Open access peer-reviewed chapter

Polymer-Based Membranes for C3+ Hydrocarbon Removal from Natural Gas

Written By

John Yang, Milind M. Vaidya, Sebastien A. Duval and Feras Hamad

Submitted: 05 February 2022 Reviewed: 24 February 2022 Published: 21 April 2022

DOI: 10.5772/intechopen.103903

From the Edited Volume

Natural Gas - New Perspectives and Future Developments

Edited by Maryam Takht Ravanchi

Chapter metrics overview

293 Chapter Downloads

View Full Metrics

Abstract

Natural gas can contain significant amounts of impurifies, including CO2, H2S, N2, He, and C3+ hydrocarbons. These C3+ hydrocarbons are valuable chemical feedstocks and can be used as a liquid fuel for power generation. Membrane-based separation technologies have recently emerged as an economically favorable alternative due to reduced capital and operating cost. Polymeric membranes for the separation and removal of C3+ hydrocarbons from natural gas have been practiced in chemical and petrochemical industries. Therefore, these industries can benefit from membranes with improved C3+ hydrocarbon separation. This chapter overviews the different gas processing technologies for C3+ hydrocarbon separation and recovery from natural gas, highlighting the advantages, research and industrial needs, and challenges in developing highly efficient polymer-based membranes. More specifically, this chapter summarizes the removal of C3H8 and C4H10 from CH4 by prospective polymer architectures based on reverse-selective glassy polymers, rubbery polymers, and its hybrid mixed matrix membranes. In addition, the effect of testing conditions and gas compositions on the membrane permeation properties (permeability and selectivity) is reviewed.

Keywords

  • glassy polymers
  • rubbery polymers
  • membrane separation
  • C3+ hydrocarbons
  • permeation property
  • C3+/CH4 separation

1. Introduction

Natural gas is one of the important and primary sources of global energy. In addition to its primary importance as a fuel, natural gas is also a raw material and source of hydrocarbons for petrochemical feed stocks. According to 2021 BP statistics, there were an estimated 6,642 trillion cubic feet (Tcf) of total world proved reserves of natural gas [1]. The worldwide natural gas production and consumption have been rising over the past 20 years. In 2020, natural gas production and consumption worldwide amounted to roughly 3.85 and 3.82 trillion cubic meters [1].

Raw natural gas contains primarily of methane (CH4) as the prevailing element but also comprises significant amounts of impurities such as nitrogen (N2), helium (He), acid gases (carbon dioxide (CO2) and hydrogen sulfide (H2S)), heavy hydrocarbons (C3+), mercaptans, water vapor, BTEX (benzene, toluene, ethylbenzene and xylene) etc. These impurities must be removed to meet the pipeline quality standard specifications for transport and processing and avoid pipelines corrosion. The operational natural gas plant delivers pipeline-quality dry natural gas that can be used as fuel by residential, commercial, and industrial consumers, or as a feed stocks for downstream chemical synthesis (Figure 1) [2].

Figure 1.

World natural gas final consumption by sector in 2019 [2].

Membrane-based separation technology has been pointed out as a key technology in the chemical industry [3, 4, 5]. It has benefits for small-to-medium scale separation where it exhibits higher energy efficiency, simplicity in operation, compact process design compared to other separation technologies. The global market for gas separation membrane estimated at US$822.1 MM in the year 2020 and is projected to reach a revised size of US$1.1 Billion by 2027, growing at the compound annual growth rate (CAGR) of 5.6% over the analysis period 2020–2027 [6]. Polymeric membrane-based separation technology for natural gas processing was first commercialized in the 1980s [7], and today it is widely used in variety of gas separation applications. Representative gas pairs needing separation in these applications is shown in Table 1 [8, 9, 10], some of the leading industrial membrane producers and their principal glassy and/or rubbery membrane materials used in gas separation is listed in Table 2 [7, 11, 12, 13, 14, 15].

ApplicationCommon gas separationProducer
• Nitrogen generation/oxygen enrichment• N2/O2• Air Products, Praxair, Air Liquide, MTR
• Hydrogen recovery• H2/CH4• Air Products, Ube, Air Liquide, Praxair, MTR
• Ammonia purge gas• H2/N2• Air Products, Ube, Air Liquide, Praxair, MTR
• Syngas ratio adjustment• H2/CO• Air Products, Ube, Air Liquide, Praxair, MTR
• Acidic gas separation• CO2/CH4• NATCO, Air Products, UOP, Uber, Air Liquide, MTR
• Natural gas dehydration• H2O/CH4• Kvaerner, Air Products, MTR
• Sour gas treatment• H2S/CH4• NATCO, Air Products, Uber, Kvaerner, MTR
• Helium separation• He/CH4• Air Products, Air Liquide, MTR
• Helium recovery• He/N2• Air Products, Air Liquide, MTR
• Heavy hydrocarbon recovery• C3+/CH4• MTR, ABB
• Air dehydration• H2O/Air• Air Products, Parker Balston, Praxair, Ube, MTR

Table 1.

Primary current industrial gas separation for polymer membranes [8, 9, 10].

Polymer typeTypical polymer usedMembrane module typeCompany
GlassyPolysulfoneHollow fiberAir Products
Polyimide/polyaramidHollow fiberAir Liquide
PolyimideHollow fiberUbe
PolyimideHollow fiberParker-Hannifin
PolyimideHollow fiberEvonik
PolyimideHollow fiberPraxair
Polyimide/polysulfoneHollow fiberGrasys
Cellulose acetateSpiral woundUOP, Kvaerner
Cellulose acetateSpiral woundW. R. Grace
Cellulose acetateSpiral woundMTR
Cellulose acetateSpiral woundFuji Film
Cellulose acetateHollow fiberSchlumberger (Natco)
Ethyl celluloseHollow fiberAir Liquide
Poly (phenylene oxide)Hollow fiberAquila
Perfluoro polymerSpiral woundABB/MTR
Tetrabromo polycarbonateHollow fiberGeneron (MG)
RubberyPoly(ether-b-amide) copolymerSpiral woundMTR
Poly(ether-b-amide) copolymerPlate and frameMTR
Poly(ether-b-amide) copolymerHollow fiberAir Liquide
PolysiloxanePlate and frameGKSS
PolysiloxaneSpiral woundMTR
PolysiloxaneHollow fiberAir Liquide

Table 2.

Principal polymer membrane materials, modules, and producers [7, 11, 12, 13, 14, 15].

Separation of C3+ hydrocarbons (e.g. propane (C3H8), butane (C4H10)) and their removal from natural gas not only is necessary to prevent condensation during transportation by reducing the dew point and heating value to pipeline specifications, but also it is economically attractive to recover C3+ hydrocarbons since they are often of greater value when used as chemical feedstocks, or as a liquid fuel for power generation. In general, both glassy and rubbery polymers are used for this application. Glassy polymeric membranes are in general diffusivity selective and traditional glassy polymer membrane based upon cellulose acetate, polyimide, Tetra Bromo polycarbonate, polysulfone have been widely utilized for a few decades for CO2 removal from natural gas. Only few glassy polymers (e.g. perfluoro-based polymers) have demonstrated remarkable abilities to separate hydrocarbons from natural gas but their application has been limited due to highly aging and plasticization [16, 17, 18, 19, 20, 21]. On the other hand, rubbery polymeric membranes are solubility selective and are mainly used in gas/vapor separation processes for separating hydrocarbons from their mixtures based on gas condensability. Commercially, poly(dimethylsiloxane) (PDMS) based rubbery siloxane membranes have been applied to separate C3H8 and C4H10 from CH4 [14, 22, 23, 24, 25] and other gas pairs (e.g. O2/N2, CO2/CH4, H2/N2, He/N2, He/H2, CO2/N2, N2/CH4, H2/CO2, He/CO2) [26, 27, 28, 29, 30, 31, 32] from natural gas; however, development of more selective, higher-flux and cost-effective rubbery membrane materials can further help improve economics of the C3+ hydrocarbon recovery.

In this chapter, the application of synthetic polymeric membranes for C3+ hydrocarbon separation and removal (C3+/CH4) from natural gas is reviewed. The review covers available glassy and rubbery polymer membrane materials, as well as its hybrid mixed matrix membranes. Their transport properties (permeability and selectivity) as well as the effect of testing conditions and feed compositions on the membrane separation performance are reviewed.

Advertisement

2. Outline of C3+ hydrocarbon separation membrane techniques

The processing of both associated gas (AG) and non-associated gas (NAG) from oil and natural gas wellheads into pipeline-quality dry natural gas can be quite complex. Generally, the Master Gas System (MGS) in the gas plants consists of three main units: (1) gas-oil separating plants (GOSPs), (2) gas plants and (3) fractionation plants as shown in Figure 2 (top). A typical natural gas treatment and separation processing plant whose simplified schematic representation shown in Figure 2 (bottom) consists five main processes to remove various impurities from raw natural gas. C3+ hydrocarbons are extracted as byproducts in the production of natural gas and oil, and natural gas processing is by far the most significant, contributing 90%+ production of C3+ hydrocarbons. There are three conventional technologies to separate C3+ hydrocarbons from natural gas: refrigeration, lean oil absorption and cryogenic [13]. These common processes are costly and energy intensive, which is reflected in the price and available capability of the finished gas (Table 3).

Figure 2.

Top: block diagram showing key process steps in gas treatments and separation plants; bottom: simplified sketch of natural gas separation processes consisting five main processes.

Processing technologyTechnology VintageAvailable capacityC3+ hydrocarbons extractionFuel consumption
Refrigeration1940–196010%large C3, most of C4Least
Lean Oil Absorption1960–197520%90%+ of C3 and 30% C2Higher
Cryogenic1975 on50%70–99% C2Higher

Table 3.

Available gas processing technology for C3+ heavy hydrocarbons recovery [12].

The refrigerated condensation process is a conventional way to separate C3+ hydrocarbon components from the gas stream through a sequence of conventional distillation columns operating down to −40°C [33]. The refrigeration process is pricey and energy intensive but can extract a large percentage of C3H8 and most of the C4+ gases. There was no major improvement of heavy hydrocarbon recovery process since 1910s until the lean oil absorption process was developed.

Lean oil absorption plants were the type of processing plant built in the late of 1960s. These plants were the next evolution from the refrigeration plants and can extract 90%+ of the C3+ in the gas stream and about 30% of the ethane by bubbling the gas through a chilled absorption oil operating at approximately 0°C. The fuel consumption of this type of plant is higher than that of the refrigeration plant [34].

Cryogenic plants became prevalent in the 1970s with the development of Turbo-expander technology with great economic advantages for natural gas liquid (NGL) recovery [35]. The 1st generation of this technology could extract 70%+ of ethane (C2) from the gas. Today, ∼99% extraction of ethane can be recovered with modified cryogenic process due to the increased pressure reduction involved in the process [34]. Highly energy intensive for regeneration, tendency for block of process of equipment and the use of flammable cryogenic fluids are one the main disadvantages of cryogenic separation.

Membrane-based separation technology can be competitive in the processing of hydrocarbon recovery from natural gas. Inorganic membranes, such as MFI-type zeolite [36, 37, 38] and MOF [39], can be appealing due to their unique properties with well-define pore structure and high chemical and mechanical stabilities. These MFI membranes exhibit high C3+/CH4 selectivity, but rather low overall permeance, high capital cost and difficulty of scaling up, hence has hardly found industrial usage. Polymeric membranes have been used for the separation and removal of C3+ hydrocarbons from natural gas, usually with moderate selectivity [14]. The recovery of C3+ hydrocarbons is currently the second biggest market for membranes in natural gas processing, after acidic gas separation [11]. Compared to conventional separation methods, membrane-based separation technologies entail low capital costs, high energy efficiency and constitute a reliable option for separating hydrocarbon mixtures.

