR) polymers were traced using thin and thick films over hundreds of hours. It was
observed that CO2 permeability of thick films did not show significant decline over time. The
thick TR films showed accelerated ageing when exposed to C2H4, C2H6 and C3H8 and
polymer matrix is rigorously plasticized by the condensable gases.[54]
Plasticization in glassy polymeric membranes is a complex process because of the nonequilibrium state of the polymer used.[55-57] The extent of plasticization depends on a number
of factors, including membrane material, morphology, thickness, temperature, pressure, feed
composition and the types of permeating gases used. Low molecular weight compounds are
often added to glassy polymers to improve the processability of the polymer.[58] These low
molecular weight compounds are referred to as plasticisers, and their addition results in a more
flexible polymer with an increased rate of segmental motion. This mechanism of increased chain
flexibility is termed as plasticization.[59-62] Enhanced inter-segmental polymer mobility can
severely affect separation performance, reduce the mechanical strength of the membrane and
speed up the aging process, resulting in membrane failure.[63, 64] Some studies have also
reported that plasticization helps during fabrication by increasing the fractional free volume of
membrane.[62] Resistance against plasticization in polymeric membranes can be developed in
a number of ways, including crosslinking, heat treatment, blending or reactively forming
interpenetrating networks.[11, 65-67] Plasticization or swelling of the polymeric membrane
matrix by highly plasticizing mixed gases is a big concern for most of the polymeric materials
present so far. Scholes et al. reported an average decrease of 8% in CO2 permeability in
presence of H2S. This was because of the polymer matrix plasticization by H2S which reduces
the gas transport resistance leading to greater flux. [68]
2.6 Membrane compaction at high pressure
The challenge of membrane compaction at high pressure may cause irreversible damage to the
membranes. High feed pressures cause physical damage to the porous support membrane
thereby damaging them. [69, 70] Most of the polymeric membranes are comprised of the porous
structures either as separating layer or as mechanical support to composite membranes. The
material characteristics as well as bulk porosity decides the mechanical stability of the porous
membranes. Structures with micro voids are less affected by compaction than the sponge like
structures. Higher operating pressures reorganize the macro void membrane structure resulting
in increased hydraulic resistance as well as reduced void volume [71]. Bulk layer where most of
the large pores and macro voids i.e. pore volume is present gets severely affected by the
compaction.[72] Surface deposits in a membrane can be easily removed by chemical cleaning
but membrane compaction is irreversible resulting in higher long-term operating costs.[73]
Ebert et al. reported significant differences in compaction behavior using TiO2 with PVDF
confirming that compaction of porous structure under pressure can be reduced by using
organic-inorganic blends or composites to improve the mechanical strength.[70]
2.7 Membrane cost
The cost of membrane depends on a number of factors such as like type of module, its material,
labour, equipment, energy and quality cost [74]. This aspect is discussed in detail at the section
5 of this article; the economics and process optimization of membrane technology.

3. Polymeric membranes
7

Polymer membranes were traditionally developed for packaging applications. In the 1950s;
polymer-based membranes made a large impact on natural gas treatment. Most of the available
literature on acid gas transport properties, especially for H2S gas transport and polymer films,
dates back to the 1950s and 1960s.[75-77] The available literature implies the need for novel
materials that can efficiently remove acid gases and make natural gases that meets the US gas
specifications.[78] The transport properties of acid gases, which are mainly based on their
permeability, selectivity, solubility and diffusion coefficients in various polymers, need to be
explored. Earlier studies reported by Robb in 1968 on acid gas permeability discussed the use
of polydimethylsiloxane (PDMS) at 25C.[79] Another approach in 1989 by Stern studied PDMS
to determine the gas permeability coefficients at 55C. These authors found H2S gas to be
relatively more permeable than CO2, which was attributed to the higher solubility and
condensability of hydrogen sulfide in the polymer. The gas permeability and the ideal selectivity
of alkyl-substituted phenoxy-phosphazene-based membrane materials were studied by Orme to
validate acid gas permeabilities. This study showed that the orderly chain packing of polymer
membranes can be influenced by proper selection of phosphazenes as pendant groups, which
in turn affects membrane performance.[80] Gas permeabilities in the range of 12.5 to 54 barrers
and 16-20 barrers were obtained for CO2 and H2S, respectively. The selective permeability for
CO2/CH4 was in the range of 4 to 12.3 barrers, while that of H2S/CH4 was 4-10 barrers.
Gas separation with polymeric membranes emerged as a commercially viable approach in the
early 1980s. The work by Vaughn also highlights the difficulty in designing materials with high
selectivity towards both CO2 and H2S.[81] Over the last two decades, a significant amount of
research was devoted to gas separation. Membrane researchers had to address various issues
to attain higher stability, better performance and economic feasibility. Higher-order technologies
are believed to contribute to the overall development of membrane technology and its
commercial applications. Thus, the current techniques have led to the development of new
commercially effective innovative materials for CO2 and H2S separation. A large amount of effort
has also been invested into fabricating longer-lasting and defect-free membranes. Lower
membrane flux and higher costs are closely related to the thickness of the membrane geometry,
and these properties have been explored widely for the development of novel materials. Here,
we discuss the various types of polymer membranes that have been used to remove acid gases
from natural gas. Glassy polymers are widely used in studies of polymer membranes for acid
gas removal. Cellulose acetate is the most widely used material for acid gas removal and has
been reported to have a CO2/CH4 selectivity of 19 and a H2S/CH4 selectivity of 24 by Chatterjee
et al.[82] These materials showed relatively low CO2/CH4 selectivities at the operating
conditions that were typical of gas fields due to the presence of heavy hydrocarbons and
swelling-induced plasticization.
Polymer membrane technology and its use for natural gas separation are largely limited to
chemical laboratories or pilot-scale research. In addition to cost concerns, further commercial
use of gas separation membranes appears to be restricted by the membranes low selectivity for
typical gas mixtures. Thus, there is a strong need for the continued development of gas
separation polymer membranes, especially given that the existing technologies, processes and
materials are inadequate to address the available opportunities to their fullest extent. It is
assumed that using highly selective CO2 and H2S polymer membranes instead of either type of
membrane alone can reduce the cost of membrane technology. However, polymer membranes
that combine higher stability, flux and selectivity have not yet been experimentally realized. This
literature review article shows that H2S gas is relatively more permeable than CO2 as a result of
8

H2Ss higher condensability and solubility in the polymers. Nevertheless, a balance between
cost and separation efficiency has yet to be established. This technical review primarily
describes polymer membranes that can address two key challenges: first, being capable of
achieving higher permselectivity for acid gases, and second, being able to handle complex and
aggressive feeds in the field while maintaining their physical and chemical material stability. A
large number of studies have demonstrated polymer membrane routes for natural gas
purification that are feasible, economic and environmentally friendly alternatives to existing
technologies. Research on this aspect is in progress at various academic and industrial
laboratories to improve the existing membranes or to invent novel polymer membranes that are
highly efficient for natural gas purification while simultaneously reducing the overall cost of the
separation process.
3.1 Dense polymeric membranes
Most of the reported transport properties for polymer membranes are derived from dense film
samples. The problem with most membrane materials is that they do not have adequate
resilience to harsh industry conditions, such as high feed pressure/temperature, and as a result,
they will fail. Most of the research in this field was thus focused on making resilient materials.
Major research objectives include better mechanical and thermal stability along with high
resistance to plasticization.[83] In addition to these properties, there is a major demand for
dense polymer membrane materials with high selectivity and permeability towards acid gases.
Polyamides in particular have excellent separation properties and durable mechanical
properties to endure high-pressure natural gas feeds. Hao et al. used the fluorine-containing
polyimide 6FDA-HAB to develop a glassy polymer for CO2- and H2S-selective membranes. The
structure of the polymer is given below the figure 1. These membranes showed a CO2/CH4
selectivity of 60 and a H2S/CH4 selectivity of 15. Kraftschik reported dense copolymer
membranes that are well suited for the instantaneous removal of CO2 and H2S from sour natural
gas streams. The copolyimide used in this study was 6FDA-DAM:DABA (3:2), and gas
permeation tests were performed at the representative aggressive conditions to replicate the
field operations with observed CO2/CH4 selectivities up to 49. Cross-linkable dense films of
polyimide-backboned 6FDA-DAM:DABA as a useful approach to stabilize the properties of the
membranes were later explored.[84] The mechanism of using polyethylene glycol (PEG) as a
cross-linking agent is shown in figure 2. It was observed that both the selectivity and the stability
of membranes were improved by crosslinking. Attractive selectivities of 22 and 27 were
obtained for H2S/CH4 and CO2/CH4, respectively, where triethyleneglycol and
tetraethyleneglycol were used as cross-linking agents.

Figure 1: Structure of fluorine-containing polyimide 6FDA-HAB. Reproduced from reference [85] with permission
from the Journal of Membrane Science.

Plasticization induced by H2S was not observed for feed pressures less than approximately 68
bars, whereas plasticization induced by CO2 occurred at feed pressures greater than
approximately 25 bars. In another study, Scholes et al. demonstrated the performance of PDMS
rubbery membranes. This study determined the interaction parameters of the system using the
FloryHuggins theory for a quaternary system.[68]
9

Figure 2: Cross-linking reaction scheme via thermally activated transesterification. Reproduced from reference [84]
with the permission of Macromolecules.

As reported by Vaughn, a series of dense polymeric membranes consisting of


polyamideimides were synthesized, and these new polymers showed ideal CO2/CH4
selectivities of nearly 50.[81, 86, 87] As listed in table 1, the CO2 and H2S permeabilities of
these polymers were shown to be higher than that of a commercial membrane (Torlon).
However, these membranes showed lower resistance to plasticization for H2S while displaying
enhanced stability towards CO2. The overall H2S/CH4 permselectivity in dense polymer
membranes is governed by solubility, while diffusion selectivity governs CO2/CH4 separations.
The work by Vaughn also highlights the difficulties in designing materials with high selectivity
towards both CO2 and H2S. A polyamideimide, 6F-PAI-1, showed higher CO2/CH4 selectivity
compared to rubbery commercial Pebax membrane materials.[87] It was also observed that
highly fluorinated polyamideimides can offer an economical and efficient platform for polymer
membrane materials that can be used for natural gas purification. Thin film composites
synthesized on ultra-porous polysulfone membrane substrates were studied to determine their
gas transport properties, and the separation of CO2 and H2S from CH4 with these composites
was reported by Sridhar. This study also used molecular dynamics simulations to verify the
theoretical studies and calculate solubility parameters, cohesive energies and sorption values of
CO2, H2S and CH4 gases in polyamide membranes.[88] The experimental study showed that
these membranes can display permeances of 15.2 GPU and 51.6 GPU for CO2 and H2S,
respectively, with selectivities of 14.4 for CO2/CH4 and 49.1 for H2S /CH4 systems.
3.2 Facilitated transport membranes
Because facilitated transport membranes (FTMs) offer both enhanced selectivity to acid gas
and increased flux, these types of membranes have received a great deal of attention.[89] A
schematic of the facilitated transport mechanism is shown in figure 3.

10

Figure 3: Schematic of the facilitated transport mechanism. Reproduced from reference [90] with the permission
from Polymer Physics.

FTMs can be divided into three general categories: fixed carrier or chained carrier membranes,
solvent-swollen polymer membranes and immobilized liquid membranes (ILMs). Fixed carrier
membranes are prepared by attaching reactive functional groups to a polymer film, while ILMs
are typically prepared by saturating the liquid carrier onto an inorganic membrane support
material. Solvent-swollen membranes are prepared by solvent-aided swelling of the polymer
followed by the introduction of a carrier species.[90] Typical flue gas streams may require
preconditioning prior to acid gas purification. To maintain a high permeability and selectivity
towards CO2 and H2S and the capability to retain this ideal characteristic without the need for
preconditioning, a moisture removal step is generally required. Facilitated liquid membranes,
which incorporate a number of liquid absorption stages, are different from any other
conventional membrane system.[91] Better separation can be achieved when desired chemical
reactions occur between the acid gases and carriers.[92] Ilconich demonstrated the production
of an ionic liquid-facilitated membrane powered by amine functionalization followed by
encapsulation on a polymeric support. [93] Promising performance for acid gas removal was
demonstrated for cross-linked nylon66 and polysulfones. This study showed that the
permeabilities of acid gases increased with decreasing feed partial pressures. High selectivities
of 400 for CO2/CH4 and 800 for H2S/CH4 were observed. It was also determined that H2S
transport is facilitated by the formation of the bisulfide ion, HS -. Membrane instability was
observed with H2S testing by Shackley, who later devised a pressure gradient-based system to
study the cross-membrane transport mechanism and stated that this mechanism is driven by
partial pressures across the membrane.[94] Quinn prepared polyelectrolyte fixed carrier
membranes consisting of poly(vinylbenzyltrimethylammoniumfluoride), PVBTAF, supported on a
microporous support.[95] However, the membrane was relatively stable with CO2, suggesting
that more work is needed on this membrane with respect to its mechanical stability. Carbon
nanotube-reinforced polyvinylamine/polyvinyl alcohol was coated on a polysulfone support to
prepare fixed-site-carrier membranes by Xuezhong et al. These membranes were tested with a
permeation rig at 40 bar for a 10% CO2 to give permeance of 0.0840.218 m3 (STP)/(m2 h bar)
with CO2/CH4 selectivity of 17.934.7; no H2S study was performed using this system. [96, 97]

11

3.3 Polymer membrane contactors


Marzouk et al. recently reported the use of microporous polymeric hollow fibers to prepare
membrane contacts to investigate the absorption rate of acid gases and the removal of acid
gases from gas streams, which were pressurized up to 50 bar.[98] Different concentrations of
aqueous sodium hydroxide, amine solutions and distilled water were used as absorption liquids.
Poly(tetrafluoroethylene) and poly(tetrafluoroethylene-co-perfluorinated alkyl vinyl ether) were
employed as hollow fibers. This study demonstrated a significant reduction (nearly 10-fold) in
membrane area and enhanced separation efficiency of the acid gases. The highly permeable
glassy polymer PVTMS (poly-[vinyltrimethylsilane]) was used along with an absorption liquid
(diethanolamine) to make a membrane contactor to study the chemical absorption of acid
gases.[99] Chemical absorption was used for regeneration, and this material showed stable
performance for at least one month. This study demonstrated the potential of gas-liquid
membrane contactors for the efficient regeneration of absorption liquids loaded with acid gases.
The use of alkanolamines in hollow fiber membrane contactors for the simultaneous separation
of H2S and CO2 from natural gas was reported in 2011 by Hedayat; in this study, PVDF and PSf
were used as hollow fibers.[100]

Figure 4: Membrane contactor setup for simultaneous absorption of H2S and CO2 from a gas mixture.
Reproduced from reference [100] with the permission from the Journal of Membrane Science.

The membrane contactors were set up as shown in figure 4. The effects of the membrane
material and absorbent liquid on acid gas removal efficiency and H2S to CO2 selectivity were
studied. It was found that adding MEA enhanced the removal efficiency. Additionally, H2S
removal was largely affected by the presence of CO2 in the feed gas; an increased CO2
concentration in the feed gas negatively influenced the efficiency of separation process.
Conversely, increasing the MEA concentration negatively affected the total mass transfer. This
study also confirmed that increased pressure favors acid gas removal, while changes in
temperature did not significantly affect removal. Studies on gas-liquid membrane contractors will
identify key opportunities and will lead to the invention of novel membranes with higher
adsorbent compatibilities. Absorbance processes are highly affected by higher temperatures
and pressures.[101] Kreulen studied acid gas removal using a liquid membrane filled with pure
methyl-di-ethanol-amine (MDEA) and found no interaction with CO2 under these
conditions.[102] A theoretical and experimental study on the effect of weak acids on the
transport properties of carbon dioxide was performed by Meldon in 1977. They found that the
transport mechanism is a function of the ionic and buffering characteristics of the weak acid;
additionally, these findings were proven to be consistent with theoretical predictions [103].
Another study was performed by Matson et al. in an attempt to improve the selectivity of this
12

system by immobilizing solutions on the pores of membranes or by sandwiching solutions in


