Download as pdf or txt
Download as pdf or txt
You are on page 1of 155

Journal Pre-proofs

Snapshot Reviews in Emerging Fields

Recent advances in polymers of intrinsic microporosity (PIMs) membranes:


Delving into the intrinsic microstructure for carbon capture and arduous in-
dustrial applications

Hui Shen Lau, Angelica Eugenia, Ying Weng, Wai Fen Yong

PII: S0079-6425(24)00066-5
DOI: https://doi.org/10.1016/j.pmatsci.2024.101297
Reference: JPMS 101297

To appear in: Progress in Materials Science

Received Date: 4 February 2024


Revised Date: 2 April 2024
Accepted Date: 11 April 2024

Please cite this article as: Lau, H.S., Eugenia, A., Weng, Y., Yong, W.F., Recent advances in polymers of
intrinsic microporosity (PIMs) membranes: Delving into the intrinsic microstructure for carbon capture and
arduous industrial applications, Progress in Materials Science (2024), doi: https://doi.org/10.1016/j.pmatsci.
2024.101297

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. 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.

© 2024 Published by Elsevier Ltd.


Recent advances in polymers of intrinsic microporosity (PIMs) membranes: Delving into

the intrinsic microstructure for carbon capture and arduous industrial applications

Hui Shen Laua, Angelica Eugeniaa, Ying Wenga, Wai Fen Yong a,b*

aSchool of Energy and Chemical Engineering, Xiamen University Malaysia, Selangor Darul

Ehsan 43900, Malaysia

bState Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and

Chemical Engineering, Xiamen University, Xiamen 361005, China

*Correspondence author:

Dr. Wai Fen Yong

E-mail address: waifen.yong@xmu.edu.my (W.F. Yong)

1
Abstract

Polymers of intrinsic microporosity (PIMs) are unique polymers known for their intrinsic micro-

scale porosity contributed by bulky and rigid contortion sites in the polymer backbone. Inherent

attributes of PIMs, such as structural diversity and good processability have made them valuable

in various applications. Herein, we outlined a comprehensive overview on the latest progress of

ladder PIMs on different industrial challenges. This review has systematically discussed the state-

of-the-art ladder PIMs redesigned on intrinsic micro-structure through five different perspectives,

including (i) architecting the polymer backbone, (ii) post-modification on polymer structure, (iii)

polymer blends and copolymerization, (iv) mixed matrix membranes (MMMs), and (v) post-

modification on membranes, aiming to address the carbon-related international treaties. A

summary of their CO2 capture performance on Robeson plots is portrayed and evaluated. In

addition, the implementation of PIMs in energy-efficient membrane-based olefin/paraffin

separation is highlighted. Subsequently, solution-processable ladder PIMs, in the form of powder,

nanofibrous, films or membranes applied in the field of environmental application, catalysis,

electrochemical energy storage and conversion, sensing, and 3D printing are emphasized. Along

with the contemplation on outlook and future perspective, this review is anticipated to path a new

avenue for the continuous development and optimization of PIMs materials in sustainable

applications.

Keywords: Polymers of intrinsic microporosity; Porous polymer; Membrane technology; Carbon

capture; Light hydrocarbon separation; Environmental application

2
1. Introduction

Microporous polymers with tailorable pore architectures, steerable post-modification, and solution

processability are desirable for targeted separation or adsorption applications. Ultrapermeable

microporous polymer, poly(1-trimethylsilyl-1-propype) (PTMSP) based on linear-chain

disubstituted glassy polyacetylenes was first discovered by Masuda et al. [1] in the early 1980s.

The PTMSP membrane possesses high free volume with an unprecedented permeability (e.g., O2

permeability of ~10000 Barrer). However, it has a low O2/N2 selectivity of 1.5 [2, 3].

Polymers of intrinsic microporosity (PIMs), a novel family of amorphous super-glassy solution-

processable microporous polymer materials have attracted tremendous research interest due to

unique intrinsic microporosity. The term ‘polymers of intrinsic microporosity’ was first introduced

by Budd and McKeown in 2004, referring to a continuous network of linked intermolecular voids

resulting from the rigid structure of the constituent macromolecules [4, 5]. PIMs consist of a wide

range of bulky ladder-like building blocks connected by a spiro-center [6]. The rigid and inflexible

backbone structures, likewise, provide high pore-free volumes due to the abundance of contortion

sites resulting from inefficient chain packing. Solution-processability refers to the ease of a

substance to be dissolved or dispersed in a solvent to achieve a desired form or concentration. It is

pivotal in industries which require precise control of distribution and concentration of substance,

such as manufacturing and pharmaceuticals. Among the developed solution-processed PIMs [7],

the PIM-1 consisted of a larger amount of interconnected pores of < 2 nm diameter and a large

surface area (e.g., 500-900 m2/g) [8, 9]. The intrinsic microporosity of PIMs is ascribed to the

presence of contortion sites and the rigid fused-ring structures induced by the spiro-center (single

tetrahedral carbon shared by two cyclic rings), which prohibit the structure-relaxation and bond-

rotation in the solid state. It possesses a high thermal stability of 350 °C [4]. PIMs are classified as

3
ladder PIMs (e.g., PIM-1 or Tröger’s base PIMs), linear PIMs (e.g., polyimides or polyamides)

and hyper-cross-linked PIMs [7]. Different from other developing microporous organic polymer

(MOPs) materials, such as hyper-cross-linked polymers (HCPs) [10], covalent-organic framework

(COFs) [11], and porous organic polymers (POPs) [12], PIMs offer better solubility in common

solvents, allowing them to be fabricated into thin film or membrane [13, 14]. Furthermore, PIMs

also exhibited superior compatibility with other materials, such as carbon-based fillers [15, 16],

metal-organic frameworks (MOFs) [17, 18], or organic polymers [19, 20].

The intrinsic porosity and high surface area of PIMs have favored tremendous interest in separation

such as water treatment, air purification and gas separation, either membrane-based or adsorption.

A surge of interest in the research of PIMs has been witnessed in the past few decades, whether in

powder, films, membranes, or nanofibrous form in various applications. Figure 1 depicts the

number of publications on PIMs in CO2 capture, adsorption, catalysis, and electrochemistry from

2013 – 2023, in which more than 900 articles and ~80 patents have been reported. Amongst them,

several review articles focus on PIMs in various applications. Zou et al. provided a comprehensive

review on the use of MOPs materials, such as PIMs, COFs, and porous aromatic frameworks

(PAFs) for membrane gas separation such as H2, CO2, O2 and other industrial-relevant gases in

2018 [21]. In another review article, Wang et al. summarized the role of state-of-the-art

functionalized ladder PIMs and PIM-PI (PIM-PI: polyimides of intrinsic microporosity) in a series

of industrial energy-intensive membrane-based gas separations such as CO2/CH4, H2S/CH4,

N2/CH4, and olefin/paraffin separations [7]. They reported another extensive review emphasizing

thermally rearranged (TR) and carbon molecular sieve (CMS) analogues of ladder PIMs and PIM-

PI membranes in the same application [22]. The same group summarized ladder PIMs,

functionalized PIMs, PIMs-based thin film composite (TFC) membranes and PIMs-based CMS

4
membranes for gas separation [23]. Ahmad et al. provided an analysis of the incorporation of

nanomaterials in retarding physical aging in the PIM-1 matrix [24]. Later, He et al. summarized

the latest progress of PIM-1 based membranes in CO2 separation based on two perspectives,

namely the design and modification of PIM-1 intrinsic structure, and fabrication of PIM-1/MMMs

[25]. In another study, Wang et al. discussed the progress of PIMs films and membranes as

versatile class of materials in electrochemical energy systems [26], while Antonangelo et al.

summarized the implementation of PIMs in catalytic processes [27]. Recently, Topuz et al.

reviewed the usage of PIMs materials as membranes and adsorbents in environment, sensing,

catalysis and energy applications [28]. Innumerable advancement was witnessed in PIMs materials

in the past 10 years.

PIMs
200 PIMs in adsorption
181 185
PIMs in carbon capture 179
PIMs in catalysis 165
157 155
PIMs in electrochemistry
143
Number of publications

150

118
111 112

100

50 48 48 46 48
43 44 41 42
36 38
28 29 27 29 27 31
21 24 2521 20
16 17 16 17 20 15
13 11 12 10
7 3 4
2 2 2 2 2 2 2 211
0
2014 2015 2016 2017 2018 2019 2020 2021 2022 2023 2024
Years

Figure 1. Number of publications in Scopus on PIMs in adsorption, carbon capture, catalysis,

and electrochemistry over the last decade [accessed on 10 March 2024].

Due to the significant importance of the unique performance of PIMs, this review aims to provide

in-depth insight into the latest development and progress of ladder PIMs from molecular

5
architecture to their potential in the form of powder, films, membranes, or nanofibrous in various

applications, including CO2 capture, light hydrocarbon separation, environmental application,

catalysis, electrochemical energy storage and conversion, sensing, and 3D printing. Notably,

collective research endeavors have yielded impressive outcomes in the realm of PIMs applications.

Herein, the recent state-of-the-art approach to redesigning the intrinsic structure of ladder PIMs

will be discussed, including (1) structural design and architecting on PIMs polymer backbone, (2)

post-modification of PIMs polymer structure, (3) polymer blends or copolymerization with PIMs,

(4) mixed matrix membranes (MMMs), and (5) post-modification on membranes, aiming to

enhance the membrane-based CO2 capture performance. Robeson plots on CO2/N2 and CO2/CH4

separation will be illustrated to assess the outperform ladder PIMs membranes for relevant

industrial applications. Additionally, recent designs of PIMs membranes are provided to realize an

energy-efficient light hydrocarbon separation. The latest advancements in PIMs in sophisticated

applications will also be elucidated. This review strives to create insightful and perceptive

guidance for scientists on the recent advancement of PIMs, fostering innovation and drive towards

sustainable applications.

2. Theoretical background

2.1. Gas transport theory

In the past few decades, membranes have gained a numerous attention in gas separation because

of their advantages of a small carbon footprint, high scalability potential, time-effective operation,

and low capital expenses [5, 14, 29]. Gas transport across a dense polymeric membrane is governed

by a solution-diffusion mechanism, in which gas transports across the membrane via a

concentration gradient. Gas transport in a membrane can be analyzed by permeability and

selectivity. The rate of gases permeate through a membrane is called gas permeability as below.

6
P=DxS Eq. (1)

where P is gas permeability (Barrer, 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg), D is the

function of diffusion coefficient (cm2/s) and S is the solubility coefficient (cm3(STP)/(cm3.atm)).