Glassy polymeric membrane-based separation process is a separation process shown in Figure 3A where gas mixtures consisting of two or more components are separated by a membrane into a “C3+ enriched” retentate stream and a “C3+ lean” permeate stream. The glassy polymer membrane separates gas mixtures by providing a permeable barrier that allows compounds to move through at specific rates, it separates gas mixtures principally by size or diffusivity. On the other hand, the feed gas mixtures are separated by a rubbery polymer membrane into a “C3+ lean” retentate stream and a “C3+ enriched” permeate stream (Figure 3B). The permeation of gas molecules across the rubbery membrane depends mainly on their solubilities or condensabilities. For example, Membrane Technology and Research (MTR) developed VaporSep® process using rubbery polymer composite membranes to treat high pressure hydrocarbon-rich feed gas mixtures (Figure 4) [11, 13]. C3+ hydrocarbons and the BTEX aromatics all permeate preferentially and are recompressed and cooled by a fan-cooled heat exchanger to condense higher hydrocarbons, while CH4 is then recirculated to the feed. The size and cost of the compressor system is often larger than membrane unit, so the selection of higher selectivity of novel rubbery membrane in this processing application can reduce capital expenditures (CAPEX) and operating expenses (OPEX) to the gas processing business.

Figure 3.

Schematic of (A) glassy and (B) rubbery polymer membrane-based separation process from natural gas.

Figure 4.

Flow scheme of a membrane C3+ hydrocarbon recovery by means of dew point control unit (MTR/polysiloxane membranes/spiral-wound modules) [13].

Commercially available rubbery polymer membranes have been applied to enhance natural gas liquid (NGL) production under C3+ rich gas feed stream [24]. Following example illustrates how nitrogen rejective (hydrocarbon selective) rubbery membranes can be utilized for enhancement of the C2+ hydrocarbons recovery via a NGL process in a gas plant. The NGL recovery process is cascaded-refrigeration process with and/or without membrane units, as shown in Figure 5. The gas plant is processing both AG stream (e.g. 450 psi) and NAG stream (e.g. 800 psi), to produce sales gas (SC) and C2+ NGL products.

Figure 5.

Schematic of the NGL recovery process: (A) without membrane; (B) utilizing rubbery membrane to divert NAG stream to feed LRU.

The core of the NGL process is the Liquid Recovery Unit (LRU), which is a cascaded refrigeration process. Figure 6 details further the cascaded-refrigeration unit. Dry natural gas feed is cooled down to −68°C (−90°F) in three cooling stages; C3 refrigeration loop is used in the first and second stages, while C2 refrigeration loop is used in the third stage.

Figure 6.

Schematic of the cascaded-refrigeration process to recover C2+ NGL from natural gas.

As noted in Figure 6, the C2 refrigeration loop is cascaded in the C3 refrigeration loop, hence referred to as cascaded-refrigeration NGL recovery process. The residue gas leaving the third cooling stage is recycled back to exchange heat with gas feeding each stage and leaving the first stage at around 450 psi (referred to as high pressure (HP) residue gas). The liquids collected in the three stages are fed to the demethanizer, which is operating at lower pressure. The C2+ NGL product is obtained as the bottom product, while the top gas product from the demethanizer (low pressure (LP) residue gas) is consolidated with HP residue stream and sent to SG network. This LRU process is ideally designed to process AG stream, which is C2+ rich and free of or very lean in nitrogen. However, under certain situations of higher demand for C2+ NGL, the LRU is designed or operated with feed that is composed partially of NAG. As noted in Table 4, the N2 content of NAG stream is higher, and the C2+ content is considerably lower, than in AG stream.

Flow (MMSCFD)Pressure (psi)%N2%C1%C2%C3%C4%C5+
AG5004800.565.517.910.93.981.25
NAG4008001480.23.561.280.670.25

Table 4.

Conditions of AG and NAG streams.

Table 5 compares the performance of the NGL process when 140 MMSCFD of the NAG stream is diverted to mix with the 500 MMSCFD AG stream. In reference (base case), Table 5 provides the performance of the NGL process when only AG stream is fed to the LRU. When portion of NAG is let down directly to mix with AG, the condensability of the feed stream to LRU is reduced, as indicated by the increase in N2 content and reduction in the C2+ content. Nevertheless, the C3+ NGL liquids production has increased as a direct result of the increase in feed to LRU, while liquid C2 production has decreased in almost similar rate rendering the cascaded-refrigeration NGL recovery process not sensitive to the increase in the flow of C2+ to LRU. On the other hand, diverting part of the NAG selectively through membrane has resulted in increasing the condensability of the feed to LRU. As a result of enhancing the condensability and flow of the feed to LRU, the NGL production has increased, especially valuable C3+ productivity. The impact of increasing the flow to LRU, as noted in Table 5, is to reduce the heating value of the sales gas. This is mainly the result of separating more condensable valuable C3+ hydrocarbons from the sales gas stream. This impact is exacerbated when the diversion of NAG is conducted selectively through the membrane, therefore increasing the shrinkage in produced sales gas.

No mixingDirect let downMembrane1
AG to LRU500500500
NAG diverted to LRU (MMSCFD)0140140
%NAG in feed to LRU021.921.9
Feed to LRU (MMSCFD)500640640
 %N20.53.52.6
 %C165.568.768.8
 %C217.914.715.1
 %C3+16.113.113.5
C2+ NGL production (Barrell/day)92,10091,90095,200
 C2 (Barrel/day)36,90035,10036,600
 C3+ (Barrel/day)55,20056,80058,600
 Extra C2+ (Barrel/day) Vs let downReference2−1800+3300
 Extra C3+ (Barrel/day) Vs let downReference2+1600+1800
Sales gas:
High Heating Value (Btu/SCF)988.1985.7975.6
Sales Gas Shrinkage (MMSCFD)Reference20.714.0

Table 5.

Impact of incorporating part of NAG in the feed to LRU of the cascaded-refrigeration NGL recovery process.

Membrane performance: C1/N2 ∼ 1.6, C2/N2 ∼ 3.2, C3/N2 ∼ 4.8, C4/N2 ∼ 5.2, C5+/N2 ∼ 9.


Base case (only AG stream is fed to the LRU).


Advertisement

3. Gas transport mechanism in polymer membranes

Polymeric membranes are generally non-porous and gas permeation through dense polymer membranes is typically described by the solution-diffusion model. Gas permeation occurs by sorption of gas on the feed side of membranes, then molecular diffusion through the membrane matrix, and evaporation of the gas from permeate side of membrane surface. Based on the solution-diffusion model, the membrane permeability, Pi (Barrer), is defined as the product of diffusivity coefficient and solubility coefficient (given by Eq. (1)) at high fugacity differences across the membrane [40]:

Pi=Di×SiE1

where Pi is the permeability coefficient measured in Barrer (1 Barrer = 1 × 10−10 cm3 (STP) cm/cm2 s cmHg), Si is the solubility coefficient (cm3 gas/cm3 polymer cmHg), and Di is the diffusion coefficient of the penetrant (cm2/s). The solubility coefficient is a thermodynamic parameter and is mainly influenced by the condensability of the penetrant gases [41], and it is inversely proportional to gas boiling point or critical temperature [42]. In contrast to gas sorption, the gas diffusion coefficients vary widely depending on the polymer materials, it is a kinetic parameter that measures the overall mobility of the penetrant molecules in the membrane, depending on various factors such as the size and shape of the gas molecules, the cohesive energy density of the polymer, the mobility of the polymer chains and the free volume size and distribution of the polymer [43]. Accordingly, the membrane selectivity α, the best measure of a membrane’s ability to separate gases i and j is defined as the ratio of the permeabilities of penetrant, Pi and Pj (given by Eq. (2)) [44].

αij=Pi/Pj=Di/Dj×Si/SjE2

As shown by Eq. (2), membrane selectivity (also known as permselectivity) can be separated into diffusivity selectivity, Di/Dj, taken as the ratio of diffusivity coefficients of the two gases and solubility selectivity, Si/Sj, taken as the ratio of solubility coefficients. Diffusivity selectivity depends primarily on the size-sieving ability of the membrane, the permeation of the smaller molecule is faster as compared to their larger counterparts. While solubility selectivity is largely driven by gas condensability and affinity with the membrane.

In general, glassy polymeric membranes generally tend to permeate smaller molecules, the diffusion coefficients decrease as the molecular size increases; whereas rubbery polymeric membranes permeate more condensable gases, the sorption coefficients generally increase as the condensability increases. Such trends are schematically illustrated in Figure 7 for glassy and for rubbery polymeric membranes, which shows the gas permeabilities for principal gas components in natural gas with different molecular size (kinetic diameter, dk, Å) [13] and condensability (critical temperature, Tc, K) [45]. The molecular sizes and relative condensabilities of C3H8 and C4H10, relative to CH4, are highlighted in Figure 7. CH4 can be separated from C3+ hydrocarbons by glassy polymers by size-selectivity; however, interaction of heavier hydrocarbons with the polymers chains tends to increase the flux resulting in decrement in C3+/CH4 selectivity in case of rich hydrocarbon natural gas. On the other hand, rubbery polymers (sorption selectivity membranes) are used to separate C3+ hydrocarbons from CH4, because of their condensabilities. For example, the selectivity of hydrocarbon vapors in PDMS membranes is dominated by the sorption component, the C3H8/CH4 and C3H8/N2 sorption selectivities (Ki/KCH4 and Ki/K) are 50-fold and 370-fold greater than that of their diffusion selectivities (Di/DCH4 and Di/DN2) (Table 6) [16, 46, 47, 48].

GasCondensability (k)permeability Pi (Barrer)1total selectivity αiN2diffusion selectivity Di/DN2sorption selectivity Ki/KN2total selectivity αi/αCH4diffusion selectivity Di/DCH4sorption selectivity Ki/KCH4
H2338902.234.120.560.746.360.12
O215580021.002.000.671.550.43
N212640011.001.000.331.550.21
CH4191120030.654.67111
CO230438009.50.6514.333.1713.07
C2H630533008.250.3324.442.750.515.24
C3H8369410010.250.1555.563.420.2311.90
C4H10242543507.900.2232.40

Table 6.

Summary of permeability, solubility, and diffusivity parameters in PDMS at 35°C [16, 46, 47, 48].

Barrer = 10−10 cm3(STP).cm/cm2. S.cm Hg.


Tested at 25°C


Figure 7.

The relative size (kinetic diameter, dk, Å) and condensability (critical temperature, Tc, K) of the principal components of natural gas.

Advertisement

4. Research and industry needs

The separation and recovery of C3+ hydrocarbons from natural gas using polymeric membrane-based technology have gained increasing attentions in chemical, petrochemical and oil and gas companies compared to conventional separation technologies. The choice of polymeric membrane materials is crucial for separation, and the key membrane performance variables are selectivity, permeability, and long-term durability. Research regarding the development of polymer materials and fabrication processes has been reported in the literature mostly with single or binary gas mixtures short duration gas permeation data; however, their industrial implementation is often hindered due to effects of multicomponent natural gas mixture permeation behavior on membrane physical aging.

Glassy polymer membranes have excellent scalability and continuously increasing being considered for more challenging hydrocarbon separations. The use of glassy polymer membranes with stiff chains, such as polyphenylene oxide (PPO), ethyl cellulose (EC), cellulose acetate (CA), polysulfone (PSF), aromatic polyimides, polymers of intrinsic microporosity (PIMs), disubstituted polyacetylenes and Si-distributed polynorbornenes etc., for separation of olefins and paraffins as well as C2-C4, aromatic, alicyclic and aliphatic hydrocarbons have been described in previous reviews [23, 49]. Unique challenges exist for these glassy polymer membranes for C3+ hydrocarbon separation and removal, most notably selectivity loss and some cases permeability loss due to physical aging and loss of separation efficiency in ultrathin membranes due to faster physical aging. For example, aging-induced permeability loss in glassy polymers is expected to be significant, especially for those condensable gases with larger molecular sizes (Figure 8) [49]. The membrane with ultrathin selective-layers could be the solution to provide economically higher hydrocarbon permeance, however, forming large-scale and defect-free membranes becomes increasing challenging in actual membrane fabrication process. In this regard, a high-flux and a good selectivity for gas separation processes are both required for a reasonable plant size and energy demand [7].

Figure 8.