between the membranes.[104, 105]
3.4 Mixed matrix membranes
While inorganic membranes demonstrate higher performance, module fabrication is still very
expensive compared to the cost of current polymeric modules. Only a few of the many hundreds
of polymers developed in research laboratories have been commercialized, suggesting that a
large number of polymeric materials are still available to be explored.[78] Even though there is a
significant improvement in the field of polymeric gas separations, commercial polymeric
membranes are still not in a position to overcome the permeability and selectivity trade-off.
Conversely, although inorganic membrane materials offer attractive transport properties, they
remain difficult and expensive to process.[106] The large-scale commercialization of inorganic
membranes is still impractical due to the high fabrication costs of the materials involved. [63]
Therefore, researchers suggested a new approach that can improve the performance of gas
separation membranes.[107] Achieving the desired properties and performance for a particular
membrane material remains a challenge, so scientists have accumulated the knowledge gained
from various systems and used it to fabricate newer membranes. These materials are popularly
known as mixed matrix membranes (MMMs). Most of the advancements in the fabrication of
new membranes address the trade-off between permeability and selectivity.[36] Although these
materials are expensive, their advantages over individual polymer and inorganic membranes
have led to innovative approaches, such as gas separation processes for the purification of
effluent streams and natural gas using MMMs. In one approach, the membranes consist of a
molecular sieving material inserted into a polymer matrix; this system has the potential to
perform economical and high performance gas separations. Mixed matrix membranes have
both the processability of polymeric materials and the superior gas transport properties of
inorganic materials. Some of these membranes have achieved CO2 separation performances
above Robesons upper bound given in figure 6. The main drawback of mixed matrix
membranes is that defects caused by poor contact may occur at the molecular sieve/polymer
interface.[108] Additionally, the high cost of scale-up and brittleness are challenges faced by
these membranes.[109] Inorganic membranes are more expensive than polymeric membranes,
but have many advantages, including temperature and wear resistance, chemical inertness,
high mechanical strength, well-defined stable pore structure and a long lifespan.[110]
Polyimide-silica membranes were analyzed by Suzuki and Yamada to study their CO2
permeability. It was found that permeability is a function of silica content and that this trend can
be used to significantly enhance CO2/CH4 selectivity.[111] Using a sol-gel method, fluorinated
poly(amideimide) and TiO2-based nanocomposite membranes were synthesized by Hu et al.,
who found that lower concentrations of TiO2 led to higher CO2/CH4 selectivity.[112]
Nanocomposite membranes of polyesterurethane and polyetherurethane were studied for their
effect on silica nanoparticles incorporation and the associated permeability behavior with
respect to CO2 and CH4. [113] Matrimid was used as base polymer to prepare dense and
asymmetric membranes with metal organic frameworks (MOFs). Three different MOFs,
[Cu3(BTC)2], ZIF-8 and MIL-53(Al), showed improved CO2/CH4 selectivity.[63] In another study,
asymmetric membranes were prepared using 6FDAODA polyimide as the organic phase and
grated zeolites with 3-aminopropylmethyldiethoxysilane as the inorganic phase. It was observed
that 25 wt.% zeolite in the MMM increased the permeability and selectivity when compared to
neat polyimide membrane.[114] Researchers evaluated the improvement in gas transport
properties using nano-metal oxides incorporated into both rubbery and glassy polymers. The
transport mechanism is largely influenced by the nature of the polymer matrix (i.e., rubbery or
13

glassy). Glassy polymers usually offer higher selectivity, which can be attributed to the
condensability of the gas species.[115] Ahn prepared glassy polysulfone MMMs to study their
mass transport properties and their effect on the filler.[116] Higher silica contents were
demonstrated to favor gas permeability properties, which can be ascribed to the larger free
volume created at the interface between the polymer and silica particles. Mesoporous silica
materials were widely used as inorganic polymer membranes in the synthesis and
functionalization of membrane materials. Synthesis conditions can be varied to obtain silica
materials with various pore geometries and the desired range pore sizes of 230 nm. The
transport mechanism of mesoporous materials is Knudsen diffusion, while solution-diffusion is
dominant in the polymer matrix. It seems that the use of inorganic materials changes the
transport mechanisms. Hydrothermal synthesis techniques or solvent evaporation techniques
were widely used for the production of mesoporous membrane materials, which were usually
grown on porous supports. Functionalization at various levels has been achieved in this field of
research over the last 15 years. The general focus of research has been on fine-tuning the
physiochemical properties of these materials so that they will be well suited for use as
membranes for gas separation. This focus has greatly aided in the evolution of these materials
for gas treatment applications such as natural gas stream purifications and other advanced
applications.[115] In polymer nanocomposite based membranes, the ordered mesoporous thin
films contribute significantly to the separation chemistry for gas mixture purification. Knudsen
diffusion is the type of mechanism that occurs in ordered mesoporous membranes, where the
pore diameters of the diffusing gas molecules are much smaller than the mean free path.
Different gas components have different molecular masses, so the Knudsen diffusion regime
and separation factor are proportional to (MA/MB), where MA and MB are the molecular masses
of components A and B, respectively. Inorganic polymeric membranes are very inexpensive, but
reduced selectivity may result from contact with acid gases. McKeown recently reported an
organic polymer membrane with intrinsic microporosity. This microporosity was achieved during
the fabrication process, which involved contorted rod-like structures. This material provides a
higher gas solubility because it has a high free volume[117].
Various approaches have been reported in the literature to help high temperature application of
polymer membranes, such as zeolite-based polymer membranes, membranes made of silica,
carbon molecular sieve membranes and modified alumina. [118-120] In another study, which
was based on a physical adsorption-based mechanism, zeolite 13X (mesoporous) and zinox
380 (micropores) were used as adsorbents to remove H2S. Zeolite 13X showed a higher level of
adsorption of H2S from natural gas. [121] A high silica CHA-type membrane supported on an alumina support was used to remove acid gases from methane.[122] Thus, zeolites offer
themselves as promising inorganic materials for MMMs. These microporous inorganic
membranes exhibit high permeability and selectivity along with improved thermal, chemical and
mechanical stability. Improving separation properties using polymer nanocomposite membranes
is considered an interesting approach due to the unique properties exhibited by polymeric and
inorganic materials, including chemical and thermal stability, good permeability, selectivity and
mechanical strength.[123]

14

Figure 5: Illustration of nanogap formation in polymer nanocomposite membranes. Reproduced from reference
[124] with the permission from Separation and Purification Technology.

As described by Jiang, porous polymethylmethacrylate can be used to immobilize arginine


complexes to form polymer nanocomposites with high CO2 sorbent capabilities. This study also
discussed complexation of arginine with polystyrene sulfonate, which resulted in enhanced CO2
adsorption capacity.[125] Polymers with intrinsic microporosity are materials with larger and
more accessible surface areas and are of greater significance to adsorption and separation
chemistry. The surface areas of these polymers can range from 300 to 1500 m2/g. Zeolites and
inorganic microporous materials are widely used for adsorption and separations. The gas
permeability capabilities of polymer nanocomposite membranes and the diffusion through these
membranes were explained by Cong [124] with the help of a nanogap formation mechanism, as
illustrated in figure 5. Poor compatibility of the polymer and nanoparticle can lead to nanogaps
surrounded by nanoparticles within the polymer, resulting in increased gas diffusivity and
permeability. This mechanism can also explain the unaltered selectivity associated with
improved permeability in polymer nanocomposites. A similar mechanism for the formation of
nanoporous cavities in the modified polyimides to improve polymer membranes is discussed
here below in the thermally rearranged polymers section.[126]
3.5 Polymers of intrinsic microporosity

Polymers of intrinsic microporosity (PIMs), which are microporous materials with dimensions
less than 2 nm with interconnected pores, have been compared to various inorganic
microporous materials such as activated carbon and zeolites. PIMs as membrane materials
demonstrate high processability and good solubility. High-molecular-weight materials with
solvent processability are decisive for the fabrication of free standing asymmetric membranes,
thin film composites or isotropic films [127]. Till date there is a general lack of H2S information
on these advanced materials. PIMs and its derivatives possess very high CO2 permeabilities,
making them potential candidate for acid gas removal in membrane based systems. They have
already surpassed the Robesons upper bound plots from 1991 and 2008 (revisited version) for
CO2/CH4 and have surface area exceeding 1000 m2/g [118, 128]. A great deal of research is
being carried out regarding the modification of PIMs, such as incorporating molecular units of
different chain lengths to produce high molecular weight copolymers, [129] changing the width
of the distribution of hole sizes and free volume of polymers [130], fuctionalizations. Another
work by Swaidan et al suggests that improving the intermolecular hydrogen-bonding network
can extensively improve the CO2/CH4 selectivity. Cross-linked PIMs membranes were formed at
175C by Du et al and they showed no CO2 plasticization for pure CO2 and CO2/CH4 mixtures
15

when tested at 20 atm pressure [131, 132]. Another approach towards manufacture of thermally
cross linked membrane by treatment at 300 C to obtain CO2 permeability of 4000 barrer and
CO2/CH4 selectivity of 54.8 was done by Li in 2012 [133]. The PIM copolymers and its
structures relationship were explained based on d-spacing and fractional free volume that was
achieved through the pendant group substitution in a work by Du et al in 2008 [134]. Recently,
Scholes et al. reported the effects of water vapor present on PIM-1 for CO2/N2 separation [135].
Seeing the higher permeabilities for CO2, we could assume higher permeabilities for H2S as
well. This concept is in concordance with the idea that CO2 and H2S removal using polymer
membranes may be complimentary to each other, which is discussed in figure 12.
Highly selective PIMs with excellent permeabilities can be a step forward towards economical
membrane based natural gas sweetening systems. Most of the research in the field of PIMs
thus indicate some common themes and is mostly limited to CO2 and CH4 separation. However,
further research in the line of functionalization with appropriate group to develop insights into
the H2S transport properties would be beneficial. Thus, PIMs are of great interest to explore
with regards to acid gas transport properties in membranes systems.
3.6 Fluorinated polymers
Since the discovery of poly(tetrafluoroethylene), fluoropolymers have been considered as strong
candidates for material applications that require greater thermal and chemical stability. The
capability of fluoropolymers to resist common organic solvents adds to their significance.
Conventionally, fluoropolymers were not used as membrane materials due to their low cost
effectiveness, processing difficulty and relatively low permeability due to their semicrystalline
nature.[136-138] Over the last few decades, perfluorinated polymers have been discovered that
are soluble, amorphous in nature and that exhibit good permeability. Teflon AF (DuPont), Hyflon
AD (Solvay Solexis) and Cytop (Asahi Glass) are some examples of such perfluorinated
polymers that can be prepared as membrane materials using a simple technique such as
solution casting. In recently reports, these fluorinated polymers have been shown to have good
transport properties. The solubility and selectivity of different fluorinated (TFE/PMVE49, Teflon
AF 2400, and Cytop) and non-fluorinated polymers (PDMS, PTMSP and PC) have been
compared and studied by accounting for comparisons between rubbery and glassy samples. It
was observed that the measured H2S/CO2 selectivity of fluoropolymers was less than 1, while
the solubility and selectivity of non-fluorinated polymers ranged between 2.9 and 3.3.
Unfavorable interactions between the fluorinated polymer and H2S seemed to deprive H2S of its
usual solubility advantage over CO2. The H2S permeabilities through fluoropolymer membranes
were significantly lower than the expected values based on molecular predictions, while H2S
permeabilities through non-fluorinated polymer membranes were in agreement with the
standard correlations.
Therefore, it may be concluded that membranes fabricated from fluoropolymers require further
study to be proven useful to process designers in NG treatment. It was demonstrated in a
recent study by Vaughn that the more highly fluorinated polymers such as 6F-PAI-1 polymer
can exhibit better plasticization resistance against H2S.[87] A recent study by Saedi investigated
the effect of preparation and operational parameters on PDMS coated PES membranes for
natural gas sweetening.[139] Iovane has experimentally studied the performance of a PEEK
membrane for CO2 and H2S removal from CH4 in a bio gas stream, which can be applied to
natural gas removal.[140] Similarly, a recent agro-biogas based study by Dolejs has reviewed
various thin film composite (TFC) membranes for removing CO2 and H2S simultaneously, which
can also be explored for natural gas [141]. Amjad-Iranagh has studied molecular simulations of
16

PSF and its composites with chitosan, hyaluronic acid, poly(amidoamine) and hydroxyl
poly(amidoamine) membranes for separation effects of acidic gases.[142]
3.7 Thermally rearranged polymers
The material property transformations that occur in polymers upon thermal treatment are very
interesting. Till date the literature has investigated on CO2 onlyand thus there is a general lack
of H2S information on these advanced materials. Thermal treatment has been widely considered
for its impact on membrane materials as well. The permeability and selectivity of polyimide
membranes that had been pyrolyzed were excellent with regards to their acid gas removal
capabilities, as discussed by Yampolskii.[77, 115] Improvements in the material properties of
thermally rearranged polymers were ascribed to the structure-property relationships connected
to the OH and SH side groups in the ortho position of the imide, which led to the formation of
insoluble polybenzoxazoles or polybenzothiazoles. Few mechanisms were reported to explain
the thermal rearrangements and molecular transformations, suggesting that further
investigations on gas-separating membranes are needed.

Figure 6: CO2/CH4 upper bound relationship.[36, 118] Reproduced from reference [36] with the permission from
the Journal of Membrane Science.

The permeability of a large number of polymer membranes that showed improved separation
characteristics were collectively assessed by Park et al. in 2007.[143] Thermally rearranged
polymers with benzoxazole-phenylene main chains exhibited exceptional CO2/CH4 separation
capabilities above the prior and present upper limit of the Robeson plot, as shown in figure 6;
these polymers are considered unique. The performance of these materials calls for a revision
of the upper bound while leaving a gap in the literature for the further exploration of H2S
transport properties. Hodgkin studied the effect of thermal treatments on hydroxyl-containing
polyimide materials. Unique structural and chemical changes occur in the polymer backbone
and at the side groups. It was found that these rearrangements can improve both the
permeability and selectivity of polymeric membranes. The advantage was attributed to the
formation of nanoporous cavities in the modified polymers.[126]

17

3.8 Polyurethanes
Polyurethanes (PUs) are thermoplastic elastomers that have hard and soft domains at the micro
phase level. The hard domain of polyurethanes consists of diisocyanate and chain extenders,
while the soft segment consists of diols, which can derive from either polyesters or polyethers.
The mechanical strength of PU polymers is driven by the hard segment, which forms the
structural framework of the polymer. The soft segment forms flexible chain structures that can
contribute to good gas permeability. The focus of research in this field has always been to find a
good balance between the hard and soft segments because the ideal hard and soft block ratios
can lead to a good separation process. Gas separation membranes formed from urethanes are
useful in separating gases in gas mixtures. Polar gases containing polar and non-polar gas
species were better separated when purified with polymer membranes. Some of the polymer
membranes that showed desirable properties with industrial significance, such as permeability,
permselectivity and durability, are discussed here.
Polyurethane urea belongs to the same family of polymers, which are prepared by careful
synthesis involving diisocynates, polyether diols and diamines. A study by Hua describes the
transport characteristics of poly(urethane-urea)s obtained through synthesis steps, as shown in
figure 7.[144] The properties of the synthesized polymers were compared with rubbery
polyurethane urea polymer networks crosslinked with ethylene glycols and propylene glycols. It
was discovered that propylene glycol based polyurethane urea has the highest permeability, but
the CO2 selectivity was found to decrease gradually. Most of the techniques reported are
patented, which included the use of polyurethane-polyether and polyurea-polyether block
copolymers containing alternating polyether soft segments and either polyurethane or polyurea
hard segments, as reported by Simmon.[145] These membranes revealed exceptionally good
permeation rates with high selectivity. A patented work by Coady reported the use of a dried
cellulose ester membrane with high permeability to H2S and CO2 to separate these gases from
a natural gas stream.[146] The separation approach uses H2S selective rubbery polymer
membranes made from poly ether urethane ureas for natural gas treatment. Chatterjee reports
the possible use for membranes made of poly(ether urethanes) and poly(ether urethane ureas),
with a structure as shown in figure 8, for the purification of low-quality natural gas streams.
Dense membranes with either poly(propylene oxide) or poly(ethylene oxide) were used as the
polyether component.

18

Figure 7: Typical polyurethane urea synthesis steps. Reproduced from reference [144] with the permission of the
Journal of Membrane Science.

The permeability of these membranes was measured at 35C and at pressures ranging from 4
to 13.6 atm. The polyurethane urea membranes showed a favorable and high H2S/CH4
selectivity (greater than 100 at 20C) and a high H2S gas permeability. Different polyurethanebased membranes used in acid gas removal from natural gas are given in table 1, which
compares the performance of these membranes based on the type of hard and soft segments
present in the polyurethane membranes.

Figure 8: Structure of poly(ether urethane urea). Reproduced from reference [82] with the permission of the Journal
of Membrane Science.

According to the comparison given in the table 1, the overall H2S /CH4 selectivity increased
when poly(ethylene oxide) was used as the poly(ether) segment in the polyurethane. This trend
is believed to be due to interactions between polar H2S molecules and the carbonyl groups of
the membrane. This mechanism was also reported by Mohammadi to explain the influence of
electron-donor amine groups.[147] A poly(ester urethane urea) membrane was used for acid
19

gas removal and was subsequently investigated to determine its permeability and selectivity for
varying feed compositions, pressures and temperatures. Permeances of 45 barrers and 95
barrers for CO2 and H2S were measured, respectively, and selectivities of 43 barrers and 16
barrers for H2S/CH4 and CO2/CH4 were measured, respectively. Amani et al. recently explored
molecular simulation as a reliable tool for predicting the gas transport behaviors of polymeric
membranes. [148] Figure 9 shows the higher solubility of acid gases when compared to other
gases that were tested at 298 K and 0-10 bar. The higher solubility and interaction energies
were due to the higher affinity of H2S with the PUU membrane and thereby higher solubility of
the H2S gas. The solubilities were in the order of H2S > CO2 > CH4 > O2 > N2. The higher
solubility of H2S and CO2 was explained by the fact that, unlike CH4, H2S has dipolar and CO2
has quadrupolar properties.

Figure 9: Adsorption isotherm of acid gases in polyurethane urea. Reproduced from reference [148] with the
permission of the Journal of Membrane Science.