The ideal selectivity (αA/B) of a fast gas A over a slow gas B is obtained by the ratio of gas

permeability of gas A and gas B, as shown by:


𝑃𝐴
𝛼𝐴/𝐵 = 𝑃𝐵 Eq. (2)

𝐷𝐴 𝑆𝐴
𝛼𝐴/𝐵 = 𝐷𝐵 × 𝑆𝐵

𝛼𝐴/𝐵 = 𝛼𝐷 × 𝛼𝑆

where αD and αS are the diffusivity selectivity and solubility selectivity, respectively.

2.2. Trade-off relationship in polymeric membrane

The permeability-selectivity trade-off upper bound was first plotted by Robeson in 1991 based on

various gas pairs over a large database focusing on polymeric membrane materials [30]. The upper

bound curves were first revisited in 2008 for newly developed materials such as PIMs and

perfluorinated polymers [31]. Structural design, architecting and tailoring of the polymer materials

have led to the establishment of a more recent upper bound in 2015 for O2/N2, H2/N2 and H2/CH4

[32], and later in 2019 for CO2/N2 and CO2/CH4 [33]. An upper bound for binary CO2/CH4 mixed

gas at 10 bar CO2 partial pressure at 35 oC was also proposed in 2018 [7]. Noteworthy, these upper

bound curves were developed based on ideal pure gas permeation data.

2.3. Polydispersity index (PDI)

Polydispersity index (PDI) or molecular weight distribution is an important physical attribute that

influences the properties and stability of polymers [34, 35]. The dispersity index represents the

ratio of average molecular weight (Mw) to the number average molecular weight (Mn). Polymer

7
materials with different properties exhibit specific dependencies on molecular dispersity profile.

However, excessively broad molecular weight distributions are prone to uncontrolled

polymerization. To obtain more uniform dispersity, some tuning methods to control and advance

material properties for a wide range of applications have been developed over the past decade [34,

36]. The varying dispersity index can exhibit complementary characteristics and purposes, making

it an important factor that affects the physical properties of polymers. A polymer with a higher

PDI generally possesses greater heterogeneity and a broader molecule weight range, which

influence its physical properties, such as increasing stiffness, viscosity, thermal stability, Tg, as

well as the effect in polymer solubility properties due to an increase in the turbidity of polymer

solution [37-39]. Despite being influential to the polymer properties, the varying of PDI does not

pose a significant effect on the thermal conductivity measured at a constant average molecular

weight [40].

2.4. Ladder PIM polymers

To overcome the trade-off and achieve a step-change in solution-diffusion membranes, altering

polymers with unique molecular architectures is the most promising approach. Ladder PIMs are

categorized based on the type of contorted units or sterically hindered building blocks, such as

spirocyclic units, bridged bicyclic moieties, Tröger’s base, and catalytic arene-norbornene

annulation (CANAL) (Figure 2). The earliest developed PIM-1 with spirocyclic and benzodioxin

structure contains bare hydrogen and nitrile groups distributed along the spirobisindane moieties.

PIM-1 was fabricated through the polymerization reaction of 5,5’,6,6’-tetrahydroxy-3,3,3’,3’-

tetramethyl-1,1’-spirobisindane (TTSBI), and 2,3,5,6-tetrafluoroterephthalonitrile (TFTPN) via

the non-reversible aromatic nucleophilic substitution. The more recent high-temperature route was

conducted at a higher temperature with a shorter duration (e.g., at 160 °C for 45 min) and resulted

8
in PIM with a higher molecular weight with narrower size distribution [14, 41]. Figure 3a

illustrates the conventional synthesis process of PIM-1 via low and high-temperature routes. PIM-

1 can also be prepared using a cheaper monomer, namely tetrachloroterephthalonitrile (TCTPN),

to replace TFTPN [42]. Various structures of chloro-based PIM-1 (e.g., cyclic, tadpole, branched,

extended and network topologies) with different network content were reported under a series of

synthesis conditions (Figure 3b). The membranes casted from higher branching and network

contents were found to demonstrate better gas separation performance, surpassing the 2008

Robeson upper bound.

Following the succession synthesis of PIM-1, several other PIMs were synthesized using different

bis(catechol) monomers, namely, PIM-2 to 10 [43, 44]. All PIMs share the similar characteristic

of having intrinsic microporosity in their nature; however, unlike PIM-1, there are limited

tetrahalide monomers that can provide a spiro-center readily available as a contortion site [44].

Therefore, limited studies reported on the synthesis of these derivatives PIMs so far. Apart from

PIM-1, the studies on several spiro-centers containing PIMs (e.g., PIM-2, PIM-7, and PIM-8) have

been conducted over the past decades [45-47]. To date, spirocyclic PIMs have been used in many

applications, including hydrogen storage [48], gas separations [49], as well as chemical sensing

equipment [50]. Another spirocyclic contortion unit, spirobifluorenes (SBF), is formed by fusing

benzo groups to the spiro-center of spirobisindane [51, 52]. The enhanced rigidity by

spirobifluorenes units would result in larger free volume inter-chain distance and higher gas

permeability.

9
Spirocyclic unit
HO O O O
CN CN CN CN

OH O O O O O
HO O O O

O O O O O
OH n n
Contortion sites CN CN
n
CN CN
PIM-1 PIM-SBF-1 Cardo-PIM-1

Bridged bicycle units


R2

CANAL
R1
*
R3
a) TMN-Trip: R1, R2 = *

O
CN
b) HMI-Trip: R1, R2 = *
O *
O

O
n
O
CN c) BTrip: R1, R2, R3 = H
CN O
O d) DM-BTrip: R1, R2 = Me; R3 = H
O e) TFM-TRip: R1 = CF3; R2, R3 = H
Contortion sites PIM-SBI-Trip f) DTFM-TRip: R1, R3 = CF3; R2 = H
Benzotriptycene PIM NC n

Contortion sites

Tröger's base N
X X'
N
N
CH3 n
N PIM-MP-TB
N n
N PIM-SBI-TB
N n
Y Y'
H 3C Contortion sites N X, X’, Y, Y’ = H, Me, Ome, i-Pr, etc.
PIM-Trip-TB N

PIM-EA-TB
N
CANAL ladder PIM
n

Figure 2. PIMs with different contortion sites.

a) Conventional PIM-1 synthesis with TTSBI & TFTPN


Low temperature O
HO CN
50–70 °C, DMF or NMP CN
F F
OH
K2CO3, 24–72 h O
HO
+F O

F High temperature
OH O
CN 160 °C, DMAc & Toluene n
CN
K2CO3, 45 min–4 h
TTSBI TFTPN PIM-1

b) PIM-1 synthesis with novel monomer, TCTPN and the microstructures predicted in step-growth polymerization
Linear PIM-1
CN
O Cl O O OH
H CN H CN
CN
HO H H
O Cl O Cl O O
O Cl O O OH

OH CN
HO O Cl O Cl O O
n n n
100–160 °C, DMAc & DCB

CN CN CN
OH
Cyclic PIM-1 Under-reacted PIM-1 residue Branched PIM-1 residue NC

TTSBI
K2CO3, < 3 h

O
O CN
O O
CN O
+ O
O
O
O

CN Cl
O
n NC Cl
Cl Cl CN O
NC O
O CN
Cl Cl
OH O
O CN
CN CN
OH O
O CN
HO
TCTPN O
HO
O
CN O
CN

Figure 3. (a) Conventional PIM-1 synthesis routes with synthesis conditions at low- and high-

temperatures. (b) PIM-1 synthesis with novel monomer, TCTPN and the microstructures

10
predicted in step-growth polymerization. Reproduced with permission from ref [42]. Copyright

2020 American Chemical Society.

The shape-persistent bridged bicyclic units, including triptycenes (Trip) [53], pentiptycenes [54],

and benzotriptycenes [33] have also been widely used to architect the backbone of PIMs for gas

separation applications with satisfying performances above the Robeson upper bound. The three-

dimensional paddle-wheel-like triptycenes PIMs have highly-rigid polymer matrixes with

prolonged resistance to aging. In addition, triptycenes-based PIMs have also been recently reported

on the removal of antidepressant chemicals in pharmaceutical wastewater [55]. Correspondingly,

the pentiptycene with double contortion sites provides opportunities for taking on both

permeability-selectivity trade-off and physical aging issues simultaneously [54]. On the other hand,

benzotriptycene-based PIMs with an additional benzo group modification on the triptycene

backbones were ascribed to extraordinary gas separation performance due to the

ultramicroporosity and extreme chain-rigidity [33].

Tröger’s base consists of amine-bridged contortion sites with a similar bicyclic structure as

triptycenes. The previously reported Tröger’s base-based PIM (e.g., PIM-EA-TB) was observed

to maintain its stability under a wet environment better than PIM-1 due to the lower affinity

towards water molecules [56]. The combination of triptycenes and Tröger’s base had also been

studied previously with excellent gas adsorption performance in CO2/CH4 mixed gas test,

benefiting from the intra-chain rigidity induced by the two bulky blocks [6]. Nevertheless, the

structural analysis confirmed the increasing FFV and polymer dilatation due to the reduction in

inter-chain packing. This phenomenon resulted in a significant loss in selectivity under the gas

permeation test. Chen et al. revealed the corolliform structures of TB-PIMs are also useful for

11
enhancing the nanofiltration effect in the desalination process [57]. The characterizations showed

excellent stability of the polymer morphology under a neutral and acidic environment.

CANAL PIM, based on the name, was initially made from the polymerization of aryl bromides

and norbornene using conventional catalysts, such as palladium [58]. The ribbon-like 2D ladder

geometry of CANAL PIMs is characterized by the methylene bridge contortions and

benzocyclobutene backbone linkages. CANAL chemistry has the advantage of using different

bromide-containing aromatics as monomers and ease of functionalization techniques [59]. Lai et

al. synthesized high molecular weight CANAL polymers using 1,4-dibromo-2,5-dimethylbenzene

and norbornadiene and further functionalized them with varying short alkyl substitutions [60].

They found that the incorporation of methyl or isopropyl substituents tuned the chain morphology

and fractional free volume (FFV), enhancing both the sorption and permeation of all tested gasses.

In another study, Hazazi et al. synthesized CANAL-Tröger’s base ladder polymers via high-

temperature pyrolysis whereby the developed membranes showed a remarkable H2/CO2 separation

performance [61].

As of today, many researchers have developed different types of PIMs with unique microporosity

and highly rigid structures, especially for carbon capture applications. The high solubility of PIMs

in various organic solvents as well as excellent thermal and mechanical properties have grounded

the fabrication of PIM films or membranes for challenging gas adsorption and separation processes.