Aging-induced permeance drop of penetrants in polyimide (6FDA-BPDA-DDBT) membrane aged for 18 months [49].

Further, other challenges that these polymer membrane materials need to withstand their C3+ hydrocarbon separation performances under natural gas feed streams and testing conditions, including a challenging high feed pressure (800+ psi), C3+ rich multicomponent hydrocarbon mixtures along with minor impurities such as CO2, N2, a trace amount of BTEX. PDMS based rubbery siloxane membranes are often used for separating C3+ hydrocarbons from natural gas. Although, many studies have investigated the permeation properties of PDMS membranes under pure gas, binary or ternary gas mixtures in the literature [44, 50, 51, 52, 53, 54, 55, 56], it is important to evaluate them under industrially relevant feed streams and varying operating conditions. Yang et al. [25]. investigated the permeation properties of conventional PDMS and modified siloxane terpolymer (Ter-PDMS) membranes under simulated typical field gas streams. Results shows that C4H10 vol% had significant influence on membrane separation performances due to the C4H10-induced swelling of PDMS based rubbery siloxane membranes. Similarly, Mushardt et al. [57] reported the permeation properties of poly(octylmethylsiloxane) (POMS)/active carbon (AC) mixed matrix membranes (MMMs) under binary and multi-component gas mixtures (Figure 9). The best separation performance was achieved under multicomponent gas mixture at highest average fugacity of C4H10 caused by increasing feed pressure and high concentration of C4H10 in feed gas mixture. These outcomes carry the message that membrane separation performance is strongly related to feed gas compositions and testing conditions. Thus, to fully exploit the use of membranes for actual C3+ hydrocarbons separation and removal, not only higher-flux and cost-effective new membranes are needed, but also it is worthy to pay more attention to the influence of gas compositions and testing conditions.

Figure 9.

Comparison of (a) C4H10 permeance and (b) C4H10/CH4 selectivity of POMS/AC MMMs in multi-component gas mixture (2 vol% C4H10) and C4H10-CH4 binary mixtures (1, 2 or 5 vol% C4H10) at 20°C [57].

Advertisement

5. Design of C3+ hydrocarbon separation polymeric membrane materials

The design and selection of polymeric membrane materials with high permeability and selectivity for gas separation is crucial. The following sections describes the potential research and development activities in the literature of designing C3+ hydrocarbon separation membrane materials. There are two types of polymeric membrane materials reported, which allow to selectively separate and remove С3+ hydrocarbons from natural gas: (1) reverse-selective glassy polymers with high free volume [58], including polyacetylene-based polymers, polymers of intrinsic microporosity (PIMs), polynorbornene-based polymers; (2) rubbery polymers, including PDMS, POMS, modified rubbery siloxanes, polyurethanes, and poly(ether-b-amide) copolymers. Summary of C3+ hydrocarbon permeation properties for these selected high free volume glassy polymers, rubbery polymers and their MMMs under pure and mixed gases can be found in Tables 79.

PolymersPure gasMixed gasRef #
Pressure/TempC3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
Pressure/TempC4H10 mol %C3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
Polyacetylene-based polymers
PTMSP250 psi/23°C78,00015,4005150 psi/25°C253,500180030[59]
250 psi/25°C228,00085,3003100927[60]
1 bar/25°C38,00015,0002.5[16]
150 psi/25°C239,249189121[61]
2.4 atm/35°C3300[62]
4.4 atm/25°C167,40031,0005.44.4 atm/25°C6127,500250051[63]
4.4 atm/35°C156,80028,0005.64.4 atm/35°C683,600220038[63]
1 atm/23°C174017,5009001.9196 atm/23°C3.7260023011.3[64]
7.5590028021.1[12]
PTMSP (L), Toluene1.4 atm/30°C876012,320446022.71.4 atm/30°C228,422214013[65]
PTMSP (L), THF1.4 atm/30°C3870467017202.32.7261207508[65]
PTMSP(2), THF241,569160926[65]
PMP1.3 atm/35°C11,7008501411.3 atm/35°C8600045013[66]
1 bar/25°C730026,00029002.59[66]
PTMGP(85% trans content)3.5 psi/30°C15,00084501.83.5 psi/30°C1.666,190213031.1[67]
3.5 psi/30°C17,00063002.73.5 psi/30°C1.670,900198035.8[67]
PTMSDPA1.4 atm/25°C49,60023,1002.11.4 atm/25°C2221,00014,70015[68]
Poly(2-hexyne)50 psi/35°C35440211.721.050 psi/35°C21590.6[69]
Poly(2-octyne)50 psi/35°C4202100518.241.250 psi/35°C2150562.7[69]
Poly(2-nonyne50 psi/35°C4501850647.028.950 psi/35°C2250554.5[69]
Poly(2-decyne)50 psi/35°C7404600848.854.850 psi/35°C2500776.5[69]
Poly(2-undecyne)50 psi/35°C84039008310.147.050 psi/35°C2600867[69]
Polymers of intrinsic microporosity
PIM-165 psi/35°C550025,10043012.858.4150 psi/25°C2420017524[17]
120 psi/25°C2460032014.4[18]
Polynorbornene-based polymers
APN-SiMe31 atm/23°C174017,5007902.222.2[70, 71, 72]
PBTCN-Si6 atm/23°C26,90033008.16 atm/23°C5.210,30091011.3[64]
PTTCN-Si1 atm/23°C349022,20012502.817.8[73]
MPTTCN-Si1 atm/23°C29019401801.610.8[73]
PTCN-Si1 atm/23°C147013,03010101.512.9[74]
PBTCN-Si1 atm/23°C753026,91033202.38.1[74]
ROMP-SiMe2OEt1 bar/25°C110.21.40.77.4[75]
ROMP-SiMe(OEt)21 bar/25°C12.875.113.315.7[75]
ROMP-Si(OEt)31 bar/25°C241.8210097.72.521.5[75]
APN-SiMe31 bar/25°C856.47041678.91.310.4[75]
APN-SiMe2OEt1 bar/25°C112.7162157.7228.1[75]
APN-SiMe(OEt)21 bar/25°C100.1110449222.6[75]
APN-Si(OEt)31 bar/25°C499.138381762.821.8[75]
Polynorbornene-based polymers
ROMP Si(OEt)3500 psi/25°C3547109213827.9[21]
800 psi/25°C394519761982.110[21]
500 psi/25°C482615191884.38.1[21]
800 psi/25°C4121321602924.27.4[21]
APN-Si(OEt)3500 psi/25°C3316371485455.813.1[21]
800 psi/25°C3431599146826.314.5[21]
500 psi/25°C4446710,0226157.316.3[21]
800 psi/25°C4644815,258811818.8[21]
ROMP TCN-Si(OMe)31 bar/22°C1027110.92.5[76]
ROMP TCN-Si(OEt)31 bar/22 °C1221260502.425.2[76]
ROMP XL TCN-Si(OEt)31 bar/22 °C1161100512.321.6[76]
ROMP XL TCN-Si(OPr)31 bar/22 °C61541001703.624.1[76]
ROMP XL TCN-Si(OBu)31 bar/22 °C130081002505.232.4[76]
APTCN-Si(OMe)31 bar/22 °C33540501302.631.2[76]
APTCN-Si(OEt)31 bar/22 °C51050001533.332.7[76]
APTCN-Si(OPr)31 bar/22 °C4404100845.248.8[76]
APTCN-Si(OBu)31 bar/22°C44036001004.436[76]
APTCN-(OPr)1 bar/25 °C6.641796.4128500 psi/25°C1.520850646.44.510.9[75]
APTCN-(OPr)1 bar/25°C800 psi/25°C1.529373957.65.112.8[75]
APTCN-(SiMe3)1 bar/25°C143.71800841.721500 psi/25°C1.5182849772806.517.8[75]
800 psi/25°C1.5234165233317.119.63[75]
APN-(OPr)3500 psi/25°C1.584.3189243.57.9[75]
APN-(OPr)3800 psi/25°C1.513432232.74.19.8[75]
APN-Si(OEt)3800 psi/25°C1.5776618,2549857.918.5[75]
APN-SiMe3 [Pd]1 bar/25 °C950890012.5[77]
APN-SiMe3 [Ni]1 bar/25 °C180015,20014.2[77]
PVNB1 bar/25 °C149374[78]
Cp-PVNB1 bar/25 °C380448.6[79]
O-PVNB1 bar/25 °C27.57.13.9[79]
S-PVNB1 bar/25 °C4.71.53.1[79]
XL-APTCN1 bar/25 °C145010,80010301.410.5[80]
APTCN-Si-b-APTCNSiOEt (3 mol%)1 bar/22 °C1030179014,6401.714.2[81]
XL-APTCN-Si-b-APTCNSiOEt (3 mol%)1 bar/22 °C960180015,4601.916.1[81]
APTCN-Si-b-APTCNSiOEt (10 mol%)1 bar/22°C830180016,7002.220.1[81]
XL-APTCN-Si-b-APTCNSiOEt (10 mol%)1 bar/22°C77013601.8[81]
APTCN-SiMe31 bar/22 °C188014,73011301.713[82]
APTCN-Si(OEt)31 bar/22 °C15351050003.332.7[82]

Table 7.

C3+ hydrocarbon properties in high free volume glassy polymeric membranes.

PolymersPure gasMixed gasRef. #
Pressure/tempC3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
Pressure/tempC4H10 mol %C3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
Polydimethylsiloxane (PDMS)
PDMS0.01 bar/25°C410012003.4[22]
250 psi/23°C2720012006[59]
240 psi/35°C410012003.4[16]
4.4 atm/35°C740012985.7150 psi/25°C212,900125010.3[83]
11 atm/25°C112,400124010[83]
11 atm/25°C212,900125010.3[83]
11 atm/25°C413,700128010.7[83]
11 atm/25°C616,100140011.5[83]
11 atm/25°C817,400145012[83]
98.6 psi/35°C10,00013507.4[23]
4.4 atm/35°C6400[62]
4.4 atm/25°C17,800120014.84.4 atm/25°C617,600125014.1[52]
4.4 atm/35°C15,900135011.84.4 atm/35°C616,100140011.5[52]
4.4 atm/5 °C13,30015008.904.4 atm/50°C613,50015508.7[52]
5 bar/35°C17,000150011.3[53]
5 atm/35°C10009011[54]
1 bar/35°C550016003.4[84]
2 bar/35°C10,00016006.3[85]
15 bar/35°C4100750012006.25[86]
PDA@PDMS6 bar/20°C160 GPU[87]
PDMS200 psi/25°C3302562794976.112.6[21]
500 psi/25°C347229709672714.4[21]
800 psi/25°C3604112,5398746.914.3[21]
200 psi/25°C4356766685816.111.5[21]
500 psi/25°C4519895338496.111.2[21]
800 psi/25°C4618411,08410825.710.2[21]
PDMS w/net0.8 bar/25°C560018,20013001414[56]
PDMS0.8 bar/25°C520017,000120014.214.2[57]
800 psi/25°C4789115,46810977.116.0[75]
2 bar/20°C690033,800117028.928.9800 psi/25°C0.54260738010334.17.1[25]
800 psi/25°C1.5597111,71011735.110.0[25]
800 psi/25°C411,01621,49616366.713.1[25]
Side-chain substituted PDMS
POMS6 bar/35 °C3510.5110[14]
POMS0.01 bar/25°C2029314[22]
POMS0.8 bar/25°C198075003006.625[56]
POMS0.4 bar/30°C15602705.8[88]
POMS0.5 bar/20°C7.510.2115.610 bar/20°C5710.31122.5[57]
POMS30 bar/20°C513.210.49127[57]
POMS40 bar/30°C59.810.56117.5[89]
POMS98.6 psi/35°C23003107.4[23]
PHexMS0.8 bar/25°C139281603405.824[56]
PDecMS0.8 bar/25°C160864802406.727[56]
PEtMS98.6 psi/35°C38004708.1[23]
PPrMS98.6 psi/35°C42005707.4[23]
PEtF3MS98.6 psi/35°C4502002.3[23]
PPhMS98.6 psi/35°C170364.7[23]
PPrMS0.01 bar/25°C812014505.6[22]
PDMSM0.4 atm/25°C83728071306.4421.6[90]
PDMSTM0.4 atm/25°C372408.44.428.6[90]
PDMSM/PDMSTM (75/25)0.4 atm/25°C54723401134.820.7[90]
Main-chain substituted PDMS
PDMS-PEtMS98.6 psi/35 °C42006007.0[23]
PDMS-PHxMS98.6 psi/35 °C34004008.5[23]
PDMS-POMS98.6 psi/35 °C28003607.8[23]
PDMS-PmPhMS98.6 psi/35 °C8001107.3[23]
PDMS-PpPMS98.6 psi/35 °C401.233.3[23]
Ter-PDMS2 bar/20 °C395021,2004309.1949.3800 psi/25°C0.5278157034695.912.2[25]
Ter-PDMS800 psi/25°C1.5361581855057.216.2[25]
Ter-PDMS800 psi/25°C4919919,52410738.618.2[25]
Polyurethane (PU)
PTMG–HDI–BDO2 bar/30°C18236.75[91]
PCL–HDI–BDO2 bar/30°C61.412.84.8[91]
PPG–HDI–BDO2 bar/30°C20044.14.5[91]
PTMG–TDI–BDA2 bar/30°C19034.85.5[91]
PDMS-PU (50/50)1 bar/30°C3006–8[92]
Poly(ether-b-amide) copolymers
Pebax 1074800 psi/25°C1.54610595.212.0[93]
Pebax 1657800 psi/25°C1.5296764.710.7[93]
Pebax 2533800 psi/25°C1.5340894516.717.6[93]
Pebax 3000800 psi/25°C1.564.5156.813.24.912.0[93]
Pebax 5513800 psi/25°C1.517.241.34.34.19.8[93]

Table 8.