Grand canonical Monte Carlo molecular dynamics were used for understanding the effect of the
polyurethane urea membranes role on the gas separation properties. The study showed that
the permeability of polymer membranes is a direct function of the urea linkages present in the
polymer. With a higher the number of urea linkages, the polymer can have increased chain
mobility and higher inter-chain distances. Additionally, a higher fractional free volume and an
increased number of hydrogen bonds were also found to be the deciding factors in increasing
the urea linkages.
3.9 Ionic liquid membranes
Ionic liquids are molten salts at ambient temperature and pressure that have peculiar physical
properties, such as negligible vapor pressure, a wide range of viscosities and high thermal and
chemical stabilities. Pomelli first reported the solubility of hydrogen sulfide in Butyl
methylimidazolium [BMIM]+ cation based ionic liquids. [149] The solubility of gases in ionic
liquids is very significant for a variety of chemical processes, including gas separations, where
solubility is strongly dependent on temperature.[136] Bulky, asymmetric organic cations include
ammonium, pyrrolidinium, pyridinium, imidazolium and phosphonium, while organic/inorganic
anions include bromide and tetrafluoroborate. Although ionic liquids exhibit many desirable
properties, the use of ionic liquids for separations is limited so far due to the high cost and high
energy consumption required to pump the liquids. Overcoming these difficulties remains a
20

challenge in this field of research. One attempt to address these difficulties led to the invention
of supported ionic liquid membranes (SILMs), which are prepared by immobilizing ionic liquids
onto the pores of a porous support. The adsorption and desorption of the permeate gas occurs
simultaneously within the SILM.[93, 137, 138, 150, 151] Ionic liquids that can simultaneously
separate both acid gases are restricted to the best performers in this study. The major drawback
with these systems is that its reduced mechanical strength due to the presence of a liquid
medium inside the polymer matrix may result in the long-term instability of membranes.
The larger free volumes may be due to repulsions between ions and alkyl chains. An interesting
study performed by Quinn suggested that H2S absorption occurs via the deprotonation of
hydrogen sulfide to form bisulfide HS.-[137] A salt hydrate, tetramethylammonium fluoride
tetrahydrate, [(CH3)4N]F4O was prepared with a high capability to reversibly adsorb H2S
quantities up to 0.30 mol H2S per mole of salt at 50C and 100 kPa. The highest H2S/CH4
selectivity shown was 140, while the selectivity for CO2 was 8. The permeabilityselectivity
trade-off was also observed, as the increased H2S content in the gas reaction mixture
suppressed CO2 permeability. Using a BMIM BF4 ionic liquid, Park reports a multi-phase
separation process and the fabrication of a SILM using room temperature ionic liquids (RTILs)
and polymers.[151] Polyvinylidene fluoride polymers and BMIM BF4 ionic liquids were used to
fabricate the SILMs. The permeability coefficients obtained with these SILMs was 30-180
barrers for carbon dioxide and 160-1100 barrers for hydrogen sulfide. The selectivity of
CO2/CH4 was in the range of 2545, whereas the permeability of H2S/CH4 was in the range of
130260. Jalili studied the solubility and diffusion of hydrogen sulfide in [EMIM][EtSO 4] ionic
liquids and found them to be greater than those of carbon dioxide.[152] Huang described
[EMIM][Pro] ionic liquids and reported that solubility correlates with anion alkalinity.[153] Huang
also reported that H2S absorption capacity decreases as follows: [EMIM][Pro]>[EMIM][Ace]>
[EMIM][Lac]. The potentials of SILMs for high pressure natural gas process due to the
mechanical strength as most of literature results were reported at low pressure such as 1
bar.[154] SILMs with high mechanical strength and which are flexible dense films with sufficient
mechanical resistance to support their potentials in high pressure natural gas feed need to be
developed.
A recent study by Jalil identifies [C2mim][OAc] ionic liquids that contain [OAc] anions (ethanoate
or acetate) that can impart exceptionally high CO2 and H2S dissolving power. This study also
reported the effect of CF3 groups in the anion; the presence of CF3 groups led to a
proportional increase in acid gas solubility. A study by Sakhaeinia et al. on [HOemim][Tf 2N]
confirmed the effect of CF3 groups in the anion of the ionic liquid and a direct correlation with
acid gas solubility.[152] The solubility of H2S is greater than that of CO2, and the solubility of
carbon dioxide and hydrogen sulfide improves with the number of trifluoromethyl (CF3) groups
in the anion. The solubility behavior of acid gases follows the order:
[HOemim][Tf2N][HOemim][OTf]> [HOemim][PF6] >[HOemim][BF4]. The influence of anions in
the ionic liquid on the permeability and selectivity of acid gas has been elucidated in studies
reporting ionic liquids with high CO2 solubility. The usage of low-cost protic ionic liquids for H2S
and CO2 absorption by Huang has demonstrated the high selectivity of H2S from CO2.[155]
Mortazavi-Manesh has studied a large number of ionic liquids for their potential in the area of
acid gas removal. Over 400 probable ionic liquids were categorized based on higher selectivity
of absorption of H2S and CO2 over CH4. They concluded that the best 58 (15% of all ionic
liquids studied) for acid gas removal predominantly contain anions BF4, NO3,CH3SO4 and
cations N4111, pmg,tmg.[156] Almost all of the ionic liquid studies based on acid gas removal
21

have reported absorption data and solubility selectivity data. [157] These ionic liquids promise a
good potential for further exploration of making membranes for natural gas sweetening.[158,
159].
An article by Scovazzo studied [emim][Tf2N], [emim][dca], [C3NH2mim][TfO] and
[C3NH2mim][Tf2N] ionic liquids, which displayed high CO2 permeabilities in the range of 500 to
1780 barrers and CO2/CH4 selectivities in the range of 12 to 30.[160] The ionic liquid studied by
Hanioka, [C3NH2mim][CF3SO3], displays a CO2 permeability of 2500 barrers and a CO2/CH4
selectivity of 120.[161] Polymerizable room temperature ionic liquids (RTILs), a source of
polyRTILs, can be synthesized by incorporating polymerizable groups at either anionic or
cationic sites on the ionic liquid structure, followed by polymerization. With these polymerizable
functional groups, RTILs can be converted into polyRTILs, which can be used for gas
separation membrane applications. Bara hypothesized that the permeability coefficients of
polyRTILs can be tuned by varying the n-alkyl substituent and its chain length. Their adsorption
capabilities were observed to be twice that of the liquid counterparts of polyRTILs. This may
confirm the potential of polyRTILs for gas separations.[162] RTILs with polymerizable functional
groups that can be attached to cationic sites include vinyl, acrylate, methyl acrylate, and styrene
groups. PolyRTILs have larger free volumes compared with their molten states, making
polyRTILs promising materials for CO2 sorption and separation.[152, 163-167]
The various types of membranes and membrane materials that are reportedly used for the acid
gas removal from natural gas streams are listed in table 1 with their respective permeability and
selectivity data.

22

(H2S)

H2S
/CH4

CO2/H2
S

Feed
T C

10-14
20-25
21
22.1
6.9

2.13
239

15-20
50
19
19.4
21

1.14
0.32

55.8

6.9

183

22.6

Poly(ether urethane)
PU3

58.8

13

271

62.2

12.2

Poly(ether urethane
urea) PU 2

197

40
35
35

Feed
Pressure
Bar
13.78
10
10

Gas composition
Pure(P) or Mixed (M)
(CH4/CO2/ H2S) %
P
M (65/29/6)
M(70.8/27.9/1.3)

0.30

35

10

M(69.4/18.1/12.5)

58

0.22

35

10

M(70.8/27.9/1.3)

280

54.9

0.22

35

10

M(69.4/18.1/12.5)

6.1

613

19

0.32

35

10

M(70.8/27.9/1.3)

195

5.6

618

18

0.32

35

10

M(69.4/18.1/12.5)

44.7
50.8
22.4
25.4

17
15
22
20

199
223
102
123

74
66
102
95

0.22
0.22
0.22
0.21

35
35
20
20

10
10
10
10

M(70.8/27.9/1.3)
M(69.4/18.1/12.5)
M(70.8/27.9/1.3)
M(69.4/18.1/12.5)

11

55

25

Up to 10

10.6
21.4
3.3
11.34
15.96
9.92
13.07

9.5
10.5
6.9
43
22
27
18

35
35
35
35

26.53
26.66
6.46
10
10
30
30

18000

0.85

23

1.38

M (Balance/4/800 ppm)
M (Balance/4/800 ppm)
M (Balance/4/800 ppm)
M (91.6/5.4/3)
M (93.2/6.2/0.6)
M (91.6/5.4/3)
M (93.2/6.2/0.6)
M (46% CO,
10.5%CO2, 1.5% H2S,
balance hydrogen)
M (36.5 % CO, 11.7 %
CO2, 0.7 % H2S,
balance hydrogen)
M (650ppm/4%CO2/bal
CH4)

P*
(CO2)

CO2/C
H4

Cellulose Acetate
Cellulose Acetate
Cellulose Acetate
Cellulose Acetate
Poly(ether urethane)
PU1

8.9
8.9
2.43
77.5

Polymer

Poly(ether urethane
urea) PU4
Poly(ester urethane
urea)
Matrimid 5218
6 FDA-IPDA
Silicone rubber
Poly(ether urethane
urea)

PTMSP

PDMS

PDMS coated PES


PEEK module

P*

Ref.
[168]
[64]
[169]

[82]

[170]
[171]

[147]

[76]

4400

0.66

21

6.89

3.3

6.9

6.55

10.6

4.18

25

10

M (97.5/2.1/0.4)

[139]

56

7.5

M(53.5/40.2/0.2)

[140]

116.07
-

43.87
-

28.86

[171]

23

membrane
TFC Sterlitech I
TFC Sterlitech II
TFC low P
membrane
TFC high P
membrane
Pebax 4011
Pebax MX 1657

Pebax MX 1074

Pebax MX 1041
Pebax 4033 SA00
Pebax 3533 SA00
Pebax 6333 SA00
Pebax 7233 SA00
Pebax MV 3000 SA
01
6FDA-DAM:DABA
(3:2)
PC

2
2

2.2
2.2

M ( 53.3/46.6/0.1)
M ( 53.3/46.6/0.1)

19

21

3.5

0.9

M ( 53.3/46.6/0.1)

5.6

14.8

6.2

17.5

0.84

M ( 53.3/46.6/0.1)

16

70

26.66

[171]

69.1

0.27

21

13.1

69
89

14.1
-

248
126

50.6
-

0.27
0.71

35
25

10
3

122

0.22

21

190 psig

155
122
39.7
84.4
243
7.4
4.1

11.2
12
11
6.5
5.7
3.9
8.2

695
553
175
312
888
37.8
7.6

50.4
54
49
24
21
20
15

0.22
0.22
0.23
0.27
0.27
0.19
0.54

35
35
35
35
35
35
35

10
10
10
10
10
10
10

M (95.79/4.12/870ppm)
M( 36.5 % CO, 11.7 %
CO2, 0.7 % H2S,
balance hydrogen)
M (70.8/27.9/1.3)
P
M( 36.5 % CO, 11.7 %
CO2, 0.7 % H2S,
balance hydrogen)
M (69.4/18.1/12.5)
M (70.8/27.9/1.3)
M (70.8/27.9/1.3)
M (70.8/27.9/1.3)
M (70.8/27.9/1.3)
M (70.8/27.9/1.3)
M (70.8/27.9/1.3)

100

10

487

49

0.21

35

4.48

[87]

55.6
50.8

32.1
31.1

25.4
23.6

14.7
14.4

0.58
0.61

35
35

49
49

6.5

4.3

21

6.89

M (70/20/10)
M (70/20/10)
M( 36.5 % CO, 11.7 %
CO2, 0.7 % H2S,
balance hydrogen)
M( 36.5 % CO, 11.7 %
CO2, 0.7 % H2S,
balance hydrogen)

PSF

3.8

PSF

9.80.34

PSF/CST-HA(50)
PSF/CSTHA(50)/PAMAM
Nylon - 6

60.30.67

45.60.78

0.088

7.20.
b
76
55.1
b
0.43
44.0
b
0.78
-

3.9

21

13

0.26

[141]

[76]
[82]
[172]
[76]

[82]

[173]

[76]

[142]

[174]

24

Cytop

17

27

21

6.89

TFE/PMVE/8CNVE

28

37

6.89

Teflon AF 1600

680

6.8

23

1.38

Teflon AF 2400

2300

5.6

23

1.38

TFC polyamide

16.6
b
15.2
c
1.18
c
0.95
c
0.95
b
12.3

12.8
14.4
-

52.3
b
51.6
c
9.6
c
8.7
c
8.3
b
11.5

40.5
49.1
1920
1730
2100
-

0.32
0.29
0.12
0.11
0.11
-

30
30
30
25

10
10
4.85
6.22
7.59
3

For CO2 - M( 36.5 %


CO, 11.7 % CO2, 0.7 %
H2S, balance hydrogen)
For H2S M (15%
H2S, 85% N2)
M( 46% CO,
10.5%CO2, 1.5% H2S,
balance hydrogen)
M( 46% CO,
10.5%CO2, 1.5% H2S,
balance hydrogen)
M( 46% CO,
10.5%CO2, 1.5% H2S,
balance hydrogen)
P
P
M (79.7/10/10.3)
M (79.7/10/10.3)
M (79.7/10/10.3)
P

196

25

0.98
8.1

25

5.29

35

4.48

PVBTAF

PEI/PEBA1657
PEI/PDMS/PEBA165
7/PDMS
PDMS
6F-PAI-1

2700
a
32.8

42

2750
6.2

6F-PAI-2

14.2

49

10.3

4.73

35

4.48

6F-PAI-3

21.6

47

5.0

10.9

4.32

35

4.48

0.7 8

52.8

0.2

14.8

103

0.6

0.02

1.9

35

4.55

6.89

4.8
-

4
-

12
2

10
-

0.4
-

30
25

2.1
1.59

12

25

1.59

157

b
a

Torlon

Cardo type
polyimide

PPOP

3.56

M (101-401 ppm
CH4)
P
P
P

[76]

[175]
[95]

[81,
87,
172,
176]

[86]

H2S in

[177]
[80]
[178]

25

PTBP
PDTBP
Pure CA
GCV- modified CA
PBI composite
membrane
[(CH3)4N]F.4H2O) /
celgard membrane
6FDA-HAB
Modified vinylidene
fluoride films

17
27

10
5.4

16
20

9.4
4

1.06
1.35

30
30

2.1
2.1

P
P

8.66

29.5

8.71

29.7

0.99

35

34.5

M (60/20/20)

27.5
129.4
136

19.1
21.8
20

39.7
204
190

27.4
34.3
27.5

0.69
0.63
0.71

35
35
35

48.3
34.5
48.3

250

3.34

140
15

0.13
-

50
35

1.15
10

M (60/20/20)
M (60/20/20)
M (60/20/20)
M (H2: 55%; CO2: 41%;
CO: 1%; CH4: 1%;
: N2 1% and H2S: 1%)
M (89.77/5.1/5.3)
P

5.3

23

M (5% H2S, (95%N2)

[182]

4.6

M (3.91% H2S, 96%


CH4)

[183]

0.005

0.15

109
-

18.7
60

813
-

Seragel membrane
b

3801140 cm
Hg

78.8

N2, O2,CH4,CO2, and


H2S in He

59.37
3.07
1.59
2.12
57.4
56.1
62.9
75.5
58.6
72.6
-

0.11
0.09
0.06
0.06
0.34
0.36
0.34
0.31
0.34
0.30
3.2-4
1.82
3.82

30
30
30
30
25
25
25
25
25
25
-

20
20
20
20
20
20
-

P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P
P

23.2

17.8

1130

6.4
34.6
7.27
3.66
19.6
20.1
20.13
23.8
20.6
22.3
-

1140
4
350
1195
-

3.75

[80]

[179]

[180]
[137]
[181]

P
Mixture of 3% each Ar,

Polymer 3

250

GENERON
PVTMS
SILAR
[MDEAH][Ac]
[MDEAH][For]
[DMEAH][Ac]
[DMEAH][For]
Bmim-BF4
Emim-CH3SO4
HOemim-NO3
N4111-Cl
Pmg-L
Tmg-TCA
[bmim][PF6]
[C2mim][eFAP]
[C8mim][PF6]

226.7
45
1600
2000
-

[184,
185]

[186]
[155]

[156]
[187]
[166]
[188]

26

[P66614][NTf2]
[C4mim][NTf2]
[hmim][Tf2N]
[emim][EtSO4]
[omim][Tf2N]

5
13
-

12.5
10
-

0.47
0.62
0.67

80
80
40
30
30

P
P
P
P
P

2mHEAPr

42

27

80

[C4mim][CH3SO3]

45

40

80

[hemim][Tf2N]

0.43

30

[hemim PF6]

0.31

30

[hemim] [TfO]

0.38

30

[hemim][BF4]

30

[192]

[bmim][Tf2N]
[bmim][BF4]

8.0
-

60
30

P
P

25-45

130260

0.33
0.26

30-180

1601100

[159]
[193]
[151,
164]

PVDF + NaCO3

PPO

7.9

ePTFE fibers

Up to 50
bar

PVDF + BMIM BF4

0.29

M (17.91159 ppm H2S


in CH4)
M (101-401 ppm H2S in
CH4)
M (2% H2S from
pressurized H2SCH4)

[164,
189]
[157]
[152]
[158]
[164,
190]
[164,
191]

[152]

[194]
[177]
[195]

: selectivity
P: permeability /permeance/flux
*units of, P:
with no superscript is Permeability in Barrer
a
: Permeance in GPU;
b
:Permeance in 10-6, cm3/cm2 s cm Hg,
c:
Flux in m3 (STP)/m2hr

Table1: Gas transport properties of various polymer membranes used for acid gas removal from natural gas.