In recent years, many studies have developed PIMs using different kinds of intrinsic microporosity

monomers and methodologies. Apart from that, post-modification functionalization, polymer

blending, and MMMs have been developed to bring PIMs membrane technologies closer to

commercialization. In addition, the other applications of PIMs in sophisticated industrial

12
applications, which benefit from their physical and chemical properties have also been widely

explored.

3. Approaches in designing PIM membranes for excellent CO2 capture

Industrial revolution hastened the emission of greenhouse gases, which exacerbated climate

change and resource scarcity. Over the past few decades, various technologies have been

developed for carbon capture, including chemical and physical adsorption, cryogenic separation,

microbiological method, as well as membrane separation [62, 63]. Membrane technology has been

an emerging carbon capture technique in recent years due to its low energy consumption, ease of

operation, and ease of scale-up [64-66]. The emergence of ultra-permeable polymers of intrinsic

microporosity (PIMs) as gas membrane constructing materials has attracted tremendous studies

over the past 20 years due to their excellent workability, wide application, and rigid structure.

Following the development of highly microporous polymers, many studies have developed new

PIMs with better performance and efficiency, especially in CO2 capture, addressing the carbon

neutrality protocol.

3.1. Structural design and architecting on PIMs polymer backbone

The permeation rate of the gas molecules through membranes is influenced by the solubility and

diffusivity coefficients of the gasses. These factors are determined based on the FFV, which is the

unoccupied volume or gap formed by polymer chains in the membrane [67]. The higher amount

of free volume generally leads to higher gas molecule permeability as a result of large interchain

spacing. However, excessive free volume might also negatively affect the selectivity values due to

weakening interaction between certain gas molecules and the polymer chain [68]. The PIM-1

membrane has a high FFV due to a fixed-contorted molecular configuration as a result of a low

concentration of sigma bonds in the polymer backbone [69].

13
Two decades ago, Budd et al. introduced two promising PIMs for gas separation, e.g., PIM-1 and

PIM-7 [70, 71]. The dominant polarizability from the aromatic structure of PIM-7 enhanced the

van der Waals interactions between the polymer and gas molecules. Additionally, PIM-7 also has

excellent thermal stability with a high degradation temperature of 480 °C. In comparison, the

presence of –CN groups in PIM-1 make it more favorable for greenhouse gas capture application

due to the enhanced affinity towards CO2 [71]. PIM-2 produced through the polycondensation of

TTSBI and decafluorobiphenyl (DFBP) has a relatively high molecular weight and is sufficiently

flexible to produce firm and robust self-standing films under chloroform medium [72]. The single

gas permeability measurements on the freshly methanol-treated PIM-2 membranes demonstrated

CO2 and N2 permeabilities of 6600 Barrer and 460 Barrer, respectively [45]. Nevertheless, the gas

solubility of PIM-2 was relatively low as compared to the other PIMs with similar high FFV. This

may be attributed to the presence of fluorine constituents that enhanced the intra and interchain

interactions through halogen bonding within the structure, caused the low polymer

interconnectivity and weakened interaction between the functional groups and the penetrant gas,

being another reason of less research interest on these PIMs for CO2 separation application.

Various rigid and bulky intrinsic microporosity groups such as iptycenes [33, 53, 54, 73],

spirobifluorenes [74], cardo [75], and fused benzo groups [53] have been introduced into the PIM

backbones to increase the rigidity of polymer chain. In addition, the rigidity of the PIM polymer

chain can also be enhanced by thermal rearrangement [76]. The existence of these bulky groups

increases the rigidity of the polymer chains, which allows better control over the pore size

distribution with a reduction in d-spacing [77]. Hence, the gas permeability of the membrane is

decreased with an increasing gas-pair selectivity [25]. Furthermore, the rigid molecular structures

improve the mechanical properties of the polymer chain, Tg and chemical resistance [25].

14
A systematic investigation by Longo et al. has shown the reduction in d-spacing of the PIM-1

molecular structure upon long-term usage at ambient temperature, followed by accelerating aging

at elevated temperatures [76]. The stiffened PIM chains collapsed, reducing the free volume, and

hence decreasing CO2 permeability. Many studies have revealed different methodologies for

structural designing the polymer backbone to enhance the separation performance of PIMs

membranes through restriction of the polymer chain movement and a comparison results is listed

in Table 1.

Iptycenes, such as triptycene and pentiptycene, with their super rigid and bulky properties, deliver

a path to foster the intrinsic microcavities in similar size distribution and control the free volume

distribution along the molecular structure of PIMs [33, 53, 54, 73]. Bezzu et al. reported on the

dioxin-forming aromatic nucleophilic substitution reaction among the newly designed monomer

Spirobisindane-triptycene (SBI-Trip) to form a new copolymer PIM-1/SBI-Trip [53]. Derived

from spirobisindane with a pair of triptycene bulky groups, PIM-1/SBI-Trip has enhanced rigidity

and polymer chain spacing. In addition, the copolymer in its pure state exhibits CO2 permeability

of > 24000 Barrer, as compared to PIM-1 with only 9650 Barrer. Similarly, Gandara et al.

synthesized a series of benzotriptycene-based PIMs, for example, PIM-TMN-Trip, PIM-HMI-Trip,

PIM-TFM-BTrip, and PIM-DTFM-BTrip which demonstrated high CO2 permeability up to 52800

Barrer [33]. All of the benzotriptycene-PIMs performed better CO2 separation than the original

PIM-Btrip benefiting from the introduction of different solubilizing groups, including

tetramethylnaphthalene (TMN), hexamethylindane (HMI), and trifluoromethyl (TFM), into the

triptycene groups. The excellent properties of ultrapermeable PIM-TMN-Trip for CO2/CH4

separation were further modelled by Stanovsky et al. CO2 treatment at 5 bar to PIM-TMN-Trip

membranes strongly elevated the ideal selectivity of CO2/CH4 from 15 at normal pressure to

15
around 37 without significant reduction in CO2 gas permeability, surpassing the 2019 Robeson

upper bound [73]. Corrado et al. reported on the Diels-Alder cycloaddition reaction to fabricate

pentiptycene-based PIMs (Figure 4a) [54]. The incorporation of 17% of pentiptycene unit into the

PIM-1 backbone demonstrated a great performance enhancement. The resulting copolymers were

classified based on S or C shape configuration and normal (np) or isopropoxy (ip) substituent.

PPIM-ip-C exhibited the highest FFV of 25.2%, contributed by the most tortuous C-shaped

backbone configuration of the series, hence leading to the highest permeability among the

pentiptycene-based PIMs. Interestingly, the C-shape configuration (e.g., PPIM-ip-C) showed

aging-enhanced permeability without compromising the selecitivity. All pentiptycene-based PIMs

tested successfully maintained the permeability and selectivity towards CO2 exceeding the 2008

Robeson upper bound.

Sun et al. synthesized the novel Cardo-PIM-1 based on the polycondensation reaction of TTSBI,

TFTPN, and (9,9-bis(3,4-dihydroxyphenyl)fluorine (BDPF) with higher and larger intrinsic

microporous structure (e.g., 0.7-0.8 nm) in comparison to the traditional PIM-1 [75]. The CO2

permeability of Cardo-PIM-1 at room temperature was above 10000 Barrer with an ideal CO2/N2

selectivity of 14.6. In another study, Felemban et al. synthesized nine different

tetraoxidethianthrene-based (TOT) PIMs derived from different bis-catechol monomers (Figure

4b) [74]. The pristine PIM-Cardo and PIM-Cardo-TOT exhibited CO2 permeability of 3110 and

1026 Barrer with selectivities of CO2/N2 and CO2/CH4 up to 23 and 19.6, respectively. In addition,

the PIM-TOT-SBF-5 membrane was synthesized by incorporating t-butyl to the spirobifluorenes

bis-catechol monomer and had a CO2 permeability of 11400 Barrer. The presence of a more rigid

spirobifluorenes unit reduced the packing efficiency and diffusion activation energy, which

contributed to higher permeability as compared to the other PIM-TOT polymers. Meanwhile,

16
Bezzu et al. investigated the effect of several substituents such as methyl, t-butyl, and fuzed benzo

groups, onto spirobifluorene-based PIMs (PIM-SBFs) [51]. According to their study, the

incorporation of t-butyl in PIM-SBF-5 broadened the pore size distribution with a greater

concentration of micropores 1 nm, which allowed high CO2 permeability of 10000 Barrer over

1439 days of aging. Kang et al. modified TTSBI with pyrrolidinated methyl and 4-

methylpiperidinated methyl via Mannich reaction and subsequently prepared PIM-Py and PIM-

MePi membranes to study the influence of the size of side groups on gas separation performance

[78]. However, it was revealed that the bulky side groups mainly filled the larger microporous

region, lowering the FFV of the polymer matrix, subsequently reducing the gas permeability

without a significant enhancement on the selectivity (i.e., improvement of CO2/CH4 and CO2/N2

selectivities less than 10%).

Various monomers have been used in synthesizing new PIMs. Dibenzomethanopentacene (DBMP)

consists of methylene bridge and large rigid cleft which provides intermolecular free volumes,

similar to triptycene [79]. Chen et al. prepared a DBMP-based monomer via a Diels-Alder reaction

of dimethyloxyantracene and norbornadiene with a ratio of 2:1 [79]. The three-step reaction of

DBMP monomer with TFTPN and spirobisindane resulted in PIM-DBMPx (e.g., x ranging from

0 to 1) (Figure 4c). PIM-DBMP0.5 membrane exhibited the best CO2 permeability up to 22200

Barrer. It maintained its permeability above 9500 Barrer after ~1400 days of aging. The copolymer

PIM-DBMP0.5 provided excellent and consistent performance for CO2 permeability, CO2/N2 and

CO2/CH4 selectivity. Devarajan et al. synthesized a range of topologically different PIM-Py

membranes derived from TTSBI and 2,3,5,6-tetrafluoro-4-pyridinecarbonitrile (TFPCN)

monomers under different solvents and reaction temperatures [80]. The polymerization processes

proceeded with the occurrence of mono-substitution chain branching that induced the formation

17
of colloidal networks. The higher occurrence of the colloidal network of PIM-Py greatly

influenced the topological balance (e.g., structural and functional groups) within the polymer

sample, which affected the gas separation performance. The PIM-Py film with 50% colloidal

network content surpassed the 2019 CO2/N2 upper bound with CO2 permeability of 5270 Barrer

and CO2/N2 selectivity of 49.