C3+ hydrocarbon properties in rubbery polymeric membranes.

mN3/(m2. h. bar).


PolymersPure GasMixed GasRef. #
FillersLoading %Pressure/tempC3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
Pressure/tempC4H10 mol %C3H8 (Barrer)C4H10 (Barrer)CH4 (Barrer)α
C3H8/CH4
α
C4H10/CH4
PTMSPEH-5 silica30150 psi/25°C254,288232623[61]
TS-610 silica50150 psi/25°C2114,655498023[61]
Silica(TEOS/OMDEOS1:0)6.21.4 atm/30°C189023904704.05.11 atm/30°C224203506.8[65]
Silica(TEOS/OMDEOS1:1)5.21.4 atm/30°C3909001802.25.01 atm/30°C27901107.4[65]
Silica(TEOS/OMDEOS1:2)2.81.4 atm/30°C44011903501.33.41 atm/30°C214401907.6[65]
Silica(TEOS/OMDEOS2:1)2.21.4 atm/30°C140421500.90.31 atm/30°C2470905.4[65]
Silica(TEOS/MTEOS1:1)3.51 atm/30°C1.4 atm/30°C2628049512.7[65]
Silica(TEOS/OTEOS1:1)6.41 atm/30°C4802409601.84.01.4 atm/30°C2385028013.7[65]
Silica(TEOS/DTEOS1:1)11 atm/30°C52030013501.74.51.4 atm/30°C212801608[65]
Silica(TEOS/HDTMOS1:1)0.61 atm/30°C1.4 atm/30°C211101407.9[65]
PTMGP(85% trans content)TiO2103.5 psi/30°C14,30074901.91 bar/30°C1.656,280184030.6[67]
TiO253.5 psi/30°C17,70082402.11 bar/30°C1.636,350121030[67]
TiO2103.5 psi/30°C18,30081202.31 bar/30°C1.666,890213031.4[67]
TiO2203.5 psi/30°C17,00079902.11 bar/30°C1.662,280199031.3[67]
PUZMS-552 atm/30°C92.531.82.91[94]
ZMS-5102 atm/30°C10132.43.12[94]
ZMS-5152 atm/30°C11033.13.33[94]
ZMS-5202 atm/30°C11732.33.63[94]
PDMSSilica0.52 bar/35°C950013507.0[85]
Silica22 bar/35°C30,000105028.6[85]
Silica0.57 bar/35°C22,500135016.7[85]
Silica27 bar/35°C24,50095025.8[85]
fumed Silica5.51.5 atm/30°C75003602110 atm/30°C329624556.5[55]
fumed Silica111.5 atm/30°C10,5006201710 atm/30°C346657526.2[55]
fumed Silica16.51.5 atm/30°C950057016.510 atm/30°C344483706.4[55]
POMSactivated carbon1040 bar/20°C54.910.28117.5[89]
activated carbon2040 bar/20°C53.210.15121[89]
activated carbon4040 bar/20°C510.06118[89]
activated carbon2040 bar/20°C5410.2120[89]
activated carbon200.5 bar/20°C4.210.12114.630 bar/20°C13.810.17122[57]
activated carbon2030 bar/20°C24.510.19124[57]
activated carbon2030 bar/20°C5910.28132.5[57]
activated carbon30 bar/20°C221310.41131.5[57]
Ter-PDMSOS-VTM-POSS501 bar/25°C162774802177.534.6800 psi/25°C1.53283967304017.116.8[95]
OS-VTM-POSS50800 psi/25°C0.54202548812627.718.7[95]

Table 9.

C3+ hydrocarbon properties in mixed matrix membranes.

mN3/(m2. h. bar).


Multicomponent gas mixture: 1% C5H12/2% C4H10/6% C3H8/10% C2H6/79% CH4/2% CO2.


Multicomponent gas mixture: 1.5%C4H10/3% C3H8/5% C2H6/66.5%% CH4/12% CO2/12% N2.


Multicomponent gas mixture: 0.5%C4H10/2% C3H8/5% C2H6/77.9%% CH4/14% N2/500ppm BTEX.


5.1 Reverse-selective glassy polymers

5.1.1 Polyacetylene-based polymers

Exceptional gas transport properties of polyacetylene-based polymers have led to considerable interest for membrane-based gas separation applications (Figure 10). These highly rigid, amorphous glassy polymers exhibit high Tg (>200°C) and high free volumes, providing ultrahigh permeabilities for C3+ hydrocarbons and attractive selectivities of C3+/CH4 (Table 7).

Figure 10.

Chemical structures of selected branched polyacetylene-based polymers (i.e. PTMSP, PMP, PTMGP, PTMSPA) and linear poly(2-alkylacetylenes) (i.e. i.e., poly(2-hexyne), poly(2-octyne), poly(2-nonyne), poly(2-decyne), and poly-(2-undecyne)).

Poly(1-trimethylsilyl-1-propyne) (PTMSP) was synthesized for the first time in 1983 [96]. Unlike conventional glassy polymers, PTMSP is the most permeable glassy polymer known for C3+ hydrocarbon separation and removal with a pure gas C4H10 permeability >160,000 Barrer, but a low pure gas C4H10/CH4 selectivity (<6) (Table 7). However, C4H10/CH4 selectivity was found to be higher under mixed gas permeation testing, indicating PTMSP may be an alternative to siloxane rubbery polymers for C3+ hydrocarbon separation. For example, Pinnau et al. [59] reported that mixed gas selectivities of C4H10 over permanent gases (CH4 and H2) are as high as 27 for C4H10/CH4 and 39 for C4H10/H2 because CH4 and H2 permeabilities in gas mixtures containing 2 vol% C4H10 are much lower than the pure CH4 and H2 permeabilities. Raharjo et al. [63] reported the mixed gas C4H10/CH4 selectivity was >8 times higher than that of pure gas measurement, ultrahigh mixed gas C4H10 permeability (127,000 Barrer) and C4H10/CH4 mixed gas selectivity (>50) were obtained under binary gas mixtures containing 6 vol% C4H10. Similar trends were observed for other polyacetylene-based polymers, such as poly(4-methyl-2-pentyne) (PMP) [64], poly(1-trimethylgermyl-1-propyne) (PTMGP) [64] and poly[1-phenyl-2-(p-trimethylsilylphenyl) acetylene] (PTMSDPA) [66]. Unlike branched polyacetylenes (i.e. PTMSP and PMP) with bulky side substitutes and higher fractional free volume (FFV = 0.28–0.29), linear-based poly(2-alkylacetylenes) with longer alkyl side chains, i.e., poly(2-hexyne), poly(2-octyne), poly(2-nonyne), poly(2-decyne), and poly- (2-undecyne), have relatively lower fractional free volume (FFV = 0.19–0.22) and permeabilities due to increased polymer chain mobility and interchain interactions [49, 67], and these linear poly(2-alkylacetylenes) membranes exhibit rather inferior performance for C4H10/CH4 mixed gas separation (Figure 11) under mixed gas testing conditions. However, the use of PTMSP as a potential membrane polymer for C3+ hydrocarbon recovery is not feasible due to fast physical aging (nonequilibrium excess free-volume relaxation of PTMSP chains) and solubility (e.g. interactions of PTMSP chains with many compounds), which can result in potential membrane destruction in the gas process streams.

Figure 11.

Mixed gas C4H10 permeability and C4H10/CH4 selectivity for selected branched polyacetylene (i.e. PTMSP, PMP, PTMGP and PTMSDPA) and linear poly(2-alkylacetylenes) (i.e. poly(2-hexyne), poly(2-octyne), poly(2-nonyne), poly(2-decyne), and poly-(2-undecyne)) membranes.

5.1.2 PIM-1

PIMs are spirobisindane-based rigid, glassy polymers with high fractional free volume (FFV = 0.26), excellent chemical stability and good solvent resistance. PIM-1 is the most studied PIMs for gas separation, and it shows high CO2 permeability with moderate CO2/CH4 selectivity, which is exceeding the upper bounds [97]. A new perspective on the role of PIMs and its functionalized ladder PIMs (e.g. PIM-polyimide copolymers) in key energy-intensive membrane-based gas separations, including O2/N2, H2/N2, H2/CH4, CO2/CH4, H2S/CH4, C2H4/C2H6, and C3H6/C3H8, was described in detail in a recent review paper by Wang et al. [98]. Similar as PTMSP, PIM-1 is also particularly permeable to hydrocarbon vapors (Table 7). Thomas et al. [16, 17] investigated the pure- and mixed-gas permeation properties of PIM-1, it exhibits C4H10/CH4 and C4H10/H2 mixed gas selectivities up to 25 and 27, respectively, under a mixture of 2 vol% C4H10 in CH4. Since PIM-1 has better solvent resistance than PTMSP, it could find applications as an advanced membrane material for the separation of organic vapor/permanent gas mixtures [17], and further membrane performance evaluation under industrially relevant feed streams and testing conditions will be the topic of future studies in the research groups.

  1. Polynorbornene-based polymers

Another promising reverse-selective glassy polymer class for C3+ hydrocarbon separation and removal are substituted polynorbornenes (PNBs). Many valuable correlations between gas-permeability and polynorbornene structure have been summarized in prior reviews [12, 19, 71]. More recently, the development in the field of next generation polynorbornene-based polymers, their limitations and challenges for targeted gas separations was discussed also in a recent review [99]. These studies show that solubility-controlled permeation of gaseous hydrocarbons (C1 − C4) is characteristic for both addition polymerization-type polynorbornenes (APNBs) with Si-containing side-groups and ring opening metathesis polymerization type polynorbornenes (ROMP PNBs) with flexible Si-O bonds in side-groups (Figure 12). Summary of C3+ hydrocarbon permeation properties for APNBs and ROMP PNBs membranes under pure- and mixed-gas can be found in Table 7. In general, the membrane C4H10/CH4 selectivity increased significantly as increasing the length of side groups (e.g. Me, Et, n-Pr, n-Bu). Among these polynorbornene-based polymers, both distributed APNBs-Si(OPr)3 and ROMP-Si(OBu)3 membranes exhibited higher C4H10/CH4 pure gas selectivities of 48.8 and 32.4, respectively [76].

Figure 12.

Structures of metathesis and addition polynorbornenes bearing different substituents for gas separation applications.