27

4. Material Structure - gas transport performance analysis


Polymeric membranes are great alternatives to existing technologies of acid gas removal from
natural gas because of their ease of formation and scalability to modules of choice, lower cost,
excellent mechanical stability at high pressure, etc. [196] Gas permeation through polymeric
membranes, which are generally nonporous, is solution-diffusion controlled. This diffusion
mechanism depends on the solubility of specific gases and their diffusion through membranes.
Separation therefore depends on the molecular size of the gas and chemical interactions
between the gas and the polymer membrane. Glassy and rubbery are the two main categories
into which polymeric membranes can be divided. This classification is based on the glass
transition temperature relative to ambient conditions. Rubbery membranes have glass transition
temperatures lower than the ambient temperature, while glassy membranes have glass
transition temperatures higher than the ambient temperature. Commercially, all industrial
permselective gas separation membrane processes utilize glassy polymeric membranes
because of their high selectivity and superior mechanical properties. The structure of glassy
membranes is rigid and glass-like because these membranes are operated below their glass
transition temperatures. Conversely, rubbery membranes are flexible and soft because these
membranes are operated above their glass transition temperatures. Generally, rubbery
polymers show high permeability but low selectivity, while glassy polymers behave oppositely
(i.e., low permeability but high selectivity). Examples of rubbery polymers that are being used
extensively in polymeric membranes for gas separation include silicone polymers and poly
(dimethylsiloxane) (PDMS), while examples of glassy polymers used for gas separation include
polyimides, polyamides, polyarylates, polyacetylenes, poly[1-(trimethylsilyl)-1-propyne]
(PTMSP), polycarbonates, cellulose acetate, poly(phenylene oxide), polysulfones and cardotype polymers. Higher selectivity, permeability, competitive flux, long-term stability, low fouling
rate and high mechanical, chemical and thermal stability are some of the main characteristics of
an ideal membrane.[197] Changing the chain packing and chain rigidity can simultaneously
increase both permeability and selectivity.[198, 199] The selection criterion for gas separation
membrane material also depends on weakly and strongly interacting components in natural gas
feeds.[186, 200] However, it is difficult for a single membrane to meet all of these requirements.
The permeation properties of gas separation membranes are significantly affected by thermal
treatment. Resistance against plasticization improves as the thermal treatment temperature is
increased. Thermal rearrangement helps by increasing chain flexibility, which in turn greatly
enhances the fractional free volume and permeability of the membrane while maintaining a high
selectivity.[143]
Table 1 presents an analysis of various types of membrane materials which were reportedly
used for acid gas removal from natural gas streams with respect to their permeability and
selectivity data. The test conditions were not always identical, and thus, it became difficult to
provide a discussion to assess material selection. The feed temperature, which plays a major
role in the performance of the membrane, was generally in the range of 25 to 40C, whereas the
feed pressure was in the range of 2 to 50 bar. The gas compositions used for testing
permeability were either pure or a mixture of acid gases along with CH4 gas. By analyzing the
reported H2S permeability performance given in table 1, we have summarized the findings in
figure 10. A dotted line is added as a trendline in the plot to identify the polymer membranes
that showed permeability above 200 barrers; it is expected to help the reader highlight polymer
membrane materials that showed the highest permeability for H2S. Thus, the materials we
highlight are polyetherurethane urea, polyether block amides (different Pebax grades),
28

supported ionic liquid membrane PVDF- BMIMBF4, SILAR and PBI composite membranes, etc.
We thus generally observe that amine based functional polymers serve as superior candidates
for acid gas removal.

Figure 10: H2S permeability of various polymer membranes

The selectivity of the gas pair H2S/CH4 as a function of H2S permeability is plotted in figure 11.
Here, we introduce the plot to the study the permeability-selectivity relationship for H2S/CH4 as
a function of H2S permeability. Each red triangle in the plot represents a data point that is
representative of a reported value in the literature (table 1) that showed simultaneous
separation properties for both acid gases.

Figure 11: selectivity- permeability relationship for H2S/CH4

The selectivity data interpretation showed us that materials such as polyetherurethane urea,
polyether block amides (different Pebax grades), supported ionic liquid membrane PVDF29

BMIMBF4, modified cellulose acetate and PBI composite membranes, etc. simultaneously
demonstrated higher H2S/CH4 selectivity and good permeability and can be classified as the
useful category of polymer materials for acid gas removal from natural gas. As our literature
survey has extensively reviewed the available published data, an imaginary dotted line has
been embedded in the picture as an H2S/CH4 upper bound for 2015.
This observation is in line with Scholes et al. article which has reviewed about the simultaneous
removal of CO2 and H2S to provide a plot of upper bound. [201] On comparison with Fig. 12, it
becomes evident that the polymer material that has high H2S permeability also had high
H2S/CO2 selectivity. Most of the highly H2S permeable polymer membranes also demonstrated
high H2S/CO2 selectivity. Those polymer materials are consistently either polyetherurethane
urea or different commercial grades of polyether block amides (different Pebax grades). It is
evident that H2S permeability can be considered as a criterion for the selection of polymer
membranes that can treat both the acid gases.

Figure 12: H2S permeability and H2S/CO2 selectivity relationship

We also highlight the necessity of further research on these suggested polymer materials and
on the significance of polymer backbone modification by crosslinking, copolymerization,
blending, etc. Blending and copolymerization were used widely as strategies to modify polymers
to achieve ideal membrane characteristics. In heterogeneous blends, the two main factors that
affect gas transport properties are the morphology of the two phase structure and the nature of
interface between the phases [202, 203]. In homogeneous miscible blends, the gas transport
properties depend on the strength of the interactions between the two components to be
blended.[204, 205] The copolymers can be categorized in terms of block, rigid-flexible block and
random [206] based on their micro-phase separation morphology.[207-209] Another study by
Car described the use of blended thin film membranes made from Pebax and polyethylene
glycol for CO2 separation.[210] The activation energy required for single gas permeation was
calculated using the Arrhenius equation after testing the membrane at various pressures and
temperatures. In a work by Yave, carbon dioxide-selective copolymer membrane materials were
synthesized from polyether block copolymer and polyethylene glycol-dimethylether (PEG30

DME).[211] In a commercial process used by UOP, PEG-DME are used under the trade name
Genosorb in the Selexol process for natural gas treatment and acid gas removal. The use of
this copolymer material increased the carbon dioxide (CO2/H2) permeability and selectivity. In a
recent study, Pebax copolymer SA01 MV 3000 was used for the simultaneous separation of
CO2 and H2S from CH4. When studied under rich and lean feed streams of acid gases, Pebax
showed selectivities of 40 and 10 for H2S/CH4 and CO2/CH4, respectively. [87]
The use of commercial copolymer membrane materials such as the Pebax polymers as
separation membranes for the purification of low-quality natural gas streams has been studied
and reported. These membrane materials are thermoplastic elastomers made of flexible
polyether and rigid polyamides; these materials exhibit H2S /CH4 selectivity and high H2S
permeability, potentially allowing natural gas to be upgraded to meet US pipeline specifications.
This study compared two different types of commercial Pebax membrane materials, namely, the
MX and SA series, both of which are polyether block amides. It was observed that the MX
series of Pebax polymers showed H2S/CH4 selectivities from 49 to 54 and CO2/CH4 selectivities
from 11 to 14 when tested at 35C and 10 atm. Pebax 4011 was reported to have a CO2/CH4
selectivity of 16 and a H2S/CH4 selectivity of 70.[147] This article also reported another
commercial material, Matrimid 5218, which has a CO2/CH4 selectivity of 10.16 and a H2S/CH4
selectivity of 9.5. For comparison, silicone rubber exhibits a CO2/CH4 selectivity of 3.3 and a
H2S/CH4 selectivity of 6.9. In various other studies, nanocomposite membranes were prepared
using Pebax 1657, multiwalled carbon nanotubes and quaternary ammonium compounds; these
membranes showed high CO2 solubility coefficients when tested.[212]
Cross-linking is a polymer structure modification that may help in reducing the chain mobility,
thus resulting in an increase of the glass transition temperature. Cross-linking has been used to
produce membranes that have designable properties with improved efficiencies and increased
resistance to plasticization and aging. There have been many studies on the synthesis of
uncross-linked and cross-linked membrane materials, especially polyimides (because of their
good gas separation and physical properties), to determine the effect of the degree of crosslinking on swelling and plasticization due to CO2. These studies found an increase in CO2/CH4
selectivity with an increasing degree of cross-linking.[39, 66, 213-224]
Modified polymers can be made by introducing bulky substituent groups along the polymer
backbone, which increases the free volume and thereby increases the permeability. However,
introducing bulky substituent groups also reduces selectivity. The substitution of rigid monomer
units or less-mobile linkages helps reduce segmental motion along the chain backbone, thereby
reducing the size of diffusional gaps and increasing selectivity. Additionally, the introduction of
large, polar substituents helps increase the selectivity by increasing inter-segmental
interactions. The simultaneous use of these changes during synthesis helps increase both
permeability and selectivity. There are different classes of nanoporous polymer networks,
including covalent organic frameworks, hyper-crosslinked polymers, conjugated microporous
polymers and polymers of intrinsic microporosity, which are based on synthetic materials of the
lighter elements of the periodic table.[225] The growing variety of synthetic routes to these
materials allows a range of different polymer networks to be formed, including crystalline and
amorphous structures. It is also possible to incorporate many different types of functional
groups in a modular fashion. Mixed matrix membranes (MMMs) were prepared with Pebax and
polyhedral oligomeric silsesquioxane (POSS) and were shown to have an improved CO2
permeability, which may be due to the large cavity of POSS and the cavitys effect on polymeric
chain packing.[226]

31

The upper bound plots given in figure10-12 have helped to identify the useful class of materials
which was suggested based on the upper bound plots for H2S/CO2. However, the target of the
separation is preventing CH4 permeation and thus figure 13 reconsiders the data available from
table 1 to extract some additional conclusions regarding useful classes of materials. It was
observed that membranes with highest H2S permeability are least permeable for CH4. The
materials with lowest CH4 permeability were plotted against materials with high CO2 and H2S
permeability. The result was in line with the depiction of kinetic diameter and condensability of
the main constituents of natural gas as described by Baker[21] in 2008. The useful category of
materials were screened based on the criteria that it has permeability of CH4 < 20 barrers;
permeability of CO2 > 100 Barrers and permeability of H2S > 600. The messages that can be
taken away from this point of view of structure and property relationship is that there are few
reported materials that has very high potential on the simultaneous acid gas removal from naturl
gas. The materials thus identified were polyetherurethane urea, polyether block amides (Pebax
grade 3533), and PBI composite membranes.

Figure 13: Upper bound plots with respect to lowest CH4 permeability

5. Economics and process optimization of membrane technology


Process optimization and economics are the most important aspects to consider when studying
membrane-based gas separation. This is because they are related to the implementation of
membranes on an industrial scale. Cost estimation is crucial not only for determining the
feasibility of membranes but also for determining its potential to replace the conventional acid
gas removal technologies, mainly amine sweetening. To study the economics of membranes,
the first step is to estimate the cost. Analogous to the cost estimation of any chemical process,
the cost of a membrane gas separation process is divided into capital cost and operating cost.
Capital cost includes the cost of the membrane material, the frame (casing, valve and tubing),
pretreatment equipment and compression equipment, a heat exchanger, and a vacuum pump.
Operating and maintenance cost includes product losses, power expenses for operating
compressors and pumps, membrane replacement, and maintenance expenses. Optimization of
the process based on the cost is the second step, where the different parameters affecting the
cost are studied and optimized. Various comprehensive studies have been conducted on the
32

economics of membrane systems and the various factors affecting the processing cost. Most of
these papers focused on membranes for CO2 removal from natural gas stream; however, some
considered both CO2 and H2S removal. Several papers have studied the economics of CO2
removal by membranes and the membranes-amine hybrid system introduced in the 1980s when
membranes first emerged for gas separation.[227] However, each studied the system for a
specific case only, which makes it difficult to compare the different papers. Additionally, the
different papers/researchers used different cost elements and factors to estimate the processing
cost; thus, the quantitative comparison is not applicable. For example, different papers have
studied systems with specific feed flow rates, compositions and pressures; with specific
membrane configuration without considering the other configurations; and with specific retentate
or permeate conditions. Moreover, the membrane system is not yet fully developed for the
economic parameters and the estimated cost values to be definite.
B.D. Bhide and S.A. Stern have eliminated those deficiencies in their published papers and
published two comprehensive papers in 1993.[228] In those papers, the different membrane
configurations were studied and optimized to achieve the lowest processing cost. The effects of
the different process and economical parameters were discussed, including feed composition,
flow rate and pressure, membrane stage cut, pressure ratio, and recycle ratio, as well as
membrane material selectivity and membrane material cost and the cost of methane losses.
Additionally, the presence of H2S and H2O were addressed, but with very low concentrations (1
mol% H2S only), as the cellulose acetate membrane would be plasticized with higher
concentrations. The separation cost of the membrane system was compared to the cost of the
conventional DEA absorption process. In 1998, Bhide published a paper in studying the hybrid
membrane-amine system with the selected optimum membrane configuration. The paper
examined the economics of the hybrid system and studied the effect of the different process
and economical parameters on the total processing cost. The base case for the three published
papers is given in table 2.
Feed composition
Feed flow rate
Feed pressure
Permeate pressure
Membrane thickness
Retentate composition
Membrane material
CH4 permeabilitya
CO2 permeabilitya
C2H6 permeabilitya
N2 permeabilitya
: (cm3(STP).cm)/(s.cm2.cmHg)

5-40 mol% CO2, 1 mol% N2 , 1 mol% C2H6,


Balance CH4
35 MMscfd
55.2 bar
1.4 bar
1000
2 mol% CO2
Cellulose Acetate
0.4257 x 10-10
9 x 10-10
0.18 x 10-10
0.4257 x 10-10

Table 2: Feed composition as summarized by Bhide. Reproduced from reference [169] with permission from the
Journal of Membrane Science.

33

As per the results from those three papers, membrane separation has a lower processing cost
when compared to amine absorption for high compositions of acid gas in the feed. This is
explained by the fact that the operating cost of amine absorption is directly related to the
amount of acid gas in the feed. That is, the higher the acid gas content, the higher the required
amine circulation rate and the higher the energy required for circulating and regenerating it.
Conversely, the diffusion rate increases for membranes at high concentrations of acid gas in the
feed [169]. The total separation cost was represented as a function of the mole fraction of CO2
in the feed, and the product specification for was CO2 2 mol%. Additionally, it can be seen that
the hybrid system has an intermediate cost as per their findings. Later, in 2002 and 2008, J.
Hao, P.A. Rice and S.A. Stern published two similar comprehensive papers studying
membranes for both CO2 and H2S separation from natural gas utilizing two types of polymeric
membrane materials with high CO2/CH4 and H2S/CH4 selectivities.[85] Various process
configurations were studied with and without recycling under a wide range of acid gas
concentration in the feed (0-40 mol% CO2 and 0-40 mol% H2S). However, the amine
sweetening technology was not considered for comparison in a hybrid system. The target
separation was to reach the pipeline specifications of 4 ppm H2S and 2% CO2.The best
configurations for the different acid gas content scenarios were then identified based on the
economical assessment.
The two papers by Hao included area calculation, cost estimation and optimization method. The
papers investigated the co-removal of CO2 and H2S with membrane systems, where two types
of membranes were contained in one module and with modules of one type of membrane
material but in different configurations. The idea of combining two membrane materials in one
module was first introduced by Ohno; Stern and co-workers verified the concept theoretically
and experimentally.[229-231] As per their results, no single configuration was assigned as the
optimum. Rather, they defined an optimum configuration for each different feed composition.

Figure14: Processing costs as a function of both CO2 and H2S concentrations in feed for the optimal process
configurations with and without recycling. (A) Single stage with CO2/CH4-selective membranes. (B) Two stages in
series without recycling, with H2S/CH4-selective membranes in rst stage and CO2/CH4-selective membrane in
second stage. (C) Single stage with H2S/CH4-selective membranes. (D) Two stages in series with recycle
streams. Reproduced from reference [232] with the permission from the Journal of Membrane Science.

34

Figure 14 below shows one of the main figures representing the processing cost per MSCF of
product for the range of CO2 and H2S studied. Ahmed presented the simulation of membrane
systems with a new two-dimensional cross-flow model in Aspen HYSYS for CO2 removal from
natural gas, which was published in 2012.[233] Visual basic was used to simulate the twodimensional cross-flow model as a user defined unit operation in Aspen HYSYS. In their paper,
different process configurations were considered and optimized and the process economics
were studied to determine the optimum configuration. Although similar configurations were
studied by F. Ahmed et.al. and B.D. Bhide, along with similar operating conditions and cost
estimation methods, each study selected a different design as the optimum. This might be
because F. Ahmed et.al. included a feed compression step, which influenced the cost.
Additionally, each study set some parameter before optimizing the configurations, which might
have created some differences. Notably, B.D. Bhide et al. represented the processing cost per
MSCF of feed, while J. Hao et al. represented it per MSCF of product. This again makes it
difficult to directly compare their results quantitatively.The operating and feed conditions and the
membrane properties used in the study are summarized in the table 3 below:
Feed composition
Feed flow rate
Feed pressure
Permeate pressure
Membrane thickness
Membrane material
Methane permeabilitya
CO2 permeabilitya
H2S permeabilitya
Membrane material
Methane permeabilitya
CO2 permeabilitya
H2S permeabilitya
a
:(cm3(STP).cm)/(s.cm2.cmHg)

0-40 mol% CO2, 0-40 mol% H2S, Balance CH4


35 MMscfd
55.2 bar
1.4 bar
1000
Poly(ether urethane urea) , H2S selective
2 x 10-10
32 x 10-10
150 x 10-10
Fluorine-containing polyimide (6FDA-HAB) , CO2
selective
0.1 x 10-10
6 x 10-10
1.5 x 10-10

Table 3: Feed condition and the membrane properties. Reproduced from reference [85] with permission from the
Journal of Membrane Science.