Zhou et al. developed a series of poly(phenyl-alkane)s PIM (PIM-xR) through methanesulfonic

acid-catalyzed Friedel-Crafts hydroxyalkylation polycondensation [81]. The fluorinated PIM-xR

(PIM-xF) membranes were transparent, flexible, and easily synthesized by solution processing.

The satisfactory mechanical properties were attributed to the good interactions between PIM-xF

chains through fluorobenzene-benzene interactions and hydrogen bonds. Most of the PIM-xF

samples exhibited greater thermal stability as compared to the pristine poly(phenyl-alkane)s PIM

with the initial thermal decomposition temperatures above 400 °C under N2 atmosphere. Among

all membranes, the best performing PIM-5F membrane exceeded the 2008 Robeson bound with

CO2 permeability of 3240 Barrer and CO2/N2 selectivity of 27.9. Correspondingly, Weng et al.

carried out a Friedel-crafts polycondensation reaction to synthesize bi-functionalized PIM from

nidrobenzaldehyde (NBA) and tetramethylspirobisindane-6,6’-diol (HSBI) [82]. The generated

HSBI-NO2 polymer showed good solubility in most conventional solvents, such as tetrahydrofuran

(THF), dimethyl sulfoxide (DMSO), and N-methylpyrrolidone (NMP). Furthermore, the high

rigidity of the HSBI building block, induced by the nitro and hydroxide groups, led to a high Tg

above 300 °C. The HSBI-NO2 membrane showed CO2 permeability and CO2/CH4 selectivity of

140 Barrer and 20, respectively, with excellent CO2 permeability stability up to 50 days of aging.

Kammakakam et al. developed PIM-polyimide ionene by attaching imidazolium cation to

spirobisindane-based PIM [83]. The bisimidazole monomer (Im-SB, IV) was initially fabricated

18
before reacting with dichloro-xylene via Menshutkin reactions, followed by a reaction with

bistriflimide anions to yield PIM-polyimide ionene. With high solubility in common polar aprotic

solvents, PIM-polyimide ionene demonstrated a CO2 permeability of 356 Barrer and high CO2

selectivity towards N2 and CH4 with the value of 46 and 55, respectively, surpassing the 2008

Robeson line. The new ionene also showed no reduction in gas separation performance over 365

days.

Tröger’s base polymers are another typical ladder-type building block with intrinsic microporosity

and robust bridged bicyclic backbone. Tröger’s base heterocycle with a twist angle of 117o is found

in a zigzag configuration with high fractional volume and excellent stability in acid or base

conditions [84-86]. In 2013, Carta et al. applied Tröger’s base polymers in gas separation whereby

the membranes showed an exceptional performance in molecular sieving and gas solubility [84].

Later on, different advancements in Tröger’s base-based PIM continue to burgeon. To enhance

polymer matrix rigidity, Vopicka et al. fused benzene rings to the bicyclic and bridge-like Tröger’s

base and ethanoanthracene in synthesizing PIM-EA-TB [56]. The methyl groups at the bridge-

head of both ethanoanthracene and Tröger’s base units increased the spaces between polymer

chains and the interconnection between pores, leading to a higher intrinsic microporosity with an

average FFV of 0.276 [56, 87]. In addition, the PIM-EA-TB matrix showed higher CO2 affinity as

compared to the pristine PIM-1. A similar enhancement in tri(amino)phenylbenzene-based TB-

PIMs was reported by Antonangelo et al. [88]. Williams et al. synthesized a highly robust and gas-

selective PIM-MP-TB from methanopentacene (MP) via double Diels-Alder addition (Figure 4d)

[89]. The rigidity of MP as compared to the other monomers such as triptycene, ethanoanthracene

and Tröger’s base, was confirmed by its sharper dihedral angles. The NLDFT calculation on PIM-

MP-TB showed an approximately low pore volume, with the major pore size distributed below 5

19
Å. The PIM-MP-TB membrane showed a CO2 permeability of 3500 Barrer and a CO2/N2

selectivity of 17.5.

a) c)

b)

d)

Figure 4. Schematic of (a) pentiptycene-based PIMs [54], (b) the synthesis of PIM-TOT and

examples of bis-catechols monomers [74], (c) PIM-DBMPx [79], and (d) methanopentacene [89].

Copyright 2018, 2019, 2021 and 2023. Reproduced with permission from National Academy of

Science, Royal Society of Chemistry and Wiley-VCH.

Table 1. Restructured ladder PIMs with different contortion units for enhanced gas separation

performance.

Mw (x103
g/mol)/ T/P Permeability (Barrer)b Selectivity
Membrane ID a Monomers Ref.
PDI (°C/atm)
(Mw/Mn) H2 O2 N2 CH4 CO2 O2/N2 CO2/N2 CO2/CH4
PIM-1/SBI-Tripi SBI-Trip, TFTPN, 225 / 2.1 25/1 8360 4230 1298 3519 24410 3.3 18.8 9.7 [53]
SBI
(Aged 401 days) 4500 1500 338 430 7700 4.4 22.8 17.9 [53]

TMN-Trip i TMN, Trip 52.3 / 3.8 25/1 18800 10400 3540 7250 52800 2.9 14.9 7.3 [33]

HMI-Trip i HMI, Trip 61.3 / 2.4 25/1 16600 8540 2560 4870 44200 3.3 17.3 9.1 [33]

TFM-Btripi TFM, Btrip - 25/1 13600 6210 1830 2280 33700 3.4 18.4 14.8 [33]
DTFM-Btripi TFM, Btrip - 25/1 14700 7770 3000 4340 42600 2.6 14.2 9.8 [33]

PIM-TMN-Trip c,i TMN, Trip - 25/5 - - - 733 11000 - - 15.0 [73]

(Aged 14 days) 25/5 - - - 271 9500 - - 35.0 [73]

20
PPIM-ip-C (Aged 2,3-dimethoxy - 25/7 - - - 138 2750 - - 20.0 [54]
225 days)d,i anthacene, TTSBI,
PPIM-np-S TFTPN 25/7 - - - 164 1970 - - 12.0 [54]
(Aged 10 days)d,i
Cardo PIM-1e TTSBI, TFTPN, 68.9 / 1.7 25/1 - - 890 - 13000 - 14.6 - [75]
BDPF
PIM-TOT-SBF-5i TOT, t-butyl SBF - 1/25 4160 1960 717 1430 11400 2.7 15.9 8.0 [74]

PIM-Cardoi - - 730 186 46 52 1026 4.0 22.3 19.6 [74]

PIM-Cardo-TOTi Cardo, TOT - 1520 490 135 187 3110 3.6 23.0 16.6 [74]

PIM-SBF-2 SBF-based 95 / 2.3 25/3 9160 3820 1150 2020 22300 3.2 19.4 11.0 [51]
monomers, TFTPN
(Aged 1295 days) 25/3 4240 910 166 184 3870 5.5 23.3 21.0 [51]
PIM-SBF-5 110 / 2.4 25/3 5590 2750 1080 2480 16400 2.6 15.2 6.6 [51]
PIM-SBF-5 f 25/3 - - - 90 10000 - - 111.0 [51]

(Aged 1439 days) 25/3 4710 1870 550 925 10000 3.4 18.0 10.8 [51]
PIM-Py TTSBI-Py, TFTPN - 35/6.8 557 127 37 52 769 3.5 20.9 14.7 [78]
PIM-MePi TTSBI-MePi, - 35/6.8 273 53 4 20 295 3.9 21.8 15.0 [78]
TFTPN
PIM-DBMP0.5i TTSBI, TFTPN, 65 / 2.3 25/1 8970 4080 1210 1840 22200 3.4 18.3 12.1 [79]
DBMP monomers
PIM-DBMP0.5f,i 25/1 - - - 400 8000 - - 20.0 [79]

PIM-DBMP0.5 g,i 25/1 - - 400 - 9000 - 23.0 - [79]

PIM-Py TTSBI, TFPCN 490 / 0.4 25/2 - - 108 223 5270 - 48.8 23.6 [80]

PIM-5F i DBS, fluorinated 270 / 3 35/6.8 - - 116 161 3240 - 27.9 20.1 [81]
benzaldehyde

HSBI-NO2 HSBI, NBA 388 / 4.11 35/5 196 28 6 7 144 4.9 - 20.0 [82]
(Aged 50 days) 35/5 166 24 5 6 130 5.2 - 21.1 [82]
HSBI-NO2h 35/5 - - - 3 69 - - 23.0 [82]

PIM-Polyimide Im-SB, Dichloro-p- - 20/3 111 - 7 6 356 - 46.4 54.9 [83]


Ionenei xylene
Trip(Me2)-TBi dimethylanthracene, 118 / 2.7 25/1 5446 1002 255 347 3718 3.9 14.6 10.7 [87]
anthranilic acid
(Aged 407 days) 25/1 2625 500 96 156 1880 5.2 19.6 12.1 [87]
EA(H2)-TB i EA, 62 / 2.3 25/1 6088 1673 358 458 6097 4.6 17.0 13.3 [87]
dichloroethylene
(Aged 119 days) 25/1 4066 902 188 196 2999 4.8 16.0 15.3 [87]
PIM-MP-TB i MP, DMM - 25/1 4050 999 200 3500 264 5.0 17.5 13.3 [89]

PIM-MP-TB- 25/1 2790 648 125 2340 158 5.2 18.7 14.8 [89]
140Ci
(Aged 118 days) 25/1 1120 134 21 633 26 6.3 29.6 24.4 [89]
(Aged 370 days) 25/1 812 88 13 378 14 6.7 28.4 25.7 [89]
a BDPF: (9,9-bis(3,4-dihydroxyphenyl)fluorine; BTrip: benzotriptycene; DBMP: dibenzomethanopentacene; DBS: dimethoxyspirobisindane;
DMM: dimethoxymethane; EA: ethanoanthracene; HMI: hexamethylindane; HSBI: tetramethylspirobisindane-6,6’-diol; Im-SB: bisimidazole; MP:
methanopentacene; NBA: nitrobenzaldehyde; SBF: spirobifluorene; SBI: spirobisindane; TFM: trifluoromethyl; TFTPN:
tetrafluoroterepthalonitrile; TMN: tetramethylnaphthalene; TOT: tetraoxidethianthrene; Trip: triptycene; TTSBI: 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-
tetramethyl-1,1′-spirobisindane. b 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. c Mixed gas composition CO2:CH4 40:60. d Mixed gas composition
CO2:CH4 20:80. e Mixed gas composition CO2:N2 50:50. f Mixed gas composition CO2:CH4 35:65. g Mixed gas composition CO2:N2 15:85. h Mixed
gas composition CO2:CH4 50:50. i Methanol treated membranes

3.2. Post-modification on PIMs polymer structure

Apart from polymer synthesis, the post-modification of PIM structure demonstrates significant

potential for improving their properties. PIMs functionalization has been studied on the conversion

21
of nitrile (–CN) groups and the hydrogen groups of the benzene ring in spirobisindane moieties.