Although many studies investigated the permeation properties using pure gases, fewer reported separation performance under multicomponent gas feed streams [21]. Sundell et al. [75] reported a route toward the production of exo ROMP and addition-type polynorbornenes and polytricyclononenes (APTCN) through the stereochemical control afforded by the reductive Mizoroki−Heck reaction. All addition-type and tricyclononene-based polymer membranes (e.g. APTCN-SiMe3 and APTCN-Si(OEt)3) show improved mixed gas C4H10/CH4 selectivities compared to APBNs-Si(OEt)3 and ROMP-Si(OEt)3 membranes (Figure 13) under the same testing conditions, due to the increased polymeric chain-spacing. These polynorbornene-based polymers demonstrate solubility-selective permeation with mixed gas selectivities that exceed commercially used PDMS.

Figure 13.

Mixed gas C4H10 permeability vs.C4H10/CH4 selectivity for APBNs-Si(OEt)3 (black squares), ROMP-Si(OEt)3 (red circles) and substituted-APTCN (blue triangles) under C3+ rich multicomponent gas mixtures containing 1.5–4 vol% C4H10 at 25°C and 800 psi.

5.2 Rubbery polymers

5.2.1 Rubbery siloxane-containing homo-and copolymers

Application of PDMS based rubbery siloxane membranes for selective removal of hydrocarbons and recovery of various organic components [3, 100, 101, 102] is considered. Summary of C3+ hydrocarbon permeation properties of PDMS, POMS and modified rubbery siloxane membranes in pure and mixed gases can be found in Table 8.

PDMS has low Tg (∼−123°C), flexible chains, and large free volume. The permeability, solubility, and diffusivity of hydrocarbons in PDMS membranes have been studied in previous literature [52, 83], and results show that feed gas compositions and testing conditions significantly affect membrane separation performance. The effect of chemical compositions on the permeation properties of PDMS membrane by introducing substituted groups in both side and main chains was compressively studied by Stern et al. [90, 103, 104, 105] and summarized in previous reviews [12, 23]. Results show that the substitution of functional groups in the side-groups of siloxane polymers has the same general effect on polymer permeability and selectivity as the substitution of such groups in the polymer main chains [104] (Figure 14). When the size of the side-groups (e.g. –CH3, –C2H5, –C3H7, –C8H17, − C10H11–, CH2CH2CF3, –C6H5, etc.) and main chains (e.g. –(CH2)2, −(CH2) 2, −(CH2)8, −m-C6H4, −p-C6H24) increased, the chains become less flexible resulting in the increases in Tg (from −123°C to −28°C) and the decreases in permeability, especially for those condensable gases with larger kinetic diameters (e.g. C3H8) [23].

Figure 14.

Chemical structures of PDMS-based siloxane polymers with different side and main chains.

Compared to PDMS with (Me2SiO)x backbone chain, membranes based on POMS with (OctMeSiO)x backbone chain exhibited enhanced C3+ hydrocarbon separation performance (C3+/CH4 selectivities) [14, 56, 57, 88, 89]. Schultz et al. [14] reported that POMS was one of two promising polymers for separation of C3+ hydrocarbons among the total 45 different polymers that were tested under C4H10/CH4 binary gas mixture (97/3 mol%). Since POMS is a soft material, it is susceptible to compression and thus the permeability goes down significantly with increase in the feed pressure as shown in Figure 15.

Figure 15.

Pressure behavior of the POMS under C4H10/CH4 (97/3 Mol%) binary gas mixture, where compression decreases flux with feed pressure ⋄ methane, • butane, x selectivity [14].

More recently, Grushevenko et al. [56] synthesized siloxane rubbery membranes by using an in-situ technique with the introduction of side alkyl groups followed by polymer chains crosslinking in the same reaction mixture. Results showed that POMS membrane demonstrated higher C4H10/CH4 ideal selectivity (α = 25) than that of PDMS membrane (α = 14). The gas permeation properties of polyhexylmethylsiloxane (PHexMS) and polydecylmethylsiloxane (PDecMS) were further studied. With the increase of substituent length from C8 to C10, C4H10 permeability (6480 Barrer) of PDecMS membranes decreased but C4H10/CH4 selectivity increased to 27 due to reducing diffusion coefficient. These results show that polyalkylmethylsiloxanes with longer side chains (e.g. PDecMS) are promising to be applied as materials for the purpose of C3+ extraction from natural gas (Figure 16).

Figure 16.

Permeability (1-CH4, 2-C3H8, 3-C4H10, 4-CO2, 5-N2) and C4H10/CH4 selectivity of polyalkylmethylsiloxane membranes (feed pressure:0.8 bar, temperature: 25°C) [56].

In addition, a novel class of silicon-organic polymers, silmethylene rubbery polymers including more robust Si-C bonds, has been developed for separation of C3+ hydrocarbons from natural gas. Alentiev et al. [90] reported a thermally initiated polymerization technique to synthesize polydimethylsilmethylene (PDMSM) and polydimethylsiltrimethylene (PDMSTM) homo-polymers that include only more robust Si–C bonds (Figure 17). These silicon-containing rubbery polymers and its copolymers have relatively high Tg (−92°C to −76°C) and excellent chemical stability (decomposition in air >240°C), compared to conventional PDMS. In addition, these novel rubbery polymers distinctly reveal solubility-controlled permeation behavior, but the permeabilities of C3+ hydrocarbons in PDMSM and PDMSTM are lower than those in PDMS.

Figure 17.

The scheme of synthesis of silmethylene rubbery polymers: polydimethylsilmethylene (PDMSM) and polydimethylsiltrimethylene (PDMSTM).

A comprehensive survey of the recent developments on the synthesis, properties, and applications of silicon-containing copolymers was provided in the literature [106], including silicone-urea and silicone-urethane copolymers, silicone-ester copolymers, silicone-amide copolymers, silicone-imide copolymers etc. These silicon-containing copolymers which combine unique properties of PDMS in the polymer backbone have been evaluated as gas separation membranes, such as CO2/N2 and CO2/CH4 separations [92, 107, 108, 109, 110], but few reported C3+ hydrocarbon separation and removal from natural gas. Gomesa et al. [92] reported the synthesis of poly (ether siloxane urethane urea) membrane materials for C4H10/CH4 separation. Results show that the higher the content of soft siloxane segments, the higher are the permeabilities of C4H10 and CH4 due to higher chain mobility, while the C4H10/CH4 single gas selectivity (6–8) does not change so much as permeability (2–320 Barrer).

A series of silicon-containing terpolymers with Si-O-Si and Si-(CH2)2- in polymer backbone and side chains were synthesized for enhanced C3+ hydrocarbon recovery from natural gas by Yang et al. [95]. The partially octyl substituted crosslinked rubbery siloxane membranes were prepared by crosslinking vinylmethylsiloxane terpolymer (Ter-PDMS) composed of (VinylMeSiO)p(Me2SiO)m(OctMeSiO)n backbone chains via an addition curing process [25]. Improved gas permeation performances were observed for this modified siloxane Ter-PDMS membrane with C4H10/CH4 ideal selectivity of 49.1 (2 bar and 20°C), with lower C4H10 permeability (19,800 Barrer) compared to conventional PDMS membrane (33,800 Barrer C4H10 permeability). Such decrease in permeability compared to conventional PDMS was induced by a reduction of chain flexibility with partial substitution of functional groups in the side chains after crosslinking in membrane matrix. Moreover, this modified siloxane Ter-PDMS membrane also showed enhanced C3+ separation performance (143% increase in C4H10/CH4 mixed gas selectivity) under high pressure (up to 800 psi) 7-component gas mixtures compared to conventional PDMS membrane [25].

5.2.2 Polyurethane rubbery polymers

Polyurethanes have found commercial applications using in a diversity of products in industry, from coatings and adhesives for automotive, shoe soles, mattresses, and foam insulation. Polyurethanes are rubbery polymers which are introduced for different gas pair separations vastly, such as CO2/N2, CO2/CH4, O2/N2. The solubility and diffusivity of gases indicate solubility domination of gas transport in these membranes. Varies researchers have shown that the permeability of polyurethane membranes increases, while the selectivity decreases with the decrease the hard segment content (urethane or urea), or the increase the soft segment (polyether or polyester) molecular weight [111, 112, 113, 114, 115].

Poly(urethane–urea) (PUU) membranes display moderate C3+/CH4 selectivity for removal of C3+ hydrocarbons from natural gas. Sadeghi and coworkers [91] evaluated the separation performance of a series of PUU membranes synthesized with different type of diisocyanates and polyols. It was found that PUU membranes with polypropylene glycol (PPG) polyol had the maximum C3H8 permeability of 200 Barrer and C3H8/CH4 selectivity of 5.5 because of its higher rubbery property and higher phase separation. In their recent paper [116], two series of PUU membranes containing aromatic and aliphatic side chains with different sizes were developed for separating of C2H6 and C3H8 from CH4. The longer aliphatic side chain based PUU provides the best hydrocarbon separation performance with C3H8 permeability of 186 Barrer and C3H8/CH4 idea selectivity of 6.5.

5.2.3 Poly(ether-b-amide) rubbery copolymers (Pebax®)

Pebax®—product of Arkema, is a group of thermoplastic elastomers, containing polyamide “hard” segment (e.g., nylon-6, nylon-12), and polyether “soft” segment (e.g., poly (ethylene oxide) [PEO], poly (tetramethylene oxide) [PTMEO]), which maintain the highly desirable combination of the toughness traditionally associated with polyamides and the flexibility/elasticity with polyether. Various Pebax® copolymers with different polyether/polyamide ratios were designed and synthesized to meet the requirements for different applications. Typical grades of commercial Pebax® 2533, 3533, 4011, 4033, 5533, 6633, 1074 and 1657 have been widely studied in literature for gas separation applications. For example, Pebax® 1657 containing 60 wt% polyether segment is a promising material showing good performance in CO2 separation from CH4 or N2 [117]. MTR has also demonstrated favorable economics for a hybrid separation process that utilized Pebax® 4011 in conjunction with an amine absorption polishing step for sour gas separation [118]. However, fewer reported C3+ hydrocarbon separation and removal from natural gas. Harrigan et al. [93] recently investigated the sour gas separation performance of five different grades of commercial Pebax® 2533, 5513, 3000, 1074 and 1657 under high pressure 6-component gas mixtures containing C2+ hydrocarbons. Results show that Pebax® 1074 and 1657 containing low of polyether segment (50–60 wt%) retained the highest CO2/CH4 selectivity (α = 16–19) but lower C3+/CH4 selectivities (C3H8/CH4 = 4.7–5.2; C4H10/CH4 = 11–12). Surprisingly Pebax® 2533 containing higher polyether segment (80 wt%) exhibited higher C3H8/CH4 selectivity of 6.7 and C4H10/CH4 selectivity of 17.6), which are greater than the C3H8/CH4 selectivity of 5.3 and C4H10/CH4 selectivity of 10.5 for commercially available PDMS membrane under the same testing conditions (800 psi, 25°C) [25]. However, Pebax® 2533 membrane shows 93% and 91% decrease in C3H8 and C4H10 permeabilities, respectively, compared with that of PDMS.