After those comprehensive papers, later publications on membranes studied only one process
configuration. Recently, in 2011, L. Peters studied both amine sweetening and membranes acid
gas removal and estimated the total capital investment of each system.
Each system was studied and optimized separately to achieve 2% CO2 in sweet gas and above
90% CO2 in acid gas.[234] As would be expected, the membranes total capital investment was
lower than that of the amines. In this paper, three different feed cases were studied; one of the
cases flow rate was 373 MMSCFD. This was the highest studied flow rate, as all of the previous
work was on a 35 MMSCFD feed flow. Ahmed, in 2013, proposed a model incorporated into
Hysys that takes into account the temperature and pressure dependence of membrane
permeances.[235] The model was validated by experimental results, and it was concluded that
neglecting the temperature and pressure effect would lead to lower processing cost estimations.
However, only one configuration was studied. Using the HYSYS simulation integrated with
35

ChemBrane, an economic feasibility analysis was performed by Xuezhong et al. for evaluating
the techno-economic feasibility. The system was studied at lower CO2 concentrations, such as
10%.The results showed that FSC membranes can be used at the cost of 5.73E3 $/Nm3 of
sweet NG produced. [96] In a very recent report by Lock in 2015, a high CO2 content was
reported, considering the fact that most natural gas fields have a high CO2 concentration. They
studied a double-staged membrane system with a recycling system, which can alter the
recycling ratio. By increasing the recycling ratio, the membrane area and compressor power
were increased. Thus, increasing the recycling ratio led to a substantial impact of low selectivity
and high CO2 feed concentration.[236]

6. Conclusions
Although many polymeric membranes and their potential applications for gas separation have
been reported, only a limited number of these membranes have been explored for the
simultaneous removal of both CO2 and H2S from natural gas streams. The permeability,
selectivity and mechanical stability of polymer membranes are the decisive factors in the
separation of acid gases from natural gas streams. Among the polymer membranes considered
in this review, the membranes that exhibited higher potential to treat both CO2 and H2S are
mainly polyetherurethane urea, polyether block amides, supported ionic liquid membranes,
modified cellulose acetate and PBI composite membranes. The functional polymers with amide
functional groups are better candidates for acid gas removal. Polymer backbone modification by
crosslinking, copolymerization, and blending are useful strategies to develop durable materials.
These polymer membrane materials have the potential to set the benchmark for further
improvement. Continuous progress will be made possible by improving the polymer membrane
fabrication processes and modifying the physical and chemical structures. Concerns over the
high cost of membranes can be reduced by creating polymer membranes with higher
permeability and selectivity for acid gases relative to CH4. Current research and development
efforts should focus on resolving the key challenges that have been identified and making new
polymer membranes which can have high gas permeability while at the same time retaining
membrane selectivity along with physical and chemical stability. Moreover, various types of
polymer membranes discussed in this review can be used on their own or in combination with
other techniques to achieve natural gas separation with desired selectivities. Therefore, we
recommend that the natural gas industry around the globe, particularly in the Middle Eastern
region, consider applying this technology without delay.
Acknowledgments
The authors thank ADNOC Gas Subcommittee for funding the project development of twostage membrane/adsorption acid gas removal process.
Notes
*Author to whom correspondence may be addressed:
Vikas Mittal; E-mail: vmittal@pi.ac.ae
Chemical Engineering Department,
The Petroleum Institute,
Abu Dhabi, United Arab Emirates.

36

References
[1] S.V. Berg, Lessons in electricity market reform: Regulatory processes and performance, The Electricity Journal,
11 (1998) 13-20.
[2] L. Langston, S.L. Lee, A Bright Natural Gas Future, Mechanical engineering (New York, N.Y. 1919), 132 (2010)
49.
[3] B. Petroleum, BP Statistical Review of World Energy June 2013, Statistical Review of World Energy, (2013).
[4] B. Petroleum, British Petroleum Energy Outlook 2035, British Petroleum Energy Outlook 2035, (2014).
[5] M.J. Economides, D.A. Wood, The state of natural gas, Journal of Natural Gas Science and Engineering, 1
(2009) 1-13.
[6] A.F. Correlj, Markets for Natural Gas, in: Reference Module in Earth Systems and Environmental Sciences,
Elsevier, 2013.
[7] W. Liss, Demand outlook: A golden age of natural gas, Chemical Engineering Progress, 108 (2012) 35-40.
[8] R.J. Reiffenstein, W.C. Hulbert, S.H. Roth, Toxicology of Hydrogen Sulfide, Annual Review of Pharmacology
and Toxicology, 32 (1992) 109-134.
[9] P. Gonzalez, Energy Use, Human, in: S.A. Levin (Ed.) Encyclopedia of Biodiversity, Elsevier, New York, 2007,
pp. 1-20.
[10] D. Dortmundt, K. Doshi, Recent developments in CO 2 removal membrane technology, UOP LLC: Des Plaines,
IL, (1999).
[11] D. Bessarabov, Membrane gas-separation technology in the petrochemical industry, Membrane Technology,
1999 (1999) 9-13.
[12] R. De Lange, K. Keizer, A. Burggraaf, Analysis and theory of gas transport in microporous sol-gel derived
ceramic membranes, Journal of Membrane Science, 104 (1995) 81-100.
[13] J. Caro, M. Noack, P. Klsch, R. Schfer, Zeolite membranes state of their development and perspective,
Microporous and Mesoporous Materials, 38 (2000) 3-24.
[14] M. Yu, R.D. Noble, J.L. Falconer, Zeolite membranes: microstructure characterization and permeation
mechanisms, Accounts of chemical research, 44 (2011) 1196-1206.
[15] G. Maier, Gas separation by polymer membranes: beyond the border, Angewandte Chemie International
Edition, 52 (2013) 4982-4984.
[16] J. Wijmans, R. Baker, The solution-diffusion model: a review, Journal of Membrane Science, 107 (1995) 1-21.
[17] T. Cnop, d. Dortmundt, m. Schott, Continued development of gas separation for highly sour service, UOP
Honeywell technical paper, (2007).
[18] R. Thiruvenkatachari, S. Su, H. An, X.X. Yu, Post combustion CO2 capture by carbon fibre monolithic
adsorbents, Progress in Energy and Combustion Science, 35 (2009) 438-455.
[19] D.R. Paul, Y.P. Yampol'skii, Polymeric Gas Separation Membranes, CRC Press, 1994.
[20] C.E. Powell, G.G. Qiao, Polymeric CO2/N2 gas separation membranes for the capture of carbon dioxide from
power plant flue gases, Journal of Membrane Science, 279 (2006) 1-49.
[21] R.W. Baker, K. Lokhandwala, Natural gas processing with membranes: an overview, Industrial & Engineering
Chemistry Research, 47 (2008) 2109-2121.
[22] R.W. Baker, Future directions of membrane gas separation technology, Industrial & Engineering Chemistry
Research, 41 (2002) 1393-1411.
[23] R. Baker, Membrane technology and applications, John Wiley & Sons, 2012.
[24] Y.-W. Jeon, D.-H. Lee, Gas Membranes for CO2/CH4 (Biogas) Separation: A Review, Environmental
Engineering Science, (2015).
[25] M. Stewart, K. Arnold, Gas Sweetening and Processing Field Manual, Gulf Professional Pub., 2011.
[26] M. Washim Uddin, M.-B. Hgg, Natural gas sweeteningthe effect on CO2CH4 separation after exposing a
facilitated transport membrane to hydrogen sulfide and higher hydrocarbons, Journal of Membrane Science, 423
424 (2012) 143-149.
[27] M. Washim Uddin, M.-B. Hgg, Effect of monoethylene glycol and triethylene glycol contamination on
CO2/CH4 separation of a facilitated transport membrane for natural gas sweetening, Journal of Membrane
Science, 423424 (2012) 150-158.
[28] S. Bhattacharya, S.-T. Hwang, Concentration polarization, separation factor, and Peclet number in membrane
processes, Journal of Membrane Science, 132 (1997) 73-90.

37

[29] G. He, Y. Mi, P. Lock Yue, G. Chen, Theoretical study on concentration polarization in gas separation
membrane processes, Journal of Membrane Science, 153 (1999) 243-258.
[30] T. Barbari, W. Koros, D. Paul, Polymeric membranes based on bisphenol-A for gas separations, Journal of
Membrane Science, 42 (1989) 69-86.
[31] T. Kim, W. Koros, G. Husk, K. O'brien, Relationship between gas separation properties and chemical structure
in a series of aromatic polyimides, Journal of Membrane Science, 37 (1988) 45-62.
[32] C.-L. Lee, H.L. Chapman, M.E. Cifuentes, K.M. Lee, L.D. Merrill, K.L. Ulman, K. Venkataraman, Effects of
polymer structure on the gas permeability of silicone membranes, Journal of Membrane Science, 38 (1988) 55-70.
[33] P. Bahukudumbi, D.M. Ford, Molecular Modeling Study of the PermeabilitySelectivity Trade-off in Polymeric
and Microporous Membranes, Industrial & Engineering Chemistry Research, 45 (2006) 5640-5648.
[34] L. Wang, Y. Cao, M. Zhou, X. Qiu, Q. Yuan, Synthesis, characterization, and gas permeation properties of
6FDA-2,6-DAT/mPDA copolyimides, Front. Chem. China, 4 (2009) 215-221.
[35] L.M. Robeson, Correlation of separation factor versus permeability for polymeric membranes, Journal of
Membrane Science, 62 (1991) 165-185.
[36] L.M. Robeson, The upper bound revisited, Journal of Membrane Science, 320 (2008) 390-400.
[37] A. Kargari, T. Kaghazchi, M. Soleimani, Extraction of gold (III) ions from aqueous solutions using polyamine
type surfactant liquid membrane, The Canadian Journal of Chemical Engineering, 82 (2004) 1301-1306.
[38] H. Lin, M. Yavari, Upper bound of polymeric membranes for mixed-gas CO2/CH4 separations, Journal of
Membrane Science, 475 (2015) 101-109.
[39] L. Cui, W. Qiu, D.R. Paul, W.J. Koros, Physical aging of 6FDA-based polyimide membranes monitored by gas
permeability, Polymer, 52 (2011) 3374-3380.
[40] Y. Huang, D.R. Paul, Physical aging of thin glassy polymer films monitored by optical properties,
Macromolecules, 39 (2006) 1554-1559.
[41] J.M. Hutchinson, Physical aging of polymers, Progress in Polymer Science, 20 (1995) 703-760.
[42] J. Kim, W.J. Koros, D.R. Paul, Physical aging of thin 6FDA-based polyimide membranes containing carboxyl
acid groups. Part I. Transport properties, Polymer, 47 (2006) 3094-3103.
[43] T.M. Murphy, D. Langhe, M. Ponting, E. Baer, B. Freeman, D. Paul, Physical aging of layered glassy polymer
films via gas permeability tracking, Polymer, 52 (2011) 6117-6125.
[44] T.M. Murphy, G.T. Offord, D.R. Paul, Fundamentals of membrane gas separation, Membrane operations:
innovative separations and transformations, (2009) 63-82.
[45] B.W. Rowe, B.D. Freeman, D.R. Paul, Physical aging of ultrathin glassy polymer films tracked by gas
permeability, Polymer, 50 (2009) 5565-5575.
[46] B.W. Rowe, B.D. Freeman, D.R. Paul, Influence of previous history on physical aging in thin glassy polymer
films as gas separation membranes, Polymer, 51 (2010) 3784-3792.
[47] B.W. Rowe, S.J. Pas, A.J. Hill, R. Suzuki, B.D. Freeman, D. Paul, A variable energy positron annihilation
lifetime spectroscopy study of physical aging in thin glassy polymer films, Polymer, 50 (2009) 6149-6156.
[48] L.C.E. Struik, Physical aging in amorphous polymers and other materials, Elsevier Amsterdam, 1978.
[49] Y. Huang, D.R. Paul, Effect of film thickness on the gas-permeation characteristics of glassy polymer
membranes, Industrial & Engineering Chemistry Research, 46 (2007) 2342-2347.
[50] L. David, A. Ismail, Influence of the thermastabilization process and soak time during pyrolysis process on the
polyacrylonitrile carbon membranes for O2//N2 separation, Journal of Membrane Science, 213 (2003) 285-291.
[51] T.C. Merkel, R.P. Gupta, B.S. Turk, B.D. Freeman, Mixed-gas permeation of syngas components in
poly(dimethylsiloxane) and poly(1-trimethylsilyl-1-propyne) at elevated temperatures, Journal of Membrane
Science, 191 (2001) 85-94.
[52] L. Zhang, Y. Xiao, T.-S. Chung, J. Jiang, Mechanistic understanding of CO 2-induced plasticization of a
polyimide membrane: A combination of experiment and simulation study, Polymer, 51 (2010) 4439-4447.
[53] W.F. Yong, F.Y. Li, T.S. Chung, Y.W. Tong, Molecular interaction, gas transport properties and plasticization
behavior of cPIM-1/Torlon blend membranes, Journal of Membrane Science, 462 (2014) 119-130.
[54] H. Wang, T.-S. Chung, D.R. Paul, Physical aging and plasticization of thick and thin films of the thermally
rearranged ortho-functional polyimide 6FDAHAB, Journal of Membrane Science, 458 (2014) 27-35.
[55] T. Visser, N. Masetto, M. Wessling, Materials dependence of mixed gas plasticization behavior in asymmetric
membranes, Journal of Membrane Science, 306 (2007) 16-28.
[56] M. Wessling, M. Lidon Lopez, H. Strathmann, Accelerated plasticization of thin-film composite membranes
used in gas separation, Separation and Purification Technology, 24 (2001) 223-233.

38

[57] X. Duthie, S. Kentish, C. Powell, K. Nagai, G. Qiao, G. Stevens, Operating temperature effects on the
plasticization of polyimide gas separation membranes, Journal of Membrane Science, 294 (2007) 40-49.
[58] B.L. Wadey, Plasticizers, Encyclopedia Of Polymer Science and Technology, (2001).
[59] Y.M. Lee, K.Y. Kim, C. Jung, H.B. Park, H.J. Kwon, S.H. Han, Gas separation using membranes comprising
polybenzoxazoles prepared by thermal rearrangement, in, 2012.
[60] M. Roussenova, M. Murith, A. Alam, J. Ubbink, Plasticization, Antiplasticization, and Molecular Packing in
Amorphous Carbohydrate-Glycerol Matrices, Biomacromolecules, 11 (2010) 3237-3247.
[61] T. Visser, M. Wessling, When do sorption-induced relaxations in glassy polymers set in?, Macromolecules, 40
(2007) 4992-5000.
[62] Y. Xiao, B.T. Low, S.S. Hosseini, T.S. Chung, D.R. Paul, The strategies of molecular architecture and
modification of polyimide-based membranes for CO2 removal from natural gasA review, Progress in Polymer
Science, 34 (2009) 561-580.
[63] S. Basu, A. Cano-Odena, I.F.J. Vankelecom, MOF-containing mixed-matrix membranes for CO2/CH4 and
CO2/N2 binary gas mixture separations, Separation and Purification Technology, 81 (2011) 31-40.
[64] Spillman, R. W., Economics of gas separation membranes, American Institute of Chemical Engineers, New
York, NY, ETATS-UNIS, 1989.
[65] A. Bos, I. Pnt, M. Wessling, H. Strathmann, Plasticization-resistant glassy polyimide membranes for CO2/CO4
separations, Separation and Purification Technology, 14 (1998) 27-39.
[66] J. Fang, H. Kita, K.-i. Okamoto, Gas permeation properties of hyperbranched polyimide membranes, Journal of
Membrane Science, 182 (2001) 245-256.
[67] H. Kita, T. Inada, K. Tanaka, K.-i. Okamoto, Effect of photocrosslinking on permeability and permselectivity of
gases through benzophenone-containing polyimide, Journal of Membrane Science, 87 (1994) 139-147.
[68] C.A. Scholes, G.W. Stevens, S.E. Kentish, The effect of hydrogen sulfide, carbon monoxide and water on the
performance of a PDMS membrane in carbon dioxide/nitrogen separation, Journal of Membrane Science, 350
(2010) 189-199.
[69] N. Ochoa, M. Masuelli, J. Marchese, Effect of hydrophilicity on fouling of an emulsified oil wastewater with
PVDF/PMMA membranes, Journal of Membrane Science, 226 (2003) 203-211.
[70] K. Ebert, D. Fritsch, J. Koll, C. Tjahjawiguna, Influence of inorganic fillers on the compaction behaviour of
porous polymer based membranes, Journal of Membrane Science, 233 (2004) 71-78.
[71] K.M. Persson, V. Gekas, G. Trgrdh, Study of membrane compaction and its influence on ultrafiltration water
permeability, Journal of Membrane Science, 100 (1995) 155-162.
[72] G. Jonsson, Methods for determining the selectivity of reverse osmosis membranes, Desalination, 24 (1977)
19-37.
[73] M.T.M. Pendergast, J.M. Nygaard, A.K. Ghosh, E.M. Hoek, Using nanocomposite materials technology to
understand and control reverse osmosis membrane compaction, Desalination, 261 (2010) 255-263.
[74] S. Kumar, A. Groth, L. Vlacic, An analytical index for evaluating manufacturing cost and performance of lowpressure hollow fibre membrane systems, Desalination, 332 (2014) 44-51.
[75] W.J. Koros, R. Mahajan, Pushing the limits on possibilities for large scale gas separation: which strategies?,
Journal of Membrane Science, 175 (2000) 181-196.
[76] T. Merkel, L. Toy, Comparison of hydrogen sulfide transport properties in fluorinated and nonfluorinated
polymers, Macromolecules, 39 (2006) 7591-7600.
[77] Y. Yampolskii, Polymeric Gas Separation Membranes, Macromolecules, 45 (2012) 3298-3311.
[78] P. Bernardo, E. Drioli, G. Golemme, Membrane gas separation: A review/state of the art, Industrial &
Engineering Chemistry Research, 48 (2009) 4638-4663.
[79] W. Robb, Thin silicone membranestheir permeation properties and some applications, Annals of the New York
Academy of Sciences, 146 (1968) 119-137.
[80] C. Orme, J. Klaehn, F. Stewart, Gas permeability and ideal selectivity of poly[bis-(phenoxy)phosphazene],
poly[bis-(4-tert-butylphenoxy)phosphazene], and poly[bis-(3,5-di-tert-butylphenoxy)1.2(chloro)0.8phosphazene],
Journal of Membrane Science, 238 (2004) 47-55.
[81] J. Vaughn, W. Koros, Effect of the Amide Bond Diamine Structure on the CO2, H2S, and CH4 Transport
Properties of a Series of Novel 6FDA-Based PolyamideImides for Natural Gas Purification, Macromolecules, 45
(2012) 7036-7049.
[82] G. Chatterjee, A. Houde, S. Stern, Poly (ether urethane) and poly (ether urethane urea) membranes with high
H2S/CH4 selectivity, Journal of Membrane Science, 135 (1997) 99-106.
[83] Y. Zhang, J. Sunarso, S. Liu, R. Wang, Current status and development of membranes for CO2/CH4
separation: A review, International Journal of Greenhouse Gas Control, 12 (2013) 84-107.