The bulky side groups occupy the polymer micropores and reduce the pore size and free volume

in the matrix. This led to trade-off of permeability and selectivity according to the principle of

molecular sieving in membrane construction. Furthermore, the incorporated functional groups also

increase the molecular diffusivity and solubility towards the polymer matrix due to chemical

bonding or high affinity between the two constituents. Meanwhile, the change in polymer

architecture by thermal or chemical treatment also controls the polymer microstructure through

free volume manipulation. A summary of the gas separation performance of membranes prepared

with post-synthetic modification on PIMs polymer structure is listed in Table 2.

Several studies reported amidoxime (AO)-functionalized PIM by replacing nitrile groups in PIM

with AO substituents via reactions with hydroxylamine [68, 79, 90]. The AO-functionalized PIMs

exhibited highly pronounced CO2/N2 and CO2/CH4 selectivities but a lower CO2 permeability, due

to the greater inter-chain cohesion induced by extensive hydrogen bonding between AO groups.

To improve the permeability of AO-PIM-1, a substitution of adamantane onto AO-PIM-1 via acyl

chloride-substitution reaction can be applied by increasing the interchain spacing of the polymer

matrix [68]. Correspondingly, a similar functionalization method has been recently applied to

modify TPE-PIM (tetraphenylethylene-based ladder PIM) into AO-TPE [91, 92]. The pure gas

permeation test on fresh AO-TPE revealed excellent CO2/CH4 and CO2/N2 selectivities of

approximately 40 and 39, respectively. In comparison with the pristine polymer, AO-TPE

displayed reduced pore size, free volume, and surface area, but with an enhanced microporosity,

which yielded a higher selectivity. Furthermore, the abundant –NH2 and –OH groups retarded

CO2-induced plasticization due to a more rigid polymer matrix and more extensive inter-chain

bonding.

22
Ionic liquids (ILs) have been widely studied as non-volatile and reversible CO2 absorbents as well

as alternative compounds for CO2 separation. This is owing to their high CO2 solubility, attributed

to the tunable interactions between cation, anion, and other substituents in the ILs with CO2 [49,

93]. Guiver et al. reported the incorporation of pseudo-ionic liquid to PIM-1 by side group

modification [49]. The PIM-1 synthesized in NMP was first modified with ZnCl2 and NaN3 to

obtain tetrazole-containing PIM (TZPIM-50), followed by further reaction with different tertiary-

amine in methanol to form an ionic bond between the nitrogen in the tetrazole and amine groups.

The fabricated TZPIM-50 membrane achieved a CO2 permeability of 5070 Barrer and an ideal

selectivity of CO2/N2 of 26.4. Rodriguez et al. presented an in-situ cross-linking and solid-state

deprotection method in amine-functionalized PIM to increase its FFV distribution as well as

sorption selectivity [94]. PIM-NH2 was initially protected with t-BOC via solid-state reaction

before deprotected under thermal and acid-catalyzed conditions to form PIM-deBOC(thermal) and

PIM-deBOC(acid). Remarkably, the CO2/N2 and CO2/CH4 selectivities for PIM-deBOC(thermal)

increased by >200%. Li et al. synthesized OH-functionalized Tröger’s base polymers, PIM-HNTB,

constructed from a tertbutyldimethylsilyl (TBS) protected 1,1’-binaphthalene-diamine monomer

(NH2-BINOL-TBS) with abundant site of contortion [95]. By adjusting the retention time, the

PIM-HNTB-30 and PIM-HNTB-60 were obtained with OH concentrations of 30% and 60%,

respectively. Among the tested membranes, these two membranes showed the highest gas

permeation performance ascribed to their good thermal stability, excellent rigidity, and amorphous

structure with pore sizes ranging from 6.0-6.5 Å.

Studies have been conducted to convert the nitrile group to carboxyl groups (–COOH) in PIMs by

undergoing acid or base hydrolysis to generate carboxylated or hydrolyzed PIMs. During base

hydrolysis, the strong base, OH- acts to attack the carbon of cyano (–C=N), forming negatively

23
charged ions which then capture the proton in the amide group and hydrolyze to carboxylic acid.

Meanwhile, an acidic environment acts as a driving force for nucleophilic attack by retaining the

protonated intermediate and releasing the more stable NH3. Diving into the molecular structure

perspective, the large –COOH group has a stronger polarity and affinity towards CO2 than –CN

groups. The easy dissociation of O-H bond in –COOH groups and the presence of unshared

electron pairs of O atoms contributed to the strong Lewis acid-base reaction. In addition, the H

atom presence in –COOH groups and O atom in CO2 promotes a strong hydrogen bond,

contributing to the improvement of CO2 adsorption selectivity [96].

The idea of carboxylating PIM-1 via hydrolysis (referred to as cPIM-1) of nitrile groups was first

introduced by Du et al. to enhance the gas pair selectivity [97]. Later, Satilmis and Budd found

that a completed base-catalyzed hydrolysis of PIM-1 (referred to as hPIM-1) yielded

indistinguishable products such as amide and carboxylate structures [98]. Nevertheless,

carboxylated or hydrolyzed PIMs membranes were found to be fragile which limits their

implementation. To overcome this hurdle, Yong and Chung reported a modified base-catalyzed

hydrolysis strategy to produce mechanically strong and flexible hPIM-1 [99]. It was revealed that

hPIM-1 membrane hydrolyzed from high molecular weight PIM-1 showed better flexibility and

tensile properties.

Rodriguez et al. reported on the facile carboxylic acid-functionalized PIM-1 with >89% conversion

via acid hydrolysis method involving both acetic acid and sulfuric acid [100]. The wide-angle X-

ray scattering (WAXS) measurements showed an overall reduction in chain spacing and free

volume of cPIM-1 due to the contraction of the polymer matrix through the incorporation of

carboxylic acid moieties. The cPIM-1 membrane had a 260% and 160% increase in CO2/CH4 and

CO2/N2 selectivities, respectively, with the CO2 permeability of approximately 2900 Barrer.

24
Despite the tight packing structure, cPIM-1 underwent plasticization at the pressure of 5 atm,

which is three times lower than the pristine PIM-1 of 15 atm. As the pressure and CO2

concentration elevated, the hydrogen bonding between the polymer chain and –COOH groups was

disturbed, leading to an increase in FFV and higher CO2 diffusion. Han et al. tailored the properties

of PIM-1 through simultaneous hydrolysis and crosslinking of nitrile groups using methane

sulfonic acid (MSA) [101]. The penetration of MSA molecules into PIM-1 membranes enhanced

the protonation of the nitrile group and facilitated the polymer chain movement and membrane

swelling. The successful cross-linking and hydrolysis reactions were attributed to the hydrogen

bonding interactions as shown in XRD analysis and the changes in solubility properties. The

formation of stable and bulky triazine rings after cross-linking of the nitrile groups led to the

insolubility of cPIM-1 in organic solvent such as dichloromethane (DCM). Peng et al. employed

hydrothermal method that achieved 90% carboxylation efficiency within 10 h in synthesizing

cPIM-1 [96]. The gel permeation colorimetry (GPC) test showed a 45.8% reduction in PDI of

cPIM-1 compared to PIM-1 (i.e., 2.53 to 1.37), indicating that the hydrothermal process would

significantly reduce the chain length while improving its homogeneity. However, it was found that

cPIM-1 demonstrated a remarkable weight loss at 300 °C, possibly due to the decarboxylation of

carboxylic groups in cPIM-1. The cPIM-1 membrane showed an improvement in CO2/N2 and

CO2/CH4 selectivities from 20 and 14 to 34 and 35, respectively. In comparison, the volume of –

COOH is larger than –CN, which increased the spatial site resistance of cPIM-1 for chain

movement and narrowed down the available pore channels for gas transport. Ma et al. synthesized

a tetraphenylethylene-based ladder network (MP1) and its COOH-functionalized analogue (MP2)

by polycondensation reaction from 4,4’,4’’,4’’’-(ethene-1,1,2,2-tetrayl)tetrakis(benzene-1,2-diol)

and TFTPN [77]. Despite the smaller surface area, BET analysis showed that the functionalized

25
MP2 exhibited a more uniform pore size distribution. A similar observation of the decomposition

of –COOH groups was found when reducing the decomposition temperatures of cPIMs (e.g., from

450 °C reduced to 320 °C).

Additionally, nitrous acid was also used in hydrolyze amide-containing PIM-1 (PIM-CONH2) to

cPIM-1 [102]. Ling et al. reported on the post-synthesis metallation of cPIM-1 using Na+, Mg2+

and Al3+ [103]. The Na+ metallated PIM (cPIM-Na) was prepared by reacting the cPIM-1 with

excess NaOH in a methanol solution. Meanwhile, the multivalent functionalized PIMs were

prepared via the cation exchange method from cPIM-Na. The metallation process was found to

significantly enhance the CO2 adsorption as well as its selectivity towards CH4 and N2.

Meanwhile, several studies have proposed the PIMs functionalization on the hydrogen groups of

benzene ring in spirobisindane moieties with bulky side groups such as methyl, bromomethyl (–

CH2Br) [104, 105], vinyl (–CH=CH2), brominated vinyl (–CHBrCH2Br) and thiophenated vinyl

(–CH2CH2SPh) [106]. Notably, Chen et al. designed bromoalkylated PIM (PIM-BM-x) by reacting

PIM-1 with different ratios of N-bromosuccinimide (NBS) and azodiisobutyronitrile (AIBN) to

tune its gas separation performance [104]. The PIM-BM-100 membrane (i.e., 100% brominated

PIM-1) achieved CO2/CH4 selectivity of 25.8, exceeding the 2008 Robeson upper bound.

Furthermore, they reported the self-cross-linking of PIM-BM-70 under temperatures of 200-

300 °C. The cross-linked PIM-BM-70 membranes demonstrated an excellent plasticization

resistance up to approximately 500 Psi, which is five times higher than the pristine PIM-BM-70.