5.3 Mixed matrix membranes (MMMs)

MMMs have been explored to develop materials for enhanced separation performance of conventional membrane. Different types of inorganic fillers, such as silica, TiO2, fumed silica, graphene oxide (GO), activated carbon (AC), MOFs, zeolites, ZIFs, POSS etc., have been incorporated into PDMS, POMS, PTMST, and PU membrane matrix for air separation [119], CO2 removal and capture [120, 121], pervaporation [122], hydrogen separation [83, 123], and hydrocarbon separation [10, 65, 94, 124, 125, 126, 127, 128, 129]. The concept of using MMMs for light and heavy hydrocarbons separation and recovery in petrochemical and chemicals industries has attracted much interest and research in recent years. Several studies, such as olefin/paraffin separations (e.g. C2H4/C2H6, C3H6/C3H8, n-C4H8/i-C4H10), aromatic/aliphatic separations (e.g. toluene/n-hexane, p-xylene/o-xylene), vapor/gas separations (e.g. C3+/N2, C3+/H2 and C3+/CH4) were discussed in a recent review by Najari and coworkers [125]. Here, summary of C3+ hydrocarbon permeation properties of mixed matrix membranes (e.g. PTMSP, PTMGP, PDMS, POMS, PU and modified rubbery siloxane) in pure and mixed gases can be found in Table 9. Khanbabaei et al. [55] prepared PDMS-fumed silica MMMs for removing C4H10 from CH4. Results show the decreased C4H10/CH4 mixed gas selectivity for both neat membrane and MMMs. The MMMs exhibited 38% increment in C4H10 permeability and simultaneous 30% increase in C4H10/CH4 selectivity under C4H10/CH4 (3/97 mol%) binary gas mixture with adding >10 wt% of fumed silica, compared to neat PDMS membrane. Mushardt et al. [89] evaluated the separation of C3+ hydrocarbons of MMMs composed of POMS and 10–40 wt% AC from permanent gas streams under varying operating conditions (e.g. feed and permeate pressure, temperature, feed gas compositions). Best performance was achieved at highest average fugacity of C4H10 caused by increasing feed pressure, high permeate pressure and high concentration of C4H10 in feed mixture. At 40 bar, POMS/20 wt% AC MMMs exhibited higher C4H10/CH4 mixed gas selectivity of 33 and C4H10 permeance of 18 mN3/(m2.h.bar) under multicomponent gas mixtures including 2 vol% of C4H10. Gomes et al. [65] reported the preparation of PTMSP/silica MMMs with incorporation of 20–40 nm silica by sol–gel copolymerization of tetraethoxysilane (TEOS) with different organoalkoxysilanes. When methyltriethoxysilane (MTEOS) or n-octyltriethoxysilane (OTEOS) were used, the obtained MMMs exhibited higher separation performance (C4H10/CH4 mixed gas selectivities of 12.7 and 13.9 at 1 bar and 30°C, respectively).

In a recent investigation, Yang et al. [95] described a method to design and produce novel, crosslinked siloxane/POSS MMMs for enhanced C3+ hydrocarbon recovery from natural gas. Dual-functional POSS nanofiller (OS-POSS-VTMO) containing silicon hydride moiety (☰–Si–H) and trimethoxylsilicon groups (–Si(OMe)3) was synthesized and used as both a crosslinking agent and nanofiller (Figure 18) in modified rubbery siloxane (Ter-PDMS) membrane matrix. Under simulated typical field gas streams and testing conditions (e.g. feed pressure up to 800 psi for multicomponent gas mixture consisting C1-C5 hydrocarbons, CO2, N2 and BTEX). The produced novel, crosslinked Ter-PDMS/POSS MMMs exhibited better separation performance (e.g. 89% and 163% increase in C3H8/CH4 and C4H10/CH4 mixed gas selectivities) and enhanced swelling resistance, compared to conventional PDMS membrane. At 800 psi, Ter-PDMS/POSS MMM had a mixed gas C4H10 permeability of 4880 Barrer and C4H10/CH4 mixed gas selectivity of 18.7, respectively.

Figure 18.

Molecular structure of dual-functionalized POSS crosslinking agent (OS-POSS-VTMO) containing silicon hydride moiety (☰–Si–H) and trimethoxylsilicon groups –(Si(OMe)3).

Advertisement

6. Conclusions and outlook

Membrane-based separations have emerged as an economically favorable alternative due to its small footprint, reduced capital, and energetic cost. Membranes for C3+ hydrocarbon recovery from natural gas have been practiced in the chemical and petrochemical industries, the improved C3+ hydrocarbon separation and removal process from natural gas will be beneficial of these industries.

This chapter summarizes C3+ hydrocarbon separation and removal from natural gas using polymeric membrane-based technologies. More specifically, it addresses the removal of C3H8 and C4H10 from CH4 by perspective polymer architectures based on reverse-selective glassy polymers with stiff chains and rubbery polymers with flexible chains. At present PDMS, POMS and poly(ether-b-amide) rubbery polymer membranes are used by industries in the field for C3+/CH4 separation. However, improved mixed gas selectivity of C3+/C1 with high flux will have a positive impact of the economics of the C3+ hydrocarbon recovery from natural gas. In case of reverse selective glassy polymers, membrane separation performance can be improved by introducing new groups or additives into membrane matrix, however, the loss of permeances due to physical aging restrains their applications in industry. This chapter also accounts the effect of the chemical structure of rubbery polymers on their permeation properties for C3+ hydrocarbons recovery from natural gas. Modified siloxane rubbery polymers-based membranes may be seen in future in the field for enhanced C3+ hydrocarbons recovery from natural gas since these membrane materials display improved mixed gas C3+/CH4 selectivity than commercial membranes. In addition, only few studies regarding MMMs approach for enhanced C3+ heavy hydrocarbons removal from natural gas are reported in the literature. More research efforts are needed in the future from the selection of right nanofillers and understanding the dynamic interactions between polymers and nanofillers regarding mixed gas permeation in the membrane matrix.

The permeation properties of the polymeric and mixed polymer matrix membrane materials have been reviewed and compared under varying operating parameters by many research groups. These results emphasize the observation that C3+ hydrocarbon separation performance is strongly related to the testing condition, feed compositions and chemical structures. However, a thorough evaluation of C3+ hydrocarbon separation properties is rarely reported in the literature under industrially relevant feed streams and testing conditions. Thus, there is a need for further research and testing of advanced membrane materials that are resistance to swelling and satisfy the practical requirements during the actual industrial gas processes.

Advertisement

Acknowledgments

The support provided by the team from Oil and Gas Treatment Division, Research and Development Center (RDC) of Saudi Aramco is greatly acknowledged. The authors also would like to thank Ahmad Bahamdan and Faisal Otaibi from RDC for their supports.