39

[84] B. Kraftschik, W.J. Koros, Cross-Linkable Polyimide Membranes for Improved Plasticization Resistance and
Permselectivity in Sour Gas Separations, Macromolecules, 46 (2013) 6908-6921.
[85] J. Hao, P. Rice, S. Stern, Upgrading low-quality natural gas with H2S and CO2-selective polymer membranes:
Part I. Process design and economics of membrane stages without recycle streams, Journal of Membrane Science,
209 (2002) 177-206.
[86] J.T. Vaughn, W.J. Koros, J.R. Johnson, O. Karvan, Effect of thermal annealing on a novel polyamideimide
polymer membrane for aggressive acid gas separations, Journal of Membrane Science, 401402 (2012) 163-174.
[87] J.T. Vaughn, W.J. Koros, Analysis of feed stream acid gas concentration effects on the transport properties
and separation performance of polymeric membranes for natural gas sweetening: A comparison between a glassy
and rubbery polymer, Journal of Membrane Science, (2014).
[88] S. Sridhar, B. Smitha, T.M. Aminabhavi, Separation of Carbon Dioxide from Natural Gas Mixtures through
Polymeric Membranes - A Review, Separation & Purification Reviews, 36 (2007) 113-174.
[89] H. Hsieh, Inorganic membranes for separation and reaction, Elsevier, 1996.
[90] T.J. Kim, L.I. Baoan, M.B. Hgg, Novel fixed-site-carrier polyvinylamine membrane for carbon dioxide capture,
Journal of Polymer Science, Part B: Polymer Physics, 42 (2004) 4326-4336.
[91] D.M. D'Alessandro, B. Smit, J.R. Long, Carbon dioxide capture: prospects for new materials, Angewandte
Chemie International Edition, 49 (2010) 6058-6082.
[92] R. Quinn, J.B. Appleby, G.P. Pez, New facilitated transport membranes for the separation of carbon dioxide
from hydrogen and methane, Journal of Membrane Science, 104 (1995) 139-146.
[93] J. Ilconich, C. Myers, H. Pennline, D. Luebke, Experimental investigation of the permeability and selectivity of
supported ionic liquid membranes for CO2/He separation at temperatures up to 125 C, Journal of Membrane
Science, 298 (2007) 41-47.
[94] S. Shackley, C. Gough, Carbon capture and its storage: an integrated assessment, Ashgate Publishing, Ltd.,
2006.
[95] R. Quinn, D. Laciak, Polyelectrolyte membranes for acid gas separations, Journal of Membrane Science, 131
(1997) 49-60.
[96] X. He, M.-B. Hgg, T.-J. Kim, Hybrid FSC membrane for CO2 removal from natural gas: Experimental, process
simulation, and economic feasibility analysis, AIChE Journal, 60 (2014) 4174-4184.
[97] X. He, T.-J. Kim, M.-B. Hgg, Hybrid fixed-site-carrier membranes for CO2 removal from high pressure natural
gas: Membrane optimization and process condition investigation, Journal of Membrane Science, 470 (2014) 266274.
[98] S.A.M. Marzouk, M.H. Al-Marzouqi, M. Teramoto, N. Abdullatif, Z.M. Ismail, Simultaneous removal of CO2 and
H2S from pressurized CO2H2SCH4 gas mixture using hollow fiber membrane contactors, Separation and
Purification Technology, 86 (2012) 88-97.
[99] J. Jahn, W.A.P. van den Bos, A. Lysenko, A. Trusov, A. Volkov, L.J.P. van den Broeke, Membrane Contactors
for High Pressure Regeneration of Absorption Liquids used for Acid Gas Removal, (2012) 165-169.
[100] M. Hedayat, M. Soltanieh, S.A. Mousavi, Simultaneous separation of H 2S and CO2 from natural gas by hollow
fiber membrane contactor using mixture of alkanolamines, Journal of Membrane Science, 377 (2011) 191-197.
[101] A. Mansourizadeh, A.F. Ismail, Hollow fiber gas-liquid membrane contactors for acid gas capture: a review,
Journal of hazardous materials, 171 (2009) 38-53.
[102] H. Kreulen, C. Smolders, G. Versteeg, v.W. Swaaij, Selective removal of H2S from sour gases with
microporous membranes. Part II. A liquid membrane of water-free tertiary amines, Journal of Membrane Science,
82 (1993) 185-197.
[103] J.H. Meldon, K.A. Smith, C.K. Colton, The effect of weak acids upon the transport of carbon dioxide in
alkaline solutions, Chemical Engineering Science, 32 (1977) 939-950.
[104] S.L. Matson, C.S. Herrick, W.J. Ward, Progress on the selective removal of H2S from gasified coal using an
immobilized liquid membrane, Industrial & Engineering Chemistry Process Design and Development, 16 (1977)
370-374.
[105] S.L. Matson, E.K.L. Lee, D.T. Friesen, D.J. Kelly, Gas separation by composite solvent-swollen membranes,
in, Bend Research, Inc., USA . 1989, pp. 11 pp. Cont.-in-part of U.S. 14,737,166.
[106] C.M. Zimmerman, A. Singh, W.J. Koros, Tailoring mixed matrix composite membranes for gas separations,
Journal of Membrane Science, 137 (1997) 145-154.
[107] M. Jia, K.-V. Peinemann, R.-D. Behling, Molecular sieving effect of the zeolite-filled silicone rubber
membranes in gas permeation, Journal of Membrane Science, 57 (1991) 289-292.

40

[108] H. Yang, Z. Xu, M. Fan, R. Gupta, R.B. Slimane, A.E. Bland, I. Wright, Progress in carbon dioxide separation
and capture: A review, Journal of Environmental Sciences, 20 (2008) 14-27.
[109] A. Brunetti, F. Scura, G. Barbieri, E. Drioli, Membrane technologies for CO2 separation, Journal of Membrane
Science, 359 (2010) 115-125.
[110] S. Akbarnezhad, S.M. Mousavi, R. Sarhaddi, Sol-gel synthesis of alumina-titania ceramic membrane:
Preparation and characterization, Indian Journal of Science & Technology, 3 (2010).
[111] T. Suzuki, Y. Yamada, Physical and gas transport properties of novel hyperbranched polyimidesilica hybrid
membranes, Polymer Bulletin, 53 (2005) 139-146.
[112] Q. Hu, E. Marand, S. Dhingra, D. Fritsch, J. Wen, G. Wilkes, Poly(amide-imide)/TiO2 nano-composite gas
separation membranes: Fabrication and characterization, Journal of Membrane Science, 135 (1997) 65-79.
[113] S. Hassanajili, E. Masoudi, G. Karimi, M. Khademi, Mixed matrix membranes based on polyetherurethane
and polyesterurethane containing silica nanoparticles for separation of CO 2/CH 4 gases, Separation and
Purification Technology, 116 (2013) 1-12.
[114] O.G. Nik, X.Y. Chen, S. Kaliaguine, Amine-functionalized zeolite FAU/EMT-polyimide mixed matrix
membranes for CO2/CH4 separation, Journal of Membrane Science, 379 (2011) 468-478.
[115] S. Matteucci, Y. Yampolskii, B.D. Freeman, I. Pinnau, Transport of gases and vapors in glassy and rubbery
polymers, Materials science of membranes for gas and vapor separation, (2006) 1-47.
[116] J. Ahn, W.-J. Chung, I. Pinnau, M.D. Guiver, Polysulfone/silica nanoparticle mixed-matrix membranes for gas
separation, Journal of Membrane Science, 314 (2008) 123-133.
[117] N.B. McKeown, P.M. Budd, Polymers of intrinsic microporosity (PIMs): organic materials for membrane
separations, heterogeneous catalysis and hydrogen storage, Chemical Society Reviews, 35 (2006) 675-683.
[118] H.B. Park, C.H. Jung, Y.M. Lee, A.J. Hill, S.J. Pas, S.T. Mudie, E. Van Wagner, B.D. Freeman, D.J. Cookson,
Polymers with cavities tuned for fast selective transport of small molecules and ions, Science, 318 (2007) 254-258.
[119] J. Brandrup, E.H. Immergut, E.A. Grulke, A. Abe, D.R. Bloch, Polymer handbook, Wiley New York, 1999.
[120] V. Bondar, B. Freeman, Y.P. Yampolskii, Sorption of gases and vapors in an amorphous glassy
perfluorodioxole copolymer, Macromolecules, 32 (1999) 6163-6171.
[121] D. Melo, J. De Souza, M. Melo, A. Martinelli, G. Cachima, J. Cunha, Evaluation of the zinox and zeolite
materials as adsorbents to remove H 2 S from natural gas, Colloids and Surfaces A: Physicochemical and
Engineering Aspects, 272 (2006) 32-36.
[122] H. Maghsoudi, M. Soltanieh, Simultaneous separation of H 2 S and CO 2 from CH 4 by a high silica CHAtype zeolite membrane, Journal of Membrane Science, 470 (2014) 159-165.
[123] T. Merkel, V. Bondar, K. Nagai, B. Freeman, Y.P. Yampolskii, Gas sorption, diffusion, and permeation in poly
(2, 2-bis (trifluoromethyl)-4, 5-difluoro-1, 3-dioxole-co-tetrafluoroethylene), Macromolecules, 32 (1999) 8427-8440.
[124] H. Cong, M. Radosz, B.F. Towler, Y. Shen, Polymerinorganic nanocomposite membranes for gas
separation, Separation and Purification Technology, 55 (2007) 281-291.
[125] R.S. Prabhakar, B.D. Freeman, I. Roman, Gas and vapor sorption and permeation in poly (2, 2, 4-trifluoro-5trifluoromethoxy-1, 3-dioxole-co-tetrafluoroethylene), Macromolecules, 37 (2004) 7688-7697.
[126] J.H. Hodgkin, M.S. Liu, B.N. Dao, J. Mardel, A.J. Hill, Reaction mechanism and products of the thermal
conversion of hydroxy-containing polyimides, European Polymer Journal, 47 (2011) 394-400.
[127] N. Du, G.P. Robertson, I. Pinnau, M.D. Guiver, Polymers of Intrinsic Microporosity Derived from Novel
Disulfone-Based Monomers, Macromolecules, 42 (2009) 6023-6030.
[128] M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, J.C. Jansen, P. Bernardo, F. Bazzarelli, N.B. McKeown,
An efficient polymer molecular sieve for membrane gas separations, Science, 339 (2013) 303-307.
[129] N. Du, G.P. Robertson, I. Pinnau, M.D. Guiver, Polymers of Intrinsic Microporosity with Dinaphthyl and
Thianthrene Segments, Macromolecules, 43 (2010) 8580-8587.
[130] T. Emmler, . Heinrich, D. Fritsch, P.M. Budd, N. Chaukura, D. Ehlers, . R tzke, F. Faupel, Free volume
investigation of polymers of intrinsic microporosity (PIMs): PIM-1 and PIM1 copolymers incorporating
ethanoanthracene units, Macromolecules, 43 (2010) 6075-6084.
[131] R. Swaidan, B.S. Ghanem, E. Litwiller, I. Pinnau, Pure-and mixed-gas CO2/CH4 separation properties of
PIM-1 and an amidoxime-functionalized PIM-1, Journal of Membrane Science, 457 (2014) 95-102.
[132] N. Du, M.M.D. Cin, I. Pinnau, A. Nicalek, G.P. Robertson, M.D. Guiver, Azidebased CrossLinking of
Polymers of Intrinsic Microporosity (PIMs) for Condensable Gas Separation, Macromolecular rapid
communications, 32 (2011) 631-636.
[133] F.Y. Li, Y. Xiao, T.-S. Chung, S. Kawi, High-Performance Thermally Self-Cross-Linked Polymer of Intrinsic
Microporosity (PIM-1) Membranes for Energy Development, Macromolecules, 45 (2012) 1427-1437.

41

[134] N. Du, G.P. Robertson, J. Song, I. Pinnau, S. Thomas, M.D. Guiver, Polymers of Intrinsic Microporosity
Containing Trifluoromethyl and Phenylsulfone Groups as Materials for Membrane Gas Separation,
Macromolecules, 41 (2008) 9656-9662.
[135] C.A. Scholes, J. Jin, G.W. Stevens, S.E. Kentish, Competitive permeation of gas and water vapour in high
free volume polymeric membranes, Journal of Polymer Science Part B: Polymer Physics, 53 (2015) 719-728.
[136] Z. Lei, C. Dai, B. Chen, Gas solubility in ionic liquids, Chemical Reviews, 114 (2014) 1289-1326.
[137] R. Quinn, J.B. Appleby, G.P. Pez, Hydrogen sulfide separation from gas streams using salt hydrate chemical
absorbents and immobilized liquid membranes, Separation Science and Technology, 37 (2002) 627-638.
[138] P. Scovazzo, J. Kieft, D.A. Finan, C. Koval, D. DuBois, R. Noble, Gas separations using nonhexafluorophosphate [PF6]- anion supported ionic liquid membranes, Journal of Membrane Science, 238 (2004)
57-63.
[139] S. Saedi, S.S. Madaeni, A.A. Shamsabadi, PDMS coated asymmetric PES membrane for natural gas
sweetening: Effect of preparation and operating parameters on performance, The Canadian Journal of Chemical
Engineering, 92 (2014) 892-904.
[140] P. Iovane, F. Nanna, Y. Ding, B. Bikson, A. Molino, Experimental test with polymeric membrane for the biogas
purification from CO2 and H2S, Fuel, 135 (2014) 352-358.
[141] P. Dolej, V. Potulka, Z. Sedlkov, V. Jandov, J. Vejraka, E. Esposito, J.C. Jansen, P. Izk,
Simultaneous hydrogen sulphide and carbon dioxide removal from biogas by waterswollen reverse osmosis
membrane, Separation and Purification Technology, 131 (2014) 108-116.
[142] S. Amjad-Iranagh, K. Golzar, H. Modarress, Molecular simulation study of PAMAM dendrimer composite
membranes, Journal of molecular modeling, 20 (2014) 1-20.
[143] H.B. Park, S.H. Han, C.H. Jung, Y.M. Lee, A.J. Hill, Thermally rearranged (TR) polymer membranes for CO2
separation, Journal of Membrane Science, 359 (2010) 11-24.
[144] H. Li, B.D. Freeman, O.M. Ekiner, Gas permeation properties of poly(urethane-urea)s containing different
polyethers, Journal of Membrane Science, 369 (2011) 49-58.
[145] J.W. Simmons, Block polyurethane-ether and polyurea-ether gas separation membranes, in, 2005.
[146] A. Coady, T. Cooley, Removal of H2S and/or CO2 from a light hydrocarbon stream by use of gas permeable
membrane, in, 1978.
[147] T. Mohammadi, M.T. Moghadam, M. Saeidi, M. Mahdyarfar, Acid gas permeation behavior through poly
(ester urethane urea) membrane, Industrial & Engineering Chemistry Research, 47 (2008) 7361-7367.
[148] M. Amani, S. Amjad-Iranagh, K. Golzar, G.M.M. Sadeghi, H. Modarress, Study of nanostructure
characterizations and gas separation properties of poly (urethane-urea) s membranes by molecular dynamics
simulation, Journal of Membrane Science, (2014).
[149] C.S. Pomelli, C. Chiappe, A. Vidis, G. Laurenczy, P.J. Dyson, Influence of the interaction between hydrogen
sulfide and ionic liquids on solubility: Experimental and theoretical investigation, The Journal of Physical Chemistry
B, 111 (2007) 13014-13019.
[150] P. Li, M.R. Coleman, Synthesis of room temperature ionic liquids based random copolyimides for gas
separation applications, European Polymer Journal, 49 (2013) 482-491.
[151] Y.-I. Park, B.-S. Kim, Y.-H. Byun, S.-H. Lee, E.-W. Lee, J.-M. Lee, Preparation of supported ionic liquid
membranes (SILMs) for the removal of acidic gases from crude natural gas, Desalination, 236 (2009) 342-348.
[152] H. Sakhaeinia, V. Taghikhani, A.H. Jalili, A. Mehdizadeh, A.A. Safekordi, Solubility of H2S in 1-(2hydroxyethyl)-3-methylimidazolium ionic liquids with different anions, Fluid Phase Equilibria, 298 (2010) 303-309.
[153] K. Huang, D.-N. Cai, Y.-L. Chen, Y.-T. Wu, X.-B. Hu, Z.-B. Zhang, Thermodynamic validation of 1-alkyl-3methylimidazolium carboxylates as task-specific ionic liquids for H2S absorption, AIChE Journal, 59 (2013) 22272235.
[154] P. Bernardo, J.C. Jansen, F. Bazzarelli, F. Tasselli, A. Fuoco, K. Friess, P. Izk, V. Jarmarov, M. arkov,
G. Clarizia, Gas transport properties of Pebax/room temperature ionic liquid gel membranes, Separation and
Purification Technology, 97 (2012) 73-82.
[155] K. Huang, X.M. Zhang, Y. Xu, Y.T. Wu, X.B. Hu, Protic ionic liquids for the selective absorption of H2S from
CO2: Thermodynamic analysis, AIChE Journal, (2014).
[156] S. Mortazavi-Manesh, M.A. Satyro, R.A. Marriott, Screening ionic liquids as candidates for separation of acid
gases: Solubility of hydrogen sulfide, methane, and ethane, AIChE Journal, 59 (2013) 2993-3005.
[157] M. Rahmati-Rostami, C. Ghotbi, M. Hosseini-Jenab, A.N. Ahmadi, A.H. Jalili, Solubility of H2S in ionic liquids
[hmim][PF6], [hmim][BF4], and [hmim][Tf2N], Journal of Chemical Thermodynamics, 41 (2009) 1052-1055.