Corresponding to PIMs containing benzodioxane backbones with diverse set of functionalities,

Rodriguez et al. studied the effect of competitive sorption phenomenon during mixed-gas

permselectivities [107]. Low pressure mixed-gas tests at 1-2 atm and 35 °C were carried out for

PIM-tBOC, PIM-COOH, PIM-1, PIM-deBOC (thermal), PIM-deBOC (acid), and PIM-NH2,

26
whereby results showed that there was an increasing order in sorption coefficients at infinite

dilution. The amine-functionalized PIM-NH2 showed the greatest improvement of 140% and 250%

in CO2/CH4 and CO2/N2 selectivities, respectively, as compared to pure gas results performed at

constant CO2 pressure. The study demonstrated that the high solubility selectivity of PIM-NH2

played a dominant role in leveraging competitive sorption and improving selectivity for CO2-based

gas pairs compared to the other functionalized polymers containing –OH and –COOH groups. In

addition, the high pressure mixed gas test proved the excellent plasticization resistance for PIM-

NH2, attributed to the more rigid matrix induced by interchain hydrogen bonding.

27
Table 2. Post-synthetic modification on PIMs polymer and its membrane gas transport properties.

Post-modification Mw (x103 g/mol)/ T/P Permeability (Barrer)b Selectivity


Membrane IDa PIMs monomer Ref.
method PDI (Mw/Mn) (°C/atm) H2 O2 N2 CH4 CO2 O2/N2 CO2/N2 CO2/CH4
AOPIM-1c TTSBI, TFTPN Amidoxime- - 35/1 575 130 27 29 981 4.8 36.0 33.8
[68]
(PIM-1) functionalized
AOPIM-1-9%Adamantane c AOPIM-1 Adamantane- - 35/1 1360 367 80 83 2484 4.6 31.2 30.1
[68]
grafted
AO-TPE TPE-PIM Amidoxime- - 35/2 137 14 3 3 101 5.2 38.8 39.8 [92]
(Aged 150 days) functionalized 35/2 113 11 2 3 81 5.2 39.5 45.5 [92]
PIM-DBMP-AOc TTSBI, TFTPN, Amidoxime- - 25/1 4530 691 106 106 3520 6.5 33.2 33.2 [79]
PIM-DBMP-AOc,d DBMP monomers functionalized 25/1 - - - 30 1300 - - 43.3 [79]
PIM-DBMP-AOc,e 25/1 - - 40 - 1600 - 40.0 - [79]
PIM-BM-70c methylated Bromomethylated 56.7 / 4.3 25/3.4 - 322 70 74 1689 4.6 24.7 22.8 [104]
PIM-BM-100c TTSBI, TFTPN PIM-M 42.7 / 3.5 25/3.4 - 228 41 43 1110 5.6 27.1 25.8 [104]
PIM-BM-70-200°C-10hc (PIM-M) Self-crosslinked 56.7 / 4.3 25/3.4 - 76 15 17 508 5.0 33.9 29.9 [104]
PIM-BM-70-250°C-5hc Bromomethylated 25/3.4 - 53 10 10 272 5.1 26.1 26.9 [104]
PIM-BM-70-250°C-10hc PIM-M 25/3.4 - 27 5 4 142 5.7 30.2 40.6 [104]
PIM-BM-70-300°C-5hc 25/3.4 - 11 2 1 48 6.5 30.1 43.9 [104]
PIM-HNTB-30c DMM, NH2- OH- 551 / 1.67 35/2 639 204 66 139 967 3.1 14.7 7.0 [95]
(Aged 250 days)c BINOL-TBS functionalization 35/2 406 102 30 58 481 3.4 16.0 8.3 [95]
PIM-HNTB-30-350°C-0.5hc 35/2 364 103 27 45 516 3.8 19.1 11.4 [95]
PIM-HNTB-60c 130 / 1.75 35/2 629 150 40 71 833 3.8 20.8 11.8 [95]
(Aged 180 days)c 35/2 280 65 17 33 341 3.8 20.1 10.4 [95]
PIM-HNTB-60-350°C-0.5hc 35/2 394 89 20 34 352 4.5 17.6 10.4 [95]
TZPIM-50-MeOH TTSBI, TFTPN NaN3/ZnCl2 86 / 2 25/6.8 - 752 192 - 5070 3.9 26.4 -
[49]
(PIM-1)
PIM-IL1 TZPIM-50 Convertion to 25/6.8 - 169 36 - 1043 4.7 29.0 - [49]
PIM-IL2 pseudo-ionic liquid 25/6.8 - 151 35 - 1010 4.3 29.0 - [49]
PIM-IL3 tetrazole-like 25/6.8 - 102 23 - 817 4.5 35.5 -
[49]
structures
PIM-NH2 TTSBI, TFTPN Amine- - 35/1 1450 430 134 210 840 3.2 6.3 4.0
[94]
(PIM-1) functionalized
PIM-t-BOC PIM-NH2, t-BOC - - 35/1 130 19 4 5 100 4.8 25.0 20.0 [94]
PIM-deBOC(acid) Acid-catalyzed - 35/1 1700 500 160 260 1400 3.1 8.8 5.4
[94]
deprotection
PIM-deBOC(thermal) Thermal - 35/1 2000 500 120 170 2300 4.2 19.2 13.5
[94]
deprotection
TTSBI, TFTPN Acid hydrolysis -
cPIM-1 35/1 270 44 9 9 290 4.9 32.2 32.2 [100]
(PIM-1)
TTSBI, TFTPN Acid hydrolysis -
cPIM-1 25/2 3185 284 107 135 3739 2.7 27.7 34.9 [101]
(PIM-1)
cPIM-1 TTSBI, TFTPN Acid hydrolysis 35.7 / 1.37 25/1 500 65 10 10 360 6.3 34.3 35.8 [96]
(PIM-1)
PIM-NH2c,f TTSBI, TFTPN Amine- - 35/3.2 - - - 300 1070 - - 3.6 [107]
(PIM-1) functionalized &
MeOH treatment
(Aged 448 days)c,f 35/3.2 - - - 44 480 - - 11.1 [107]

28
PIM-NH2 c,f,g TTSBI, TFTPN Amine- - 35/2.4 - - - 33.5 845 - - 25.2 [107]
(PIM-1) functionalized &
MeOH treatment
(Aged 448 days)c,f,g - - - 18 509 - - 27.8 [107]
a DAT: 2,6-diaminotriptycene; DATCA: 2,6-diaminotriptycene-14-carboxylic acid; DBMP: dibenzomethanopentacene; DMM: dimethoxymethane; NaN3: sodium azide; NH2-BINOL:1,1'-binaphthalene-
diamine; –t-BOC: tert-butoxycarbonyl; TBS: tertbutyldimethylsilyl; TFTPN: tetrafluoroterepthalonitrile; TPE-PIM: Tetraphenylethylene-based ladder PIM; TTSBI: 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-
tetramethyl-1,1′-spirobisindane; ZnCl2: zinc chloride. b 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. c Methanol treated membranes. d Mixed gas composition CO2:CH4 35:65. e Mixed gas composition
CO2:N2 15:85. f Mixed-gas tests performed at a 1.6 atm partial pressure of CO2. g Mixed gas composition CO2:CH4 50:50.

29
3.3. Polymer blends or copolymerization with PIMs

Polymer blends have been recognized as one of the most time-efficient and low-cost approaches

to enhance PIM membrane gas separation performances for their simple methodologies [19, 108,

109]. In general, pristine PIM membrane has high CO2 permeability, but with low to moderate

selectivity. The concept of the PIM polymer blend is to combine the advantages of PIM with

different materials into a new dual-polymer system with synergetic and improved properties,

including gas permeability enhancement [110, 111], selectivity improvement [112, 113], and anti-

plasticization characteristics [114]. One of the straightforward methods to determine the

miscibility of the blend is through optical inspections [108, 112]. Besides, Tg analysis is commonly

used to determine the miscibility of a blend. Generally, the homogeneous blending system can be

obtained when there is only one Tg evaluated [109, 115]. Studies have shown that blending PIM

with polyimide increased the CO2 selectivity against N2 or CH4, followed by mechanical properties

improvements [20, 108]. Table 3 summarizes the gas separation performance of recently

developed PIMs-based blend membranes.

Poly-phosphazene (MEEP80) is an amorphous polymer composed of a phosphazene backbone

subjected to 80 mol% substitution with MEE ether group. In general, polyphosphazenes possess

larger densities as compared to PIM-1 due to their rubbery and packed structure. Sekizkardes et

al. reported excellent miscibility of the PIM-1 blending with ether-functionalized MEEP80 [114].

The PIM-1-based blend system was prepared using a series of annealing, compression, and

relaxation steps based on the molecular dynamics (MD) simulation developed by Larsen et al.

[116]. The blend of PIM-1 and flexible MEEP80 polyphosphazene showed a significant

improvement in toughness and better solubility in solvents such as chloroform and THF. The

composite PIM-1/25%MEEP80 and PIM-1/50%MEEP80 exceeded the 2008 Robeson upper

30
bound by showing a significant enhancement in CO2/N2 selectivity with the CO2 permeability of

2450 and 2000 Barrer respectively. In another work, Sekirkardes et al. examined the effect of ether

side chain concentration in the polyphosphazene by blending PIM-1 and MEEP0, MEEP80, and

MEEP100, respectively (Figure 5a) [19]. It was found that the higher concentrations of MEE in

polyphosphazenes improved the miscibility with PIM-1. Among these membranes, MEEP100 has

the smallest FFV of 12%, followed by 14% and 16% for MEEP80 and MEEP0, consecutively.

The lower free volume of MEEP100 was attributed to the replacement of bulky phenoxy groups

with methoxyethoxy groups. Furthermore, MEEP100 with an average pore size of 0.1 nm

exhibited the smallest tendency towards self-aggregation due to lower interaction between MEEP

molecules. In the 20% mol CO2/20% mol N2/Ar mixed gas permeation test, the PIM-

1/10%MEEP100 membrane showed a high CO2 permeability of 5340 Barrer and CO2/N2

selectivity of 24. The PIM-1/MEEP100 membrane with a higher concentration of MEE groups

contributed to a superior gas transport performance due to a higher CO2 affinity with pendant ether

groups than with the phenoxy group.

Zhao et al. blended PIM-1 with Tröger’s base polymer and the resultant membranes showed good

miscibility with improved physic-chemical characteristics and gas separation performance [117].

The Tröger’s base polymer exhibited a homogeneous morphological structure and excellent

miscibility with PIM-1 due to the interaction between the N atom and with –CN group in PIM-1.