References

  1. 1. The BP Statistical Review of World Energy. 70th ed. 2021. Available from: https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2021-full-report.pdf
  2. 2. Statistics report “Key World Energy Statistics 2021”, IEA September 2021. Available from: https://iea.blob.core.windows.net/assets/52f66a88-0b63-4ad2-94a5-29d36e864b82/KeyWorldEnergyStatistics2021.pdf
  3. 3. Baker RW. Membrane Technology and Applications. 2nd ed. New York: John Wiley and Sons; 2004
  4. 4. Bernardo P, Drioli E, Golemme G. Membrane gas separation: A review/state of the art. Industrial and Engineering Chemistry Research. 2009;48:4638-4663
  5. 5. Baker RW. Membrane gas-separation: Applications. Membrane Operations: Innovation Separation and Transformations. Wiley-VCH Verlag GmbH Co; 2009;8:167-194
  6. 6. Gas Separation Membranes—Global Market Trajectory and Analytics. Global Industry Analyst Inc; 2021. Available from: https://www.businesswire.com/news/home/20220120005591/en/Global-Gas-Separation-Membranes-Market-Trajectory-Analytics-Report-2021-Asia-Pacific-Poised-to-Drive-Market-Growth-Europe-and-US-Hold-Significant-Share—Forecast-to-2026—ResearchAndMarkets.com
  7. 7. Baker RW. Future directions of membrane gas separation technology. Industrial and Engineering Chemistry Research. 2002;41:1393-1411
  8. 8. Ismail AF, Ridzuan N, Rahman SA. Latest development on the membrane formation for gas separation. Songklanakarin Journal of Science and Technology. 2002;24:1025-1042
  9. 9. Sanders DF, Smith ZP, Guo R, Robeson LM, McGrath JE, Paul DR, et al. Energy-efficient polymeric gas separation membranes for a sustainable future: A review. Polymer. 2013;54:4729-4761
  10. 10. Ravanchi RM, Kaghazchi T, Kargari A. Application of membrane separation processes in petrochemical industry: A review. Desalination. 2009;235:199-244
  11. 11. Scholes CS, Stevens GW, Kentish SE. Membrane gas separation applications in natural gas processing. Fuel. 2012;96:15-28
  12. 12. Yampolskii Y. Polymeric gas separation membranes. Macromolecules. 2012;45:3298-3311
  13. 13. Baker RW, Lokhandwala K. Natural gas processing with membranes: An overview. Industrial and Engineering Chemistry Research. 2008;47:2109-2121
  14. 14. Schultz J, Peinemann KV. Membranes for separation of higher hydrocarbons from methane. Journal of Membrane Science. 1996;110:37-45
  15. 15. Toy LG. Pinnau I. Natural Gas Treatment Process Using PTMSP Membrane. U.S. Patent 55017221
  16. 16. Merkel TC, Bondar V, Nagai K, Freeman BD. Gas sorption, diffusion, and permeation in poly(dimethylsiloxane). Journal of Polymer Science Part B: Polymer Physics. 2000;38:273-296
  17. 17. Thomas S, Pinnau I, Dub N, Guiver MD. Pure- and mixed-gas permeation properties of a microporous spirobisindane-based ladder polymer (PIM-1). Journal of Membrane Science. 2009;333:125-131
  18. 18. Thomas S, Pinnau I, Du N, Guiver MD. Hydrocarbon/hydrogen mixed-gas permeation properties of PIM-1, an amorphous microporous spirobisindane polymer. Journal of Membrane Science. 2009;338:1-4
  19. 19. Yampolskii Y, Starannikova L, Belov N, Bermeshev M, Gringolts M, Finkelshtein E. Solubility controlled permeation of hydrocarbons: New membrane materials and results. Journal of Membrane Science. 2014;453:532-545
  20. 20. Jue ML, Lively RP. Targeted gas separations through polymer membrane functionalization. Reactive and Functional Polymers. 2015;86:88-110
  21. 21. Vaughn JT, Harrigan DJ, Sundell BJ, Lawrence JA, Yang J. Reverse selective glassy polymers for C3+ hydrocarbon recovery from natural gas. Journal of Membrane Science. 2017;522:68-76
  22. 22. Stern SA, Shah VM, Hardy BJ. Structure-permeability relationships in silicone polymers. Journal of Polymer Science Part B: Polymer Physics. 1987;25:1263-1298
  23. 23. Semenova SI. Polymer membranes for hydrocarbon separation and removal. Journal of Membrane Science. 2004;231(1–2):189-207
  24. 24. Abang T, Vaidya MM, Prada IC, Duval SA, Ballaguet J-PR, Mohammad AF, et al. Membrane application for NGL production enhancement. In: GPA Europe / GPA GCC joint Annual Conference, Athens, 21-23 September, 2016. Available from: https://gpaeurope.com/library/secure/membrane-application-ngl-production-enhancement
  25. 25. Yang J, Vaidya MM, Harrigan DJ, Duval SA, Hamad F, Bahamdan AA. Modified rubbery siloxane membranes for enhanced C3+ hydrocarbon recovery from natural gas: Pure and multicomponent gas permeation evaluation. Separation and Purification Technology. 2020;242:116774
  26. 26. Rao HX, Liu FN, Zhang ZY. Preparation and oxygen/nitrogen permeability of PDMS crosslinked membrane and PDMS/tetraethoxysilicone hybrid membrane. Journal of Membrane Science. 2007;303:132-139
  27. 27. Haider B, Dilshad MR, Rehman MAU, Schmitz JV, Kaspereit M. Highly permeable novel PDMS coated asymmetric polyethersulfone membranes loaded with SAPO-34 zeolite for carbon dioxide separation. Separation and Purification Technology. 2020;248:116899
  28. 28. Haider B, Dilshad MR, Rehman MAU, Akram MS, Kaspereit M. Highly permeable innovative PDMS coated polyethersulfone membranes embedded with activated carbon for gas separation. Journal of Natural Gas Science and Engineering. 2020;81:103406
  29. 29. Bakonyi P, Bogdán F, Kocsi V, Nemestóthy N, Bélafi-Bakó K, Buitrón G. Investigating the effect of hydrogen sulfide impurities on the separation of fermentatively produced hydrogen by PDMS membrane. Separation and Purification Technology. 2016;157:222-228
  30. 30. Liang C-Z, Yong W-F, Chung T-S. High-performance composite hollow fiber membrane for flue gas and air separations. Journal of Membrane Science. 2017;541:367-377
  31. 31. Moradi MR, Chenar MP, Noie SH. Using PDMS coated TFC-RO membranes for CO2/N2 gas separation: Experimental study, modeling, and optimization. Polymer Testing. 2016;56:287-298
  32. 32. Scholes CA, Bacus J, Chen GQ, Tao WX, Li G, Qader A, et al. Pilot plant performance of rubbery polymeric membranes for carbon dioxide separation from syngas. Journal of Membrane Science. 2012;389:470-477
  33. 33. Uwitonze H, Hwang KS, Lee I. Improving NGL recovery process with dividing-wall column for offshore applications. Chemical Engineering and Processing: Process Intensification. 2020;147:107747
  34. 34. White Paper: “Liquid Hydrocarbon Drop Out in Natural Gas Infrastructure”. NGC+ Liquid Hydrocarbon Drop Out Task Group; 2005. Available from: https://www.beg.utexas.edu/files/cee/legacy/NGC_HDP_Paper.pdf
  35. 35. Mehra R. Conversion of Lean Oil Absorption Process to Extraction Process for Conditioning Natural Gas. U.S. Patent 4696688
  36. 36. Dragomirova R, Stöhr M, Hecker C, Lubenau U, Paschek D, Wohlra S. Desorption-controlled separation of natural gas alkanes by zeolite membranes. RSC Advances. 2014;4:59831-59834
  37. 37. Yu L, Grahn M, Hedlund J. Ultra-thin MFI membranes for removal of C3+ hydrocarbons from methane. Journal of Membrane Science. 2018;551:254-260
  38. 38. Arruebo M, Coronas J, Menéndez M, Santamaría L. Separation of hydrocarbons from natural gas using silicalite membranes. Separation Science and Technology. 2001;25:275-286
  39. 39. Ribeiro MA, Ferreira A, Santos JC, Seo YK, Lee UH, Chang JS, et al. Separation of C3/C4 hydrocarbon mixtures by adsorption using a mesoporous iron MOF: MIL-100(Fe). Microporous and Mesoporous Materials. 2012;153:178-190
  40. 40. Wijmans JW, Baker RW. The solution-diffusion model: A review. Journal of Membrane Science. 1995;107:1-21
  41. 41. Bhide BD, Voskericyan A, Stern SA. Hybrid processes for the removal of acid gases from natural gas. Journal of Membrane Science. 1998;140:27-49
  42. 42. Denbeigh K. The Principles of Chemical Equilibrium. Cambridge: Cambridge University Press; 1960
  43. 43. Singh-Ghosal A, Koros AJ. Energetic and entropic contributions to mobility selectivity in glassy polymers for gas separation membranes. Industrial and Engineering Chemistry Research. 1999;38:3647-3654
  44. 44. Raharjo RD, Freeman BD, Sanders ES. Pure and mixed gas CH4 and n-C4H10 sorption and dilation in poly (dimethylsiloxane). Journal of Membrane Science. 2007;292:45-61
  45. 45. Lau CH, Li P, Chung TS, Paul DR. Reverse-selective polymeric membranes for gas separations. Progress in Polymer Science. 2013;38:740-766
  46. 46. Jiang X, Kumar A. Performance of silicone-coated polymeric membrane in separation of hydrocarbons and nitrogen mixtures. Journal of Membrane Science. 2005;25:179-188
  47. 47. Merkel TC, Bondar VI, Nagai K, Freeman BD, Pinnau. Gas sorption, diffusion, and permeation in poly(dimethylsiloxane). Journal of Polymer Science Part B: Polymer Physics. 2000;38:415-434
  48. 48. Khanbabaei G, Vasheghani-Farahani E, Rahmatpour A. Permeability, solubility, and diffusivity of methane and butane in diphenylsiloxane-dimethylsiloxane copolymer membranes. Journal of Macromolecular Science Part B Physics. 2011;50:2376-2392
  49. 49. Iyer GM, Liu L, Zhang C. Hydrocarbon separation by glassy polymer membranes. Journal of Polymer Science. 2020;58:2482-2517
  50. 50. Jordan SM, Koros WJ. Permeability of pure and mixed gases in silicone rubber at elevated pressures. Journal of Polymer Science Part B: Polymer Physics. 1990;28:795-809
  51. 51. Pinnau I, He Z. Pure-and mixed-gas permeation properties of poly-dimethyl siloxane for hydrocarbon/methane and hydrocarbon/hydrogen separation. Journal of Membrane Science. 2004;244:227-233
  52. 52. Raharjo RD, Freeman BD, Paul DR, Sarti GC, Sanders ES. Pure and mixed gas CH4 and n-C4H10 permeability and diffusivity in poly(dimethylsiloxane). Journal of Membrane Science. 2007;306(9):75-92
  53. 53. Sadrzadeh M, Shahidi K, Mohammadi T. Effect of operating parameters on pure and mixed gas permeation properties of a synthesized composite PDMS/PA membrane. Journal of Membrane Science. 2009;342:327-340
  54. 54. Ghadimi A, Sadrzadeh M, Shahidi K, Mohammadi T. Ternary gas permeation through a synthesized PDMS membrane: Experimental and modeling. Journal of Membrane Science. 2009;344:225-236
  55. 55. Khanbabaei G, Vasheghani-Farahani E, Rahmatpour A. Pure and mixed gas CH4 and n-C4H10 permeation in PDMS-fumed silica nanocomposite membranes. Chemical Engineering Journal. 2012;191:369-377
  56. 56. Grushevenko EA, Borisov IL, Bakhtin DS, Bondarenko GN, Levin IS, Volkov AV. Silicone rubbers with alkyl side groups for C3+ hydrocarbon separation. Reactive and Functional Polymers. 2019;134:156-165
  57. 57. Mushardt H, Muller M, Shishatskiy S, Wind J, Brinkmann T. Detailed investigation of separation performance of a MMM for removal of higher hydrocarbons under varying operating conditions. Membranes. 2016;6:16
  58. 58. Low ZX, Budd PM, McKeown NB, Patterson DA. Gas permeation properties, physical aging, and its mitigation in high free volume glassy polymers. Chemical Reviews. 2018;118:5871-5911
  59. 59. Pinnau I, Toy LG. Transport of organic vapors through poly(1-trimethylsilyl-1-propyne). Journal of Membrane Science. 1996;116:199-209
  60. 60. Pinnau I, Casillas CG, Morisato A, Freeman BD. Hydrocarbon/hydrogen mixed gas permeation in poly(1-trimethylsilyl-1-propyne) (PTMSP), poly(1-phenyl-1-propyne) (PPP), and PTMSP/PPP blends. Journal of Polymer Science Part B: Polymer Physics. 1996;34:2613-2621
  61. 61. Pinnau I, He Z. Filled Super Glassy Membrane. U.S. Patent 6316684
  62. 62. Prabhakar RS, Merkel TC, Freeman BD, Imizu T, Higuchi A. Sorption and transport properties of propane and perfluoropropane in poly(dimethylsiloxane) and poly(1-trimethylsilyl-1-propyne). Macromolecules. 2005;38:1899-1910
  63. 63. Raharjo RD, Freeman BD, Paul DR, Sanders ES. Pure and mixed gas CH4 and n-C4H10 permeability and diffusivity in poly(1-trimethylsilyl-1-propyne). Polymer. 2007;48:7329-7344
  64. 64. Grinevich Y, Starannikova L, Yampolskii Y, Gringolts M, Finkelshtein E. Solubility controlled permeation of hydrocarbons in novel highly permeable polymers. Journal of Membrane Science. 2011;378:250-256
  65. 65. Gomes D, Nunes SP, Peinemann K-V. Membranes for gas separation based on poly(1-trimethylsilyl-1-propyne)–silica nanocomposites. Journal of Membrane Science. 2005;246:13-25
  66. 66. Morisato A, Pinnau I. Synthesis and permeation properties of poly (4-methyl 2-pentyne). Journal of Membrane Science. 1996;121:243-250
  67. 67. Yave W, Peinemann KV, Shishatskiy S, Khotimskiy V, Chirkova M, Matson S, et al. Synthesis, characterization, and membrane properties of poly(1-trimethylgermyl-1-propyne) and its nanocomposite with TiO2. Macromolecules. 2007;40:8991-8998
  68. 68. Hu Y, Shiotsuki M, Sanda F, Freeman BD, Masuda T. Synthesis and properties of Indan-based polyacetylenes that feature the highest gas permeability among all the existing polymers. Macromolecules. 2008;41:8525-8532
  69. 69. Pinnau I, Morisato A, He Z. Influence of side-chain length on the gas permeation properties of poly(2-alkylacetylenes). Macromolecules. 2004;37:2823-2828
  70. 70. Finkelshtein E, Makovetsii K, Gringolts M, Rogan Y, Golenko T, Starannikova L, et al. Addition-type polynorbornenes with Si(CH3)3 side groups: Synthesis, gas permeability, and free volume. Macromolecules. 2006;39:7022-7029