42

[158] A.H. Jalili, M. Safavi, C. Ghotbi, A. Mehdizadeh, M. Hosseini-Jenab, V. Taghikhani, Solubility of CO2, H2S,
and Their Mixture in the Ionic Liquid 1-Octyl-3-methylimidazolium Bis (trifluoromethyl) sulfonylimide, The Journal of
Physical Chemistry B, 116 (2012) 2758-2774.
[159] M. Ramdin, S.P. Balaji, J.M. Vicent-Luna, J.J. Gutierrez-Sevillano, S. Calero, T.W. De Loos, T.J. Vlugt,
Solubility of the Precombustion Gases CO2, CH4, CO, H2, N2, and H2S in the Ionic Liquid [bmim][Tf2N] from
Monte Carlo Simulations, The Journal of Physical Chemistry C, (2014).
[160] P. Scovazzo, Determination of the upper limits, benchmarks, and critical properties for gas separations using
stabilized room temperature ionic liquid membranes (SILMs) for the purpose of guiding future research, Journal of
Membrane Science, 343 (2009) 199-211.
[161] S. Hanioka, T. Maruyama, T. Sotani, M. Teramoto, H. Matsuyama, K. Nakashima, M. Hanaki, F. Kubota, M.
Goto, CO2 separation facilitated by task-specific ionic liquids using a supported liquid membrane, Journal of
Membrane Science, 314 (2008) 1-4.
[162] C.H. Lau, P. Li, F. Li, T.-S. Chung, D.R. Paul, Reverse-selective polymeric membranes for gas separations,
Progress in Polymer Science, 38 (2013) 740-766.
[163] T.K. Carlisle, E.F. Wiesenauer, G.D. Nicodemus, D.L. Gin, R.D. Noble, Ideal CO2/Light Gas Separation
Performance of Poly(vinylimidazolium) Membranes and Poly(vinylimidazolium)-Ionic Liquid Composite Films,
Industrial & Engineering Chemistry Research, 52 (2012) 1023-1032.
[164] P.J. Carvalho, J.A.P. Coutinho, The polarity effect upon the methane solubility in ionic liquids: a contribution
for the design of ionic liquids for enhanced CO2/CH4 and H2S/CH4 selectivities, Energy & Environmental Science,
4 (2011) 4614.
[165] H.Z. Chen, P. Li, T.S. Chung, PVDF/ionic liquid polymer blends with superior separation performance for
removing CO 2 from hydrogen and flue gas, International Journal of Hydrogen Energy, 37 (2012) 11796-11804.
[166] A.H. Jalili, M. Shokouhi, G. Maurer, M. Hosseini-Jenab, Solubility of CO2 and H2S in the ionic liquid 1-ethyl-3methylimidazolium tris(pentafluoroethyl)trifluorophosphate, The Journal of Chemical Thermodynamics, 67 (2013)
55-62.
[167] M.A. Malik, M.A. Hashim, F. Nabi, Ionic liquids in supported liquid membrane technology, Chemical
Engineering Journal, 171 (2011) 242-254.
[168] W. Schell, C. Wensley, M. Chen, K. Venugopal, B. Miller, J. Stuart, Recent advances in cellulosic membranes
for gas separation and pervaporation, Gas separation & purification, 3 (1989) 162-169.
[169] B. Bhide, A. Voskericyan, S. Stern, Hybrid processes for the removal of acid gases from natural gas, Journal
of Membrane Science, 140 (1998) 27-49.
[170] D. Amirkhanov, A. Kotenko, V. Rusanov, M. Tul'skii, Polymeric membranes for hydrogen sulfide separation
from natural gas, Polymer science. Series A, Chemistry, physics, 40 (1998) 206-212.
[171] R.W. Baker, K.A. Lokhandwala, Sour gas treatment process including membrane and non-membrane
treatment steps, in: United States Patent No. 5407466, 1995.
[172] X. Ren, J. Ren, M. Deng, Poly(amide-6-b-ethylene oxide) membranes for sour gas separation, Separation
and Purification Technology, 89 (2012) 1-8.
[173] B. Kraftschik, W.J. Koros, J.R. Johnson, O. Karvan, Dense film polyimide membranes for aggressive sour gas
feed separations, Journal of Membrane Science, 428 (2013) 608-619.
[174] W. Heilman, V. Tammela, J. Meyer, V. Stannett, M. Szwarc, Permeability of polymer films to hydrogen sulfide
gas, Industrial & Engineering Chemistry, 48 (1956) 821-824.
[175] S. Sridhar, B. Smitha, S. Mayor, B. Prathab, T.M. Aminabhavi, Gas permeation properties of polyamide
membrane prepared by interfacial polymerization, Journal of Materials Science, 42 (2007) 9392-9401.
[176] R.T.C. W.J. Koros, Separation of gaseous mixtures using polymer
membranes, in: Handbook of Separation Process Technology, John Wiley &
Sons, New York, 1987.
[177] M. Chenar, H. Savoji, M. Soltanieh, T. Matsuura, S. Tabe, Removal of hydrogen sulfide from methane using
commercial polyphenylene oxide and Cardo-type polyimide hollow fiber membranes, Korean J. Chem. Eng., 28
(2011) 902-913.
[178] R. McCaffrey, D. Cummings, Gas separation properties of phosphazene polymer membranes, Separation
Science and Technology, 23 (1988) 1627-1643.
[179] C.S.K. Achoundong, N. Bhuwania, S.K. Burgess, O. Karvan, J.R. Johnson, W.J. Koros, Silane Modification of
Cellulose Acetate Dense Films as Materials for Acid Gas Removal, Macromolecules, 46 (2013) 5584-5594.
[180] K.A. Berchtold, R.P. Singh, J.S. Young, K.W. Dudeck, Polybenzimidazole composite membranes for high
temperature synthesis gas separations, Journal of Membrane Science, 415-416 (2012) 265-270.
[181] H.K. S.A. Stern, A.Y. Houde, G. Zhou, Process for Separating Carbon Dioxide from Methane.

43

, in: US Patent 5,591,250, 1997.


[182] R.L. Heyd, F. McCandless, Separation of H2S from N2 By selective permeation through polymeric
membranes, Journal of Membrane Science, 2 (1977) 375-389.
[183] M. Saeidi, M. Tavakol Moghadam, M. Mahdyarfar, T. Mohammadi, Gas permeation properties of Seragel
membrane, AsiaPacific Journal of Chemical Engineering, 5 (2010) 324-329.
[184] C.J. Orme, M.K. Harrup, T.A. Luther, R.P. Lash, K.S. Houston, D.H. Weinkauf, F.F. Stewart, Characterization
of gas transport in selected rubbery amorphous polyphosphazene membranes, Journal of Membrane Science, 186
(2001) 249-256.
[185] C.J. Orme, F.F. Stewart, Mixed gas hydrogen sulfide permeability and separation using supported
polyphosphazene membranes, Journal of Membrane Science, 253 (2005) 243-249.
[186] O.V. Malykh, A.Y. Golub, V.V. Teplyakov, Polymeric membrane materials: new aspects of empirical
approaches to prediction of gas permeability parameters in relation to permanent gases, linear lower hydrocarbons
and some toxic gases, Advances in colloid and interface science, 164 (2011) 89-99.
[187] M.B. Shiflett, A. Yokozeki, Separation of CO2 and H2S using room-temperature ionic liquid [bmim][PF6], Fluid
Phase Equilibria, 294 (2010) 105-113.
[188] M. Safavi, C. Ghotbi, V. Taghikhani, A.H. Jalili, A. Mehdizadeh, Study of the solubility of CO2, H2S and their
mixture in the ionic liquid 1-octyl-3-methylimidazolium hexafluorophosphate: Experimental and modelling, The
Journal of Chemical Thermodynamics, 65 (2013) 220-232.
[189] R. Lungwitz, S. Spange, A hydrogen bond accepting (HBA) scale for anions, including room temperature
ionic liquids, New Journal of Chemistry, 32 (2008) 392-394.
[190] L. Crowhurst, P.R. Mawdsley, J.M. Perez-Arlandis, P.A. Salter, T. Welton, Solventsolute interactions in ionic
liquids, Physical Chemistry Chemical Physics, 5 (2003) 2790-2794.
[191] H. Salari, A.R. Harifi-Mood, M.R. Elahifard, M.R. Gholami, Solvatochromic probes absorbance behavior in
mixtures of 2-hydroxy ethylammonium formate with methanol, ethylene glycol and glycerol, Journal of solution
chemistry, 39 (2010) 1509-1519.
[192] A.H. Jalili, A. Mehdizadeh, M. Shokouhi, A.N. Ahmadi, M. Hosseini-Jenab, F. Fateminassab, Solubility and
diffusion of CO2 and H2S in the ionic liquid 1-ethyl-3-methylimidazolium ethylsulfate, The Journal of Chemical
Thermodynamics, 42 (2010) 1298-1303.
[193] A.H. Jalili, M. Rahmati-Rostami, C. Ghotbi, M. Hosseini-Jenab, A.N. Ahmadi, Solubility of H2S in ionic liquids
[bmim][PF6],[bmim][BF4], and [bmim][Tf2N], Journal of Chemical & Engineering Data, 54 (2009) 1844-1849.
[194] D. Wang, W.K. Teo, K. Li, Removal of H2S to ultra-low concentrations using an asymmetric hollow fibre
membrane module, Separation and Purification Technology, 27 (2002) 33-40.
[195] S.A.M. Marzouk, M.H. Al-Marzouqi, N. Abdullatif, Z.M. Ismail, Removal of percentile level of H2S from
pressurized H2SCH4 gas mixture using hollow fiber membrane contactors and absorption solvents, Journal of
Membrane Science, 360 (2010) 436-441.
[196] K. Kamide, S.-i. Manabe, H. Makino, T. Nohmi, H. Narita, T. Kawai, Surface diffusion flow on the pore wall
when gas permeates through a polymeric membrane, Polymer Journal, 15 (1983) 179-193.
[197] S.A. Stern, Y. Mi, H. Yamamoto, A.K.S. Clair, Structure/permeability relationships of polyimide membranes.
Applications to the separation of gas mixtures, Journal of Polymer Science Part B: Polymer Physics, 27 (1989)
1887-1909.
[198] L.M. Costello, W.J. Koros, Effect of structure on the temperature dependence of gas transport and sorption in
a series of polycarbonates, Journal of Polymer Science Part B: Polymer Physics, 32 (1994) 701-713.
[199] W.J. Koros, G.K. Fleming, Membrane-based gas separation, Journal of Membrane Science, 83 (1993) 1-80.
[200] W. Koros, M. Hellums, Gas separation membrane material selection criteria: differences for weakly and
strongly interacting feed components, Fluid Phase Equilibria, 53 (1989) 339-354.
[201] C.A. Scholes, S.E. Kentish, G.W. Stevens, Effects of minor components in carbon dioxide capture using
polymeric gas separation membranes, Separation & Purification Reviews, 38 (2009) 1-44.
[202] J. Barrie, K. Munday, Gas transport in heterogeneous polymer blends: I. Polydimethylsiloxane-g-polystyrene
and polydimethylsiloxane-b-polystyrene, Journal of Membrane Science, 13 (1983) 175-195.
[203] J. Barrie, M. Williams, H. Spencer, Gas transport in heterogeneous polymer blends: III. Alternating block
copolymers of poly (bisphenol-a carbonate) and polydimethylsiloxane, Journal of Membrane Science, 21 (1984)
185-202.
[204] Y. Maeda, D.R. Paul, Selective gas transport in miscible PPO-PS blends, Polymer, 26 (1985) 2055-2063.
[205] D. Paul, Gas transport in homogeneous multicomponent polymers, Journal of Membrane Science, 18 (1984)
75-86.

44

[206] K. Nagai, M. Mori, T. Watanabe, T. Nakagawa, Gas permeation properties of blend and copolymer
membranes composed of 1trimethylsilyl1propyne and 1phenyl1propyne structures, Journal of Polymer Science
Part B: Polymer Physics, 35 (1997) 119-131.
[207] A. Barnabeo, W. Creasy, L. Robeson, Gas permeability characteristics of nitrilecontaining block and random
copolymers, Journal of Polymer Science: Polymer Chemistry Edition, 13 (1975) 1979-1986.
[208] A. Senuma, Gas permeability coefficients of ethylenevinyl acetate copolymermodified poly
(dimethylsiloxane) membranes. Doublecolumn approach for twophase materials, Macromolecular Chemistry and
Physics, 201 (2000) 568-576.
[209] H.B. Park, S.Y. Ha, Y.M. Lee, Percolation behavior of gas permeability in rigid-flexible block copolymer
membranes, Journal of Membrane Science, 177 (2000) 143-152.
[210] A. Car, C. Stropnik, W. Yave, K.-V. Peinemann, PEG modified poly(amide-b-ethylene oxide) membranes for
CO2 separation, Journal of Membrane Science, 307 (2008) 88-95.
[211] W. Yave, A. Car, K.V. Peinemann, Nanostructured membrane material designed for carbon dioxide
separation, Journal of Membrane Science, 350 (2010) 124-129.
[212] R.S. Murali, S. Sridhar, T. Sankarshana, Y.V.L. Ravikumar, Gas Permeation Behavior of Pebax-1657
Nanocomposite Membrane Incorporated with Multiwalled Carbon Nanotubes, Industrial & Engineering Chemistry
Research, 49 (2010) 6530-6538.
[213] R.A. Hayes, Polyimide gas separation membranes, in, 1988.
[214] Y. Liu, R. Wang, T.-S. Chung, Chemical cross-linking modification of polyimide membranes for gas
separation, Journal of Membrane Science, 189 (2001) 231-239.
[215] M.E. Rezac, E. Todd Sorensen, H.W. Beckham, Transport properties of crosslinkable polyimide blends,
Journal of Membrane Science, 136 (1997) 249-259.
[216] C. Staudt-Bickel, W. J Koros, Improvement of CO2/CH4 separation characteristics of polyimides by chemical
crosslinking, Journal of Membrane Science, 155 (1999) 145-154.
[217] C. Wright, D. Paul, Feasibility of thermal crosslinking of polyarylate gas-separation membranes using
benzocyclobutene-based monomers, Journal of Membrane Science, 129 (1997) 47-53.
[218] J.D. Wind, C. Staudt-Bickel, D.R. Paul, W.J. Koros, The effects of crosslinking chemistry on CO2
plasticization of polyimide gas separation membranes, Industrial & Engineering Chemistry Research, 41 (2002)
6139-6148.
[219] W.J. Koros, D. Wallace, J.D. Wind, S.J. Miller, C. Staudt-Bickel, D.Q. Vu, Crosslinked and crosslinkable
hollow fiber mixed matrix membrane and method of making same, in, Google Patents, 2004.
[220] J.D. Wind, D.R. Paul, W.J. Koros, Natural gas permeation in polyimide membranes, Journal of Membrane
Science, 228 (2004) 227-236.
[221] J. Kim, W.J. Koros, D.R. Paul, Effects of CO 2 exposure and physical aging on the gas permeability of thin
6FDA-based polyimide membranes: Part 2. with crosslinking, Journal of Membrane Science, 282 (2006) 32-43.
[222] A.M. Hillock, W.J. Koros, Cross-linkable polyimide membrane for natural gas purification and carbon dioxide
plasticization reduction, Macromolecules, 40 (2007) 583-587.
[223] A.M. Hillock, S.J. Miller, W.J. Koros, Crosslinked mixed matrix membranes for the purification of natural gas:
effects of sieve surface modification, Journal of Membrane Science, 314 (2008) 193-199.
[224] J.D. Wind, C. Staudt-Bickel, D.R. Paul, W.J. Koros, Solid-state covalent cross-linking of polyimide
membranes for carbon dioxide plasticization reduction, Macromolecules, 36 (2003) 1882-1888.
[225] R. Dawson, A.I. Cooper, D.J. Adams, Nanoporous organic polymer networks, Progress in Polymer Science,
37 (2012) 530-563.
[226] Y. Li, T.-S. Chung, Molecular-level mixed matrix membranes comprising Pebax and POSS for hydrogen
purification via preferential CO2 removal, International Journal of Hydrogen Energy, 35 (2010) 10560-10568.
[227] R. McKee, M. Changela, G. Reading, CO2 removal-memebrane plus amine, Hydrocarbon Processing, 70
(1991) 63-65.
[228] B. Bhide, S. Stern, Membrane processes for the removal of acid gases from natural gas. I. Process
configurations and optimization of operating conditions, Journal of Membrane Science, 81 (1993) 209-237.
[229] M. Ohno, O. Ozaki, H. Sato, S. Kimura, T. Miyauchi, Radioactive rare gas separation using a separation cell
with two kinds of membrane differing in gas permeability tendency, Journal of Nuclear Science and Technology, 14
(1977) 589-602.
[230] J. Perrin, S. Stern, Modeling of permeators with two different types of polymer membranes, AIChE Journal,
31 (1985) 1167-1177.
[231] J. Perrin, S. Stern, Separation of a heliummethane mixture in permeators with two types of polymer
membranes, AIChE Journal, 32 (1986) 1889-1901.