Besides, notable improvements to the mechanical properties of PIM/TB were also observed with

approximately 25% and 67% increments in tensile strength and Young’s modulus as compared to

the pristine PIM-1. PIM/TB blending membrane (80:20) resulted in the highest gas separation

performance with CO2 permeability of 5765 Barrer and ideal selectivity values of 17.4 and 11.3

for CO2/N2 and CO2/CH4, respectively. With a solution casting method, Esposito et al. developed

31
PIM-EA(Me2)-TB with Matrimid 5218 polyimide membranes. PIM-EA(Me2)-TB consists of

Tröger’s base and ethanoanthracene, which form a rigid polymer backbone [110]. The SEM

analysis showed less than 1% of the separate domain area suggesting good miscibility between the

polymers. The PIM-EA-(H2)-TB/Matrimid 5218 membrane resulted in a 20-fold elevation in CO2

permeability compared to the original Matrimid 5218. The mixed gas permeability test on the

blended membrane resulted in CO2/N2 and CO2/CH4 selectivities of 43 and 27, respectively with

no significant performance change in the pressure of 1-7 bars. In the following years, they

fabricated PIM-EA-(Me2)-TB/Matrimid 5218 TFC membranes on different types of support [76].

Despite the reasonable CO2 permeability of 100 Barrer, the equimolar blend of PIM-EA-(Me2)-

TB/Matrimid 5218 on polyacrylonitrile (PAN) support reached a doubled selectivity of CO2/N2

compared to the neat PIM-EA-(Me2)-TB. Chen et al. reported the multi-covalent crosslinking of

the blended bromomethyl PIM (PIM-BM-70) and Tröger’s base membranes [118]. The PIM-

BM/TB blend membranes were heated to 300 oC to form a crosslinked PIM-BM/TB with tailorable

porosity and outstanding molecular sieving properties. The proposed cross-linking mechanism was

initiated by the formation of quaternary ammonium salts followed by alkylation, polymer chain

scission, and intra-structural rearrangement (Figure 5b). The increase in temperature as well as

reaction time led to a more substantial drop in gas permeability, with the tradeoff of higher

selectivities towards smaller-sized molecules, including H2, O2, and CO2. Almansour et al.

reported on a novel method to improve the physical aging of PIM-1 by blending it with different

loadings of colloidal network-rich bis(phenyl)fluorine-based Cardo-PIM-1 [119]. The PIM-1

membrane blended with 5 wt% Cardo-PIM-1 reached CO2 permeability of 15187 Barrer, and

CO2/CH4 selectivity of 17.4, overpassing that of the pure PIM-1 membrane. However, a higher

loading caused polymer rigidification and lower mobility, leading to poorer separation. The

32
membranes preserved around 65% of their initial permeability after 230 days of aging and

surpassed the 2008 Robeson upper bound with CO2/CH4 selectivity of 13. The existence of Cardo-

PIM-1 in the PIM-1 matrix increased both the diffusion and solubility coefficients, hence

enhancing the free volume and improving the gas separation performance.

In addition, polybenzimidazole (PBI) also has been known for its good mechanical properties and

film-forming ability. The excellent miscibility between PIMs and PBI due to the interaction

between benzimidazole and –CN moieties of PIM has been reported in previous studies [111, 115,

120]. Sanchez-Lainez et al. fabricated both dense and asymmetric membranes based on PIM-EA-

(H2)-TB/PBI at different proportions via the phase inversion method [115]. The resultant

membrane showed good miscibility with the presence of a single Tg. Wang et al. blended PBI with

bromomethylated PIM-Br to prepare gas separation membranes [111]. The thermal treatment of

PIM-Br/PBI membrane at 300 °C formed covalent cross-linking via Friedel-Crafts reaction of

bromomethyl, which was confirmed by the reduction of N+Br- peak intensity in the XPS result

(Figure 5c). After the thermal treatment, the CO2 permeability of PIM-Br/PBI membrane consisted

of 5 wt% PBI improved from 1645 to 3313 Barrer with the performance near to the 2008 Robeson

upper bound. Moreover, the cross-linked membrane achieved a comparable CO2/CH4 selectivity

of 13 with CO2 plasticization pressure of >40 atm.

To improve CO2 selectivity, Hou et al. blended different concentrations of poly(arylene fluorine

ether ketone) (Am-PAFEK) with PIM-1 [112]. The SEM and XRD results showed a morphology

change and reduced intensity of each pristine polymer after blending (Figure 5d,e). Furthermore,

the surface morphologies of the blend membranes were smooth with no aggregation, indicating

the good compatibility of PIM-1 and Am-PAFEK. The highest gas transport performance was

33
obtained by PIM-1/Am-PAFEK (80:20) with CO2 permeability of 752 Barrer and CO2/N2 and

CO2/CH4 selectivities of 26 and 24, respectively.

Chung’s group have reported a detailed investigation of PIMs-based blend membranes for gas

separation, including Matrimid [108, 109], Ultem [121], sulfonated polyphenylenesulfone (sPPSU)

[122]. To improve the miscibility of the blend, they modified the PIM-1 to cPIM-1. They

incorporated cPIM-1 with Torlon [123], 3,3′4,4′-benzophenone tetracarboxylic dianhydride

and 80% methylphenylene-diamine + 20% methylenediamine (BTDA-TDI/MDI) (P84) [124] and

Matrimid [20] to tailoring membranes, extending from gas separation to pervaporation. For

example, the polarized light microscopy (PLM) analysis demonstrated miscibility of PIM-1 in

PIM-1/polymer blends with PIM-1 content above 90% only [108], while cPIM-1 showed full

miscibility for cPIM-1 of all ranges from 5, 10, 30, 50, 70, 90, 95% [20]. Moreover, the

incorporation of sPPSU enhanced the CO2 selectivity significantly, owing to the strong hydrogen

bonds formed by sulfonic acid groups in the polymer matrix [122]. In addition, the PIM-1/sPPSU

membranes displayed improved anti-plasticization properties tests up to 30 atm under CO2/CH4

binary system. Meanwhile, the cPIM-1/Matrimid blend membranes demonstrated a significant

enhancement of plasticization pressure from <10 atm to 15 atm for small loading of cPIM-1 (e.g.,

5-10 wt%) [20]. The improvement of the plasticization resistance is attributed to the hydrogen

bonding formed after the incorporation cPIM-1. Aside from that, the blended membranes showed

comparable gas separation performance when compared to the pristine PIM-1, with a slight

reduction in CO2 permeability and an increase in selectivity against N2 and CH4.

Nevertheless, polymer blends of PIM with other polymers face the challenge of miscibility at the

molecular level, which can be addressed via copolymer polycondensation. Hossain et al.

34
synthesized random copolymers made of PIM-1 and Ellagic acid [(PIM-co-EA)x] through a facile

one-way polycondensation [125]. The film consisting of 1 mol% of planar hydrophilic ellagic acid

showed increasing in total free volume due to the disturbance of regular PIM-1 chain orientation,

hence enhancing the gas diffusivity. The (PIM-co-EA)1 exhibited excellent CO2 permeability of

5483 Barrer with CO2/N2 and CO2/CH4 selectivities of 23 and 18, respectively. Meanwhile, the

higher Ellagic acid concentration resulted in a more closely packed structure that led to poor gas

separation selectivity. A similar phenomenon was observed by Ponomarev et al. in their study

when copolymerized PIM-1 with 2,3,6,7-antharacenetetrayl moiety [126]. In another work,

Bengtson et al. employed both polymer blends and polycondensation methods to synthesize PIM-

1-PEG (PEG: polyethylene glycol) copolymers using PEG with different molecular weights (e.g.,

PEG-200, PEG-1000, and PEG-2000) [127]. The blend membranes were constructed by dissolving

varying amounts of PEG and PIM-1 in chloroform. Another synthesis method uses PEG-

substituted trihydroxy-benzoic compounds (THBA-PEG-OMe2) at different concentrations to

terminate the PIM-1 polycondensation reaction. All tested copolymers experienced reduced gas

permeability with a slight increase in selectivity. The improved gas separation performance of the

copolymers compared to the parent PIM-1 membrane corresponded to the preceding study on PIM-

1/PEG blend membranes by Wu et al. [128].

35
a) b)

c) d) e)

Figure 5. (a) Chemical structure of different MEEP and PIM-1 [19]. (b) PIM-BM/TB cross-

linking mechanism [118]. (c) XPS analysis for different cross-linked PIM-Br/PBI [111]. (d)

SEM and (e) XRD patterns of pristine and PIM-1/Am-PAFEK

membranes [112]. Copyright 2020 and 2021. Reproduced with permission from Elsevier and

Springer Nature.

Table 3. Gas separation performance of recently developed PIMs-based blend or copolymerized

membranes.

Blending T/P Permeability (Barrer)b Selectivity


Membrane IDa Ref.
Method (°C/atm) H2 O2 N2 CH4 CO2 O2/N2 CO2/N2 CO2/CH4
PIM-1/25%MEEP80 Physical 40/1 - - - - 2450 - 21.0 - [114]
PIM-1/50%MEEP80 blending 40/1 - - - - 2000 - 29.0 - [114]
PIM-1/10%MEEP80c Physical 22/1 - - - - 3100 - 24.0 - [19]
PIM-1/10%MEEP100c blending 22/1 - - - - 5200 - 24.0 - [19]
PIM/TB (80:20)d Physical 35/4 3489 1126 332 511 5765 3.4 17.4 11.3 [117]
PIM/TB (60:40)d blending 35/4 1905 538 144 212 2585 3.7 18.0 12.2 [117]
PIM/TB (40:60)d 35/4 1417 394 99 127 1824 4.0 18.4 14.4 [117]
PIM/TB (20:80)d 35/4 1032 183 38 45 727 4.8 19.1 16.2 [117]