  71. 71. Yampolskii Y, Starannikova L, Belov N, Gringolts M, Finkelshtein E, Shantarovich V. Addition—type polynorbornene with Si(CH3)3 side groups: Detailed study of gas permeation, free volume and thermodynamic properties. Membrane Gas Separation. 1st ed. John Wiley and Sons Ltd; 2010;3:43-57
  72. 72. Gringolts ML, Bermeshev MV, Starannikova LE, Rogan YV, Yampol’skii YP, Finkel’shtein ES. Synthesis and gas separation properties of metathesis polynorbornenes with different positions of one or two SiMe3 groups in a monomer unit. Polymer Science, Series A. 2009;51:1233-1240
  73. 73. Bermeshev MV, Syromolotov AV, Gringolts ML, Starannikova LE, Yampolskii YP, Finkelshtein ES. Synthesis of high molecular weight poly[3-{tris(trimethylsiloxy)silyl}tricyclononenes-7] and their gas permeation properties. Macromolecules. 2011;44:6637-6640
  74. 74. Gringolts M, Bermeshev M, Yampolskii Y, Starannikova L, Shantarovich V, Finkelshtein E. New high permeable addition poly(tricyclononenes) with Si(CH3)3 side groups. Synthesis, gas permeation parameters, and free volume. Macromolecules. 2010;43:7165-7162
  75. 75. Sundell BJ, Lawrence JA III, Harrigan DJ, Lin S, Headrick TP, O’Brien JT, et al. Exo-selective, reductive heck derived polynorbornenes with enhanced molecular weights, yields, and hydrocarbon gas transport properties. ACS Macro Letters. 2020;9:1363-1368
  76. 76. Alentiev DA, Egorova ES, Bermeshev MV, Starannikova LE, Topchiy MA, Asachenko AF, et al. Anus tricyclononene polymers bearing tri(n-alkoxy)silyl side groups for membrane gas separation. Journal of Materials Chemistry A. 2018;6:19393-19408
  77. 77. Alentiev DA, Bermeshev MV, Starannikova LE, Bermesheva EV, Shantarovich VP, Bekeshev VG, et al. Stereoselective synthesis and polymerization of Exo-5-trimethylsilylnorbornene. Journal of Polymer Science, Part A: Polymer Chemistry. 2018;56:1234-1248
  78. 78. Wozniak AI, Bermesheva EV, Andreyanov FA, Borisov IL, Zarezin DP, Bakhtin DS, et al. Modifications of addition poly(5-vinyl-2-norbornene) and gas-transport properties of the obtained polymers. Reactive and Functional Polymers. 2020;149:104513
  79. 79. Wozniak AI, Borisov IL, Bermesheva EV, Zarezin DP, Volkov AV, Finkelstein ES, et al. Influence of the nature of chemical modification of addition poly(5-vinyl-2-norbornene) on the gas permeability of hydrocarbons. Polymer Science, Series B. 2020;62:218-224
  80. 80. Alentiev DA, Dzhaparidze DM, Bermeshev MV, Starannikova LE, Filatova MP, Yampolskii YP, et al. Addition copolymerization of silicon-containing tricyclononene with 2,5-norbornadiene dimer. Polymer Science, Series B. 2019;61:812-816
  81. 81. Alentiev DA, Egorova ES, Bermeshev MV, Starannikova LE, Yampolskii YP, Finkelshtein ES. Crosslinking of addition copolymers from tricyclononenes bearing (CH3)3Si- and (C2H5O)3Si-groups as a modification of membrane gas separation materials. Polymer Engineering and Science. 2019;59:2502-2507
  82. 82. Alentiev DA, Egorova ES, Bermeshev MV, Starannikova LE, Topchiy MA, Asachenko AF, et al. Janus tricyclononene polymers bearing tri(n-alkoxy)silyl side groups for membrane gas separation. Journal of Materials Chemistry A. 2018;6:19393-19408
  83. 83. Pinnau I, He Z. Pure- and mixed-gas permeation properties of polydimethylsiloxane for hydrocarbon/methane and hydrocarbon/hydrogen separation. Journal of Membrane Science. 2004;244:227-233
  84. 84. Sadrzadeh M, Shahidi K, Mohammad T. Preparation and C3H8/gas separation properties of a synthesized single layer PDMS membrane. Separation Science and Technology. 2010;45:592-603
  85. 85. Farno E, Ghadimi A, Kasiri N, Mohammadi T. Separation of heavy gases from light gases using synthesized PDMS nano-composite membranes: Experimental and neural network modeling. Separation and Purification Technology. 2011;81:400-410
  86. 86. Chen B, Ruan X, Xiao W, Jiang X, He G. Synergy of CO2 removal and light hydrocarbon recovery from oil-field associated gas by dual-membrane process. Journal of Natural Gas Science and Engineering. 2015;26:1254-1263
  87. 87. Fang M, Zhang H, Chen J, Wang T, Liu J, Li X, et al. A facile approach to construct hierarchical dense membranes via polydopamine for enhanced propylene/nitrogen separation. Journal of Membrane Science. 2016;499:290-300
  88. 88. Grushevenko EA, Borisov IL, Bakhtin DS, Legkov SA, Bondarenko GN, Volkov AV. Membrane material based on octyl-substituted polymethylsiloxane for separation of C3/C1 hydrocarbons. Petroleum Chemistry. 2017;57:334-340
  89. 89. Mushardt H, Kramer V, Hülagü D, Brinkmann T, Kraume M. Development of solubility selective mixed matrix membranes for gas separation. Chemie Ingenieur Technik. 2014;86:83-91
  90. 90. Alentiev A, Economou IG, Finkelshtein E, Petrou J, Raptis VE, Sanopoulou M, et al. Transport properties of silmethylene homo-polymers and random copolymers: experimental measurements and molecular simulation. Polymer. 2004;45:6933-6944
  91. 91. Khosravi A, Sadeghi M. Separation performance of poly(urethane–urea) membranes in the separation of C2 and C3 hydrocarbons from methane. Journal of Membrane Science. 2013;434:171-183
  92. 92. Gomesa D, Peinemann KV, Nunes SP, Kujawski W, Kozakiewicz J. Gas transport properties of segmented poly (ether siloxane urethane urea) membranes. Journal of Membrane Science. 2006;281:747-753
  93. 93. Harrigan DJ, Yang J, Sundell BJ, Lawrence JA, O’Brien JT, Ostraat ML. Sour gas transport in poly(ether-b-amide) membranes for natural gas separations. Journal of Membrane Science. 2020;595:117497
  94. 94. Tirouni I, Sadeghi M, Pakizeh M. Separation of C3H8 and C2H6 from CH4 in polyurethane–zeolite 4Å and ZSM-5 mixed matrix membranes. Separation Science and Technology. 2015;141:394-402
  95. 95. Yang J, Vaidya MM, Tammana VVR, Harrigan D. Modified Siloxane Composite Membranes for Heavy Hydrocarbon Recovery. U.S. Patent 10293301
  96. 96. Masuda T, Isobe E, Higashimura T, Takada K. Poly[1-(trimethylsilyl)-1-propyne]: A new high polymer synthesized with transition-metal catalysts and characterized by extremely high gas permeability. Journal of the American Chemical Society. 1983;105:7473-7474
  97. 97. Budd PM, Msayib KJ, Tattershall CE, Ghanem BS, Reynolds KJ, McKeown NB, et al. Gas separation membranes from polymers of intrinsic microporosity. Journal of Membrane Science. 2005;251:263-269
  98. 98. Wang Y, Ma X, Ghanem BS, Alghunaimi F, Pinnau I, Han Y. Polymers of intrinsic microporosity for energy-intensive membrane-based gas separations. Materials Today Nano. 2018;3:69-95
  99. 99. Wang X, Wilson TJ, Alentiev DA, Gringolts M, Finkelshtein E, Bermeshev M, et al. Substituted polynorbornene membranes: A modular template for targeted gas separations. Polymer Chemistry. 2021;12:2947-2977
  100. 100. Zhan X, Wang M, Gao T, Lu J, He Y, Li L. A highly selective sorption process in POSS-g-PDMS mixed matrix membranes for ethanol recovery via pervaporation. Separation and Purification Technology. 2020;236:116238
  101. 101. Hu M, Wu Z, Sun L, Guo S, Li L, Lia J, et al. Improving pervaporation performance of PDMS membranes by interpenetrating polymer network for recovery of bio-butanol. Separation and Purification Technology. 2019;228:115690
  102. 102. Si Z, Liu C, Li G, Wang Z, Li J, Xue T, et al. Epoxide-based PDMS membranes with an ultrashort and controllable membrane-forming process for 1-butanol/water pervaporation. Journal of Membrane Science. 2020;612:118472
  103. 103. Shan VM, Hardy BJ, Stern SA. Solubility of carbon dioxide, methane and propane in silicon polymers. Effect of polymer side chains. Journal of Polymer Science Part B: Polymer Physics. 1986;24:2033-2047
  104. 104. Stern SA, Shah VM, Hardy BJ. Structure-permeability relationships in silicone polymer. Journal of Polymer Science Part B: Polymer Physics. 1987;25:1263-1298
  105. 105. Shan VM, Hardy BJ, Stern SA. Solubility of carbon dioxide, methane and propane in silicon polymers. Effect of polymer backbone chains. Journal of Polymer Science Part B: Polymer Physics. 1993;31:313-317
  106. 106. Yilgor E, Yilgor I. Silicone containing copolymers: Synthesis, properties and applications. Progress in Polymer Science. 2014;39:1165-1195
  107. 107. Ghosh A, Sen SK, Dasgupta B, Banerjee S, Voit B. Synthesis, characterization and gas transport properties of new poly (imide siloxane) copolymers from 4,4-(4,4-isopropylidenediphenoxy) bis(phthalicanhydride). Journal of Membrane Science. 2010;364:211-218
  108. 108. Powell CE, Qiao G, Kebtish SE. Gas Separation Membranes and Processes for the Manufacture Therefor U.S. Patent Publication 2010/0313752
  109. 109. Hong T, Chatterjee S, Mahurin SM, Fan F, Tian Z, Jiang D, et al. Impact of tuning CO2-philicity in polydimethylsiloxane-based membranes for carbon dioxide separation. Journal of Membrane Science. 2017;530:213-219
  110. 110. Madhavan K, Reddy BSR. Poly(dimethylsiloxane-urethane) membranes: Effect of hard segment in urethane on gas transport properties. Journal of Membrane Science. 2006;283:357-365
  111. 111. Sadeghi M, Talakesh MM, Ghalei B, Shafiei M. Preparation, characterization and gas permeation properties of a polycaprolactone based polyurethane– Silica nanocomposite membrane. Journal of Membrane Science. 2013;427:21-29
  112. 112. Talakesh MM, Sadeghi M, Pourafshari Chenar M, Khosravi A. Gas separation properties of poly(ethylene glycol)/poly(tetramethylene glycol) based polyurethane membranes. Journal of Membrane Science. 2012;415–416:469-477
  113. 113. Sadeghia M, Semsarzadehb MA, Barikanic M, Ghaleib B. The effect of urethane and urea content on the gas permeation properties of poly(urethane–urea) membranes. Journal of Membrane Science. 2010;354:40-47
  114. 114. Sadeghi M, Semsarzadeh MA, Barikani M, Ghalei B. Study on the morphology and gas permeation property of polyurethane membranes. Journal of Membrane Science. 2011;385–386:76-85
  115. 115. Isfahani AP, Ghalei B, Bagheri R, Kinoshita Y, Kitagawa H, Sivaniah E, et al. Polyurethane gas separation membranes with ethereal bonds in the hard segments. Journal of Membrane Science. 2016;513:58-66
  116. 116. Fakhar A, Sadeghi M, Dinaru M, Lammertink R. Comparative assessment of hydrocarbon separation performance of bulky poly(urethane-urea)s toward rubbery membranes. Journal of Natural Gas Science and Engineering. 2021;27:104356
  117. 117. Liu SL, Shao L, Chua ML, Lau CH, Wang H, Quan S. Recent progress in the design of advanced PEO-containing membranes for CO2 removal. Progress in Polymer Science. 2013;38:1089-1120
  118. 118. Amo K et al. Final Report to DOE “Low-Quality Natural Gas Sulfur Removal/Recovery”. Morgantown, WV: MTR; 1998
  119. 119. Ogbole EO, Lou J, Ilias S, Desmane V. Influence of surface treated SiO2 on the transport behavior of O2 and N2 through polydimethylsiloxane nanocomposite membrane. Separation and Purification Technology. 2017;175:358-364
  120. 120. Ha H, Park J, Ando S, Kim CB, Nagai K, Freeman BD, et al. Gas permeation and selectivity of poly(dimethylsiloxane)/graphene oxide composite elastomer membranes. Journal of Membrane Science. 2016;518:131-140
  121. 121. Molki B, Aframehr WM, Bagheri R, Salimi J. Mixed matrix membranes of polyurethane with nickel oxide nanoparticles for CO2 gas separation. Journal of Membrane Science. 2018;549:588-601
  122. 122. Zhan X, Lu J, Xu H, Liu J, Liu X, Cao X, et al. Enhanced pervaporation performance of PDMS membranes based on nanosized octa[(trimethoxysilyl)ethyl]-POSS as macro-crosslinker. Applied Surface Science. 2019;473:785-798
  123. 123. Rezakazemi M, Shahidi K, Mohammadi T. Hydrogen separation and purification using crosslinkable PDMS/zeolite A nanoparticles mixed matrix membranes. International Journal of Hydrogen Energy. 2012;37:14576-14589
  124. 124. Najari S, Saeidi S, Gallucci F. Mixed matrix membranes for hydrocarbons separation and recovery: A critical review. Reviews in Chemical Engineering. 2019;37(3):363-406
  125. 125. Sadrzadeh M, Amirilargani M, Shahidi K, Mohammadi T. Pure and mixed gas permeation through a composite polydimethylsiloxane membrane. Polymers for Advanced Technologies. 2011;22:586-597
  126. 126. Kim H, Kim H-G, Kim S, Kim SS. PDMS-silica composite membranes with silane coupling for propylene separation. Journal of Membrane Science. 2009;344:211-218
  127. 127. Shen G, Zhao J, Guan K, Shen J, Jin W. Highly efficient recovery of propane by mixed-matrix membrane via embedding functionalized graphene oxide nanosheets into polydimethylsiloxane. AICHE Journal. 2017;63:3501-3510
  128. 128. Woo M, Choi J, Tsapatsis M. Poly(1-trimethylsilyl-1-propyne)/MFI composite membranes for butane separations. Microporous and Mesoporous Materials. 2008;110:330-338
  129. 129. Sundell BJ, Lawrence JA III, Harrigan DJ, Vaughn JT, Pilyugina TS, Smith DR. Alkoxysilyl functionalized polynorbornenes with enhanced selectivity for heavy hydrocarbon separations. RSC Advances. 2016;6:51619-51628

Written By

John Yang, Milind M. Vaidya, Sebastien A. Duval and Feras Hamad

Submitted: 05 February 2022 Reviewed: 24 February 2022 Published: 21 April 2022