45

[232] J. Hao, P.A. Rice, S.A. Stern, Upgrading low-quality natural gas with H2S- and CO2-selective polymer
membranes: Part II. Process design, economics, and sensitivity study of membrane stages with recycle streams,
Journal of Membrane Science, 320 (2008) 108-122.
[233] F. Ahmad, K.K. Lau, A.M. Shariff, G. Murshid, Process simulation and optimal design of membrane
separation system for CO2 capture from natural gas, Computers & Chemical Engineering, 36 (2012) 119-128.
[234] L. Peters, A. Hussain, M. Follmann, T. Melin, M.-B. Hgg, CO2 removal from natural gas by employing amine
absorption and membrane technologyA technical and economical analysis, Chemical Engineering Journal, 172
(2011) 952-960.
[235] F. Ahmad, K. Lau, A. Shariff, Y. Fong Yeong, Temperature and pressure dependence of membrane
permeance and its effect on process economics of hollow fiber gas separation system, Journal of Membrane
Science, 430 (2013) 44-55.
[236] S.S.M. Lock, K.K. Lau, A.M. Shariff, Effect of recycle ratio on the cost of natural gas processing in
countercurrent hollow fiber membrane system, Journal of Industrial and Engineering Chemistry, 21 (2015) 542-551.

46

Polymer Membranes for Acid Gas Removal from Natural Gas


Gigi George, Nidhika Bhoria, Sama AlHallaq, Ahmed Abdala, Vikas Mittal *
The Petroleum Institute, Abu Dhabi, United Arab Emirates.

1.

Figure 1: Structure of fluorine-containing polyimide 6FDA-HAB. Reproduced from reference [81] with permission
from the Journal of Membrane Science.

47

2.

Figure 2: Cross-linking reaction scheme via thermally activated transesterification. Reproduced from reference [80]
with the permission of Macromolecules.

48

3.

Figure 3: Schematic of the facilitated transport mechanism. Reproduced from reference [86] with the permission
from Polymer Physics.

49

4.

Figure 4: Membrane contactor setup for simultaneous absorption of H2S and CO2 from a gas mixture.
Reproduced from reference [96] with the permission from the Journal of Membrane Science.

50

5.

Figure 5: Illustration of nanogap formation in polymer nanocomposite membranes. Reproduced from reference
[120] with the permission from Separation and Purification Technology.

51

6.

Figure 6: CO2/CH4 upper bound relationship.[34, 114] Reproduced from reference [34] with the permission from
the Journal of Membrane Science.

52

7.

Figure 7: Typical polyurethane urea synthesis steps. Reproduced from reference [140] with the permission of the
Journal of Membrane Science.

53

8.

Figure 8: Structure of poly(ether urethane urea). Reproduced from reference [78] with the permission of the Journal
of Membrane Science.

54

9.

Figure 9: Adsorption isotherm of acid gases in polyurethane urea. Reproduced from reference [144] with the
permission of the Journal of Membrane Science.

55

10.

Figure 10: H2S permeability of various polymer membranes

56

11.

Figure 11: selectivity- permeability relationship for H2S/CH4

57

12.

Figure 12: H2S permeability and H2S/CO2 selectivity relationship

58

13.

Figure 13: Upper bound plots with respect to lowest CH4 permeability

59

14.

Figure14: Processing costs as a function of both CO2 and H2S concentrations in feed for the optimal process
configurations with and without recycling. (A) Single stage with CO2/CH4-selective membranes. (B) Two stages in
series without recycling, with H2S/CH4-selective membranes in rst stage and CO2/CH4-selective membrane in
second stage. (C) Single stage with H2S/CH4-selective membranes. (D) Two stages in series with recycle
streams. Reproduced from reference [228] with the permission from the Journal of Membrane Science.

60

P*
Polymer
(CO2)

CO2/
H2S

Feed
Press
ure
Bar

H2S) %

P*

CO2/
CH4

(H2

H2S
/CH4

S)

Gas composition

Fe
ed
T
C

Pure(P) or Mixed
(M) (CH4/CO2/

Ref
.

Cellulose
Acetate

10-14

1520

40

13.78

[16
8]

Cellulose
Acetate

8.9

20-25

50

[64
]

Cellulose
Acetate

8.9

21

19

[16
9]

Cellulose
Acetate

2.43

22.1

2.1
3

19.4

1.14

35

10

M (65/29/6)

77.5

6.9

239

21

0.32

35

10

M(70.8/27.9/1.3)

55.8

6.9

183

22.6

0.30

35

10

M(69.4/18.1/12.5)

58.8

13

271

58

0.22

35

10

M(70.8/27.9/1.3)

62.2

12.2

280

54.9

0.22

35

10

M(69.4/18.1/12.5)

197

6.1

613

19

0.32

35

10

M(70.8/27.9/1.3)

195

5.6

618

18

0.32

35

10

M(69.4/18.1/12.5)

44.7

17

199

74

0.22

35

10

M(70.8/27.9/1.3)

50.8

15

223

66

0.22

35

10

M(69.4/18.1/12.5)

22.4

22

102

102

0.22

20

10

M(70.8/27.9/1.3)

25.4

20

123

95

0.21

20

10

M(69.4/18.1/12.5)

Poly(ester
urethane urea)

11

55

25

Up to
10

Matrimid 5218

10.6

9.5

26.53

M (Balance/4/800
ppm)

Poly(ether
urethane) PU1

Poly(ether
urethane) PU3

Poly(ether
urethane urea)
PU 2

Poly(ether
urethane urea)
PU4

6 FDA-IPDA

21.4

10.5

26.66

M (Balance/4/800
ppm)

Silicone
rubber

3.3

6.9

6.46

M (Balance/4/800
ppm)

[82
]

[17
0]

[17
1]

61

Poly(ether
urethane urea)

11.34

43

35

10

M (91.6/5.4/3)

15.96

22

35

10

M (93.2/6.2/0.6)

9.92

27

35

30

M (91.6/5.4/3)

13.07

18

35

30

M (93.2/6.2/0.6)

[14
7]

M (46% CO,
PTMSP

18000

0.85

23

1.38

10.5%CO2, 1.5%

H2S, balance
hydrogen)
M (36.5 % CO,

4400

0.66

21

6.89

PDMS

[76
]

11.7 % CO2, 0.7


% H2S, balance
hydrogen)
M

3.3

6.9

6.55

(650ppm/4%CO2
/bal CH4)

PDMS coated
PES

116.07

PEEK module
membrane

[17
1]

43.87

28.
a
86

10.6

4.18

25

10

M (97.5/2.1/0.4)

[13
9]

56

7.5

M(53.5/40.2/0.2)

[14
0]

TFC Sterlitech
I

2.2

M ( 53.3/46.6/0.1)

TFC Sterlitech
II

2.2

M ( 53.3/46.6/0.1)

[14
1]

TFC low P
membrane

19

21

3.5

0.9

M ( 53.3/46.6/0.1)

TFC high P
membrane

5.6

14.8

6.2

17.5

0.84

M ( 53.3/46.6/0.1)

Pebax 4011

16

70

26.66

M
(95.79/4.12/870pp
m)

Pebax MX
1657

[17
1]

M( 36.5 % CO,
69.1

0.27

21

13.1

11.7 % CO2, 0.7


% H2S, balance
hydrogen)

[76
]

62

69

14.1

248

50.6

0.27

35

10

M (70.8/27.9/1.3)

[82
]

89

126

0.71

25

[17
2]

M( 36.5 % CO,
11.7 % CO2, 0.7

122

0.22

21

190
psig

155

11.2

695

50.4

0.22

35

10

M
(69.4/18.1/12.5)

122

12

553

54

0.22

35

10

M (70.8/27.9/1.3)

Pebax MX 1041

39.7

11

175

49

0.23

35

10

M (70.8/27.9/1.3)

Pebax 4033
SA00

84.4

6.5

312

24

0.27

35

10

M (70.8/27.9/1.3)

Pebax 3533
SA00

243

5.7

888

21

0.27

35

10

M (70.8/27.9/1.3)

Pebax 6333
SA00

7.4

3.9

37.
8

20

0.19

35

10

M (70.8/27.9/1.3)

Pebax 7233
SA00

4.1

8.2

7.6

15

0.54

35

10

M (70.8/27.9/1.3)

Pebax MV 3000
SA 01

100

10

487

49

0.21

35

4.48

55.6

32.1

25.
4

14.7

0.58

35

49

M (70/20/10)

31.1

23.
6

Pebax MX 1074

6FDADAM:DABA
(3:2)

50.8

14.4

0.61

35

49

% H2S, balance
hydrogen)

M (70/20/10)

[76
]

[82
]

[87
]

[17
3]

M( 36.5 % CO,
PC

6.5

4.3

21

6.89

11.7 % CO2, 0.7


% H2S, balance
hydrogen)
M( 36.5 % CO,

PSF

3.8

3.9

21

13

PSF

9.80.3
b
4

7.2
0.

[76
]

11.7 % CO2, 0.7


% H2S, balance
hydrogen)
P

[14
2]

63

76

PSF/CSTHA(50)

60.30.
b
67

55.
10
.43

0.26

PSF/CSTHA(50)/PAMA
M

Nylon - 6

45.60.
b
78

44.
00
.78
b

0.088

[17
4]

For CO2 - M(
36.5 % CO, 11.7
% CO2, 0.7 %

H2S, balance
Cytop

17

27

21

6.89

hydrogen)
For H2S M
(15% H2S, 85%

N2)
M( 46% CO,
TFE/PMVE/8C
NVE

28

37

6.89

10.5%CO2, 1.5%

H2S, balance
hydrogen)
M( 46% CO,

Teflon AF 1600

680

6.8

23

1.38

10.5%CO2, 1.5%

H2S, balance
hydrogen)

M( 46% CO,
Teflon AF 2400

2300

5.6

23

1.38

10.5%CO2, 1.5%

[76
]

H2S, balance
hydrogen)

TFC polyamide

16.6

15.2
PVBTAF

12.8

52.
b
3

14.4

51.
b
6

9.6

1.18

40.5

0.32

49.1

0.29

1920

0.12

10

30

10

4.85

M (79.7/10/10.3)

[17
5]

64

0.95

8.7

0.95

PEI/PEBA1657

12.3

PEI/PDMS/PEB
A1657/PDMS

157

1730

0.11

30

6.22

M (79.7/10/10.3)

8.3

2100

0.11

30

7.59

M (79.7/10/10.3)

11.
b
5

25

25

196
b

275
a
0

0.98

25

42

6.2

8.1

5.29

35

4.48

49

10.3

4.73

35

4.48

5.0

10.9

4.32

35

4.48

3.56

35

4.55

PDMS

2700

6F-PAI-1

32.8

6F-PAI-2

14.2

6F-PAI-3

21.6

47

0.7 8

52.8

[95
]

[81
,
87,
17
2,
17
6]

0.2
Torlon

14.8

103

0.6

1.9

[86
]

0.02

Cardo type
polyimide

M (101-401 ppm

6.89

4.8

12

10

0.4

30

2.1

25

1.59

12

25

1.59

PTBP

17

10

16

9.4

1.06

30

2.1

PDTBP

27

5.4

20

1.35

30

2.1

8.66

29.5

8.7
1

29.7

0.99

35

34.5

M (60/20/20)

27.5

19.1

39.
7

27.4

0.69

35

48.3

M (60/20/20)

129.4

21.8

204

34.3

0.63

35

34.5

M (60/20/20)

136

20

190

27.5

0.71

35

48.3

M (60/20/20)

PPOP

Pure CA

GCV- modified
CA

H2S in CH4)

[17
7]
[80
]
[17
8]

[80
]

[17
9]

65

M (H2: 55%;
PBI composite
membrane

0.15

0.0
a
05

250

3.34

CO2: 41%; CO:


1%; CH4: 1%;

[18
0]

: N2 1% and

[(CH3)4N]F.4H2
O) / celgard
membrane

H2S: 1%)
109

18.7

813

140

0.13

50

1.15

M (89.77/5.1/5.3)

[13
7]

6FDA-HAB

60

15

35

10

[18
1]

Modified
vinylidene
fluoride films

5.3

23

4.6

23.
b
2

3801140
cm Hg

Seragel
membrane

3.75

M (5% H2S,
(95%N2)

[18
2]

M (3.91% H2S,
96% CH4)

[18
3]

Mixture of 3%
each Ar, N2,

17.8

113
a
0

78.8

6.4

114
b
0

59.37

45

34.6

3.07

PVTMS

1600

7.27

350

1.59

SILAR

2000

3.66

119
5

2.12

[MDEAH][Ac]

0.11

30

[MDEAH][For]

0.09

30

[DMEAH][Ac]

0.06

30

[DMEAH][For]

0.06

30

Bmim-BF4

19.6

57.4

0.34

25

20

20.1

56.1

0.36

25

20

Polymer 3

250

226.7
GENERON

Emim-CH3SO4

O2,CH4,CO2,
and H2S in He

[18
4,
18
5]

[18
6]

[15
5]

66

HOemim-NO3

20.13

62.9

0.34

25

20

N4111-Cl

23.8

75.5

0.31

25

20

Pmg-L

20.6

58.6

0.34

25

20

Tmg-TCA

22.3

72.6

0.30

25

20

[15
6]

[bmim][PF6]

3.2-4

[18
7]

[C2mim][eFAP]

1.82

[16
6]

[C8mim][PF6]

3.82

[18
8]

[P66614][NTf2]

12.5

80

[C4mim][NTf2]

13

10

80

[hmim][Tf2N]

0.47

40

[15
7]

[emim][EtSO4]

0.62

30

[15
2]

[omim][Tf2N]

0.67

30

[15
8]

[16
4,
19
0]
[16
4,
19
1]

2mHEAPr

42

27

80

[16
4,
18
9]

[C4mim][CH3S
O3]

45

40

80

[hemim][Tf2N]

0.43

30

[hemim PF6]

0.31

30

[hemim] [TfO]

0.38

30

[hemim][BF4]

0.29

30

[19
2]

[bmim][Tf2N]

8.0

0.33

60

[15

[15
2]

67

9]
[bmim][BF4]

PVDF + BMIM
BF4

30-180

25-45

160
110
0

130260

0.26

30

[19
3]

[15
1,
16
4]

M (17.91159
PVDF + NaCO3

ppm H2S in

CH4)
PPO

ePTFE fibers

7.9

Up to
50 bar

M (101-401 ppm

H2S in CH4)
M (2% H2S from
pressurized H2S

CH4)

[19
4]

[17
7]

[19
5]

: selectivity
P: permeability /permeance/flux
*units of, P:
with no superscript is Permeability in Barrer
a
: Permeance in GPU;
b
:Permeance in 10-6, cm3/cm2 s cm Hg,
c:
Flux in m3 (STP)/m2hr

Table1: Gas transport properties of various polymer membranes used for acid gas removal from natural gas.

68

Feed composition

5-40 mol% CO2, 1 mol% N2 , 1 mol% C2H6,


Balance CH4

Feed flow rate

35 MMscfd

Feed pressure

55.2 bar

Permeate pressure

1.4 bar

Membrane thickness

1000

Retentate composition

2 mol% CO2

Membrane material

Cellulose Acetate
0.4257 x 10-10

CH4 permeability

9 x 10-10

CO2 permeability

0.18 x 10-10

C2H6 permeability
N2 permeabilitya

0.4257 x 10-10
a

: (cm3(STP).cm)/(s.cm2.cmHg)

Table 2: Feed composition as summarized by Bhide. Reproduced from reference [169] with permission from the
Journal of Membrane Science.

69

Feed composition

0-40 mol% CO2, 0-40 mol% H2S, Balance CH4

Feed flow rate

35 MMscfd

Feed pressure

55.2 bar

Permeate pressure

1.4 bar

Membrane thickness

1000

Membrane material

Poly(ether urethane urea) , H2S selective

Methane permeabilitya

2 x 10-10

CO2 permeabilitya

32 x 10-10

H2S permeabilitya

150 x 10-10

Membrane material

Fluorine-containing polyimide (6FDA-HAB) , CO2


selective

Methane permeabilitya

0.1 x 10-10

CO2 permeabilitya

6 x 10-10

H2S permeabilitya

1.5 x 10-10
a

:(cm3(STP).cm)/(s.cm2.cmHg)

Table 3: Feed condition and the membrane properties. Reproduced from reference [85] with permission from the
Journal of Membrane Science.

70

Highlights

1. Polymer membranes for simultaneous separation of CO2 & H2S from natural gas.
2. Examining simultaneously polymer membrane material, performance and economics.
3. Polymer material selection for CO2 & H2S removal (Table 1).
4. Material - gas transport performance analysis (Section 4).

71

You might also like