36
PIM-EA(Me2)-TB/ Physical 25/1 328 41 7 9 198 6.0 29.0 21.6 [110]
Matrimid (50:50) blending
PIM-BM/TBd Physical 35/3.4 1925 423 110 112 2007 3.8 18.2 17.9 [118]
PIM-BM/TB 80°C- blending 35/3.4 1392 216 44 47 987 4.9 22.4 21.0 [118]
20hd
PIM-BM/TB 200°C- 35/3.4 721 88 18 12 391 4.9 21.7 32.6 [118]
20hd
PIM-BM/TB 250°C- 35/3.4 582 104 25 16 431 4.2 17.2 26.9 [118]
5hd
PIM-BM/TB 250°C- 35/3.4 356 33 4 1.8 149 7.5 33.9 82.7 [118]
20hd
1.0Cardo-PIM-1 Physical 25/2 - - - 1510 12628 - - 8.4 [119]
2.0Cardo-PIM-1 blending 25/2 - - - 1692 14644 - - 8.7 [119]
5.0Cardo-PIM-1 25/2 - - - 873 15191 - - 17.4 [119]
10.0Cardo-PIM-1 25/2 - - - 789 10651 - - 13.5 [119]
PIM-Br/PBI-95:5d Physical 35/6.8 1526 360 87 113 1645 4.1 18.9 14.6 [111]
(Aged 400 days)d blending 35/6.8 412 79 18 31 543 4.3 29.6 17.2 [111]
PIM-Br/PBI-95:5- 35/6.8 2897 759 210 260 3317 3.6 15.8 12.7 [111]
300d
(400 d) 35/6.8 2172 569 158 200 2518 3.6 16.0 12.6 [111]
PIM-1/Am-PAFEK Physical 35/2 - 6 1 1 31 - 30.5 28.1 [112]
(20:80) blending
PIM-1/Am-PAFEK 35/2 - 25 5 6 148 - 29.6 26.4 [112]
(40:60)
PIM-1/Am-PAFEK 35/2 - 59 13 14 354 - 27.2 25.9 [112]
(60:40)
PIM-1/Am-PAFEK 35/2 - 121 28 31 752 - 26.7 24.3 [112]
(80:20)
PIM-co-EA1 Poly- 30/1 - - 238 289 5483 - 23.0 19.0 [125]
condensation
PIM-1-PEG-2000- Physical 30/1 - 314 100 177 2513 3.2 25.5 14.2 [127]
5wt% blending
PIM-1-PEG-2000- 30/1 - 74 20 37 660 3.5 30.9 18.0 [127]
11wt%
PIM-1-(THBA-PEG- Poly- 30/1 - 1265 4450 780 8700 2.8 19.2 11.2 [127]
OMe2)-0.9wt% condensation
a Am-PAFEK: poly(arylene fluorine ether ketone); EA: Ellagic acid; EA(Me ): ethanoanthracene; MEEP: ether-functionalized
2
poly-phosphazene; PEG: Polyethylene glycol; (THBA-PEG-OMe2): PEG-substituted trihydroxy-benzoic compounds; PIM-BM:
bromomethyl PIM; PIM-Br: bromomethylated PIM; PBI: polybenzimidazole; TB: Tröger’s base. b 1 Barrer = 1 x 10-10
cm3(STP)cm/cm2scmHg. c Mixed gas composition CO2:N2:Ar 20:20:60. d Methanol treated membranes

3.4. Incorporation of nanoparticles to form PIMs-based mixed matrix membranes

(MMMs)

Mixed matrix membranes (MMMs) comprising rigid nanofiller phases dispersed in a continuous

polymer phase, bring about the synergistic advantageous properties of both phases (e.g., high gas

permeability of polymers and high selectivity of nanofillers) in ameliorating the separation

performance of the membranes. The embedment of nanofillers in the polymer matrix precisely

adjusts the dynamic properties of the polymer such as contorted molecular configuration,

37
segmental chain packing and FFV, restricting the polymer chain rearrangement and loss of FFV

as a result of insufficient single bonds in the polymer backbones during construction of MMMs

[25], offering enhanced physical, mechanical and thermal properties of the MMMs for feasible gas

separation application.

A vast number of filler materials have been reported to facilitate gas transport in PIMs-based

membranes. Fillers are classified according to their porosity: porous or non-porous. The role of

porous filler materials in MMMs is primarily related to their physical structure and pore size,

allowing precise molecular sieves and faster gas diffusion according to their kinetic sizes and

molecular shapes. With precise pore apertures, the MMMs with porous fillers at optimum loading

would exhibit promising permeability and selectivity. On one other hand, the incorporation of non-

porous fillers in MMMs creates a longer transportation pathway for the gas molecules to travel

through the MMMs by generating a tortuous path, thus increasing gas-pair selectivity. Non-porous

fillers may also enhance the permeability of MMMs by creating additional FFV through disruption

of polymer chain packing. However, it may create non-selective voids that will compromise the

selectivity [24].

In general, fillers are incorporated in membranes by either solution mixing or in-situ method,

constructing the MMMs. Solution mixing method can be categorized into three types: (i) polymer

first dissolved in a specific solvent followed by the addition of filler, (ii) fillers dispersed in specific

solvents prior to the addition of polymer (required additional priming steps), and (iii) polymer and

filler each dispersed in separated solvents before mixing [129-131]. This method is widely applied

to prepare MMMs due to its simplicity. Meanwhile, in-situ growth of filler and one-pot

polymerization are two types of in-situ preparation methods of MMMs. The former method adds

filler precursor to the polymer solution to promote the simultaneous growth of fillers in the dope

38
solution [130], while the latter method mixes fillers to the monomer of polymer for one-step

condensation polymerization [132]. In comparison, the in-situ method promotes better interfacial

compatibility between polymer and filler, attributed to the formation of chemical bonding during

polymerization between the two phases which fosters the development of defect-free MMMs.

Likewise, gel membranes (or soft membranes) are prepared by the addition of liquid-like small

molecules, combining the beneficial characteristics of enhanced operational stability akin to solid

membranes and enhanced permeability akin to liquid membranes. This gel type MMMs possess

exceptional high free volume due to the increased chain mobility, promoting the diffusion of gas

molecules across the membranes [133, 134]. A recent reported PIMs-based gel membrane has

demonstrated excellent H2 permeability enhancement of 78% [135], yet their performance on CO2

separation is underexplored.

Among the PIMs materials, ladder PIMs have been widely studied in tailoring MMMs for gas

separation. To date, the fillers that have been reported for PIM-1 based MMMs include silica- and

carbon-based fillers, MOFs, microporous organic materials such as porous organic frameworks

(POFs), COFs, and HCPs. Table 4 depicts the state-of-the-art PIMs-based MMMs for CO2 capture.

3.4.1. Addressing the challenges on the fabrication of PIMs-based MMMs

In practice, the fabrication of high-performance PIMs-based MMMs often encountered critical

challenges which include (1) dispersibility of fillers in a PIMs polymer matrix, (2) interfacial

compatibility between two phases, (3) optimum size and loading of fillers, and (4) permeability-

selectivity trade-off observed in different polymer-fillers combination in MMMs. All these

challenges are reciprocal, whereby the careful selection and optimization of fillers will reflect the

resultant separation performance of the membranes.

39
The poor dispersibility of fillers in PIMs matrix can lead to agglomeration and sedimentation.

Agglomeration occurs especially at high filler loadings, when the fillers adhere to each other due

to van der Waals forces, electrostatic interactions, and Brownian motion, forming sub-microsized

entities. Sedimentation happens when the dope solution has low polymer concentration and

viscosity. The dispersibility of fillers in MMMs is also influenced by the high aspect ratio of fillers,

processing conditions and solvent selection. In systems with poor filler dispersibility, a “sieve-in-

a-cage” phenomenon may occur, creating non-selective voids within the particles which leading

to the formation of undesirable channels between phases. Gas molecules can then transport through

these extra channels, reducing the selectivity of the membranes.

Poor filler dispersibility and poor polymer-filler compatibility in PIMs-based MMMs are often co-

existence which adversely affect the membrane selectivity. Likewise, insufficient interaction

between polymer and filler results in unfavorable morphology between phases, leading to the

formation of interfacial voids that deteriorate the performance of the membrane. These voids arise

from inadequate interaction between polymer and fillers which result in non-selective bypass for

gas permeation. Furthermore, partial blockage of filler pores and polymer rigidification around the

fillers also exacerbates poor interfacial morphology, hampering the interaction between phases and

compromising the gas permeability. Consequently, the resultant membranes would demonstrate

poor mechanical strength and stability.

Ensuring excellent polymer-filler dispersibility and good compatibility are crucial in fostering the

performance and selectivity of PIMs-based membranes in gas separation applications. To address

these issues, efforts have been made by designing the polymer-filler system via several approaches,

such as decorating the fillers with polymer [136-138], cross-linking the polymer and fillers [132],

tailoring the fillers [139-144] and polymers [101, 144, 145] with functional groups, incorporating

40
molecular solders between polymer-filler interface [146-148], and regulating the structure and

morphology of fillers [149]. Each of these designing approaches and their corresponding

influences on the morphology and performances of MMMs are outlined in the following

subsections.

The performance of PIMs-based MMMs is also affected by the size and loading of fillers. Smaller

particles at the nanoscale are prone to provide a higher surface-area-to-volume ratio compared to

larger fillers which allow better polymer-filler compatibility while at the same time providing more

efficient gas barrier properties. There is also a high tendency of small size fillers to agglomerate

or aggregate at high concentrations [150]. A number of studies have been conducted to evaluate

the effect of particle size on the gas separation performance of resultant PIMs-based MMMs [151-

154]. Meanwhile, an increase in filler loading established a percolative network to enhance gas

separation performance, eliminating loss of selectivity due to non-selective voids. However, high

filler loadings will lead to brittle MMMs with poor mechanical properties. The optimum fillers

loading in PIMs-based MMMs reported from the literature are in the range of 0.05 wt% to 30 wt%

depending on the type of fillers (Table 4). Interestingly, recent work on PIMs-based MMMs has

reached an exceptionally high loading of up to 82.5 wt% based on a symbiosis-inspired synthesis

method, exceeding the threshold limit of filler loading in glassy polymers [154].

3.4.2. Silica-based fillers

Silica particles were the first non-porous fillers introduced into the PIM-1 matrix [156] and

subsequently cross-linked with PIMs [157]. Despite this, the absence of low or no exterior

functional groups on fumed silica nanoparticles created a huge challenge to polymer-filler

compatibility, in which the non-optimum interfaces often lead to the generation of non-selective

41
bypass across the membranes [158, 159]. Sakaguchi et al. performed surface modification on silica

nanoparticles with methyl, amine and carboxylic acid groups to improve the polymer-filler

compatibility as well as the CO2 transport properties of MMMs [139]. Their results revealed that

the silica nanoparticles functionalized with 3,5-dimethylbenzoic acid provided an efficient gas

permeable pathway in the MMMs, resulting in a five-fold enhancement of CO2 permeability

compared to the pristine PIM-1 membranes.

Polyhedral oligomeric silsesquioxanes (POSS) is a cage-like hybrid molecule with the basic

formula RnSinO1.5n, n = 6, 8, 12. It is one of the smallest porous particles with a dimension of 1 –

3 nm [160]. Being a zero-dimensional nanoparticle with precise structure, POSS have diversified

functional groups, excellent organic/inorganic compatibility, and as well as high degree of

molecular-level mixing in polymer matrix. Yong et al. first incorporated DiSilanolIsobutyl POSS

(SO1440) into the PIM-1 matrix to suppress the aging and plasticization behavior of the

membranes [161]. In addition to the separation performance enhancement with only 2 wt% loading

of POSS, the PIM-1/POSS MMMs showed an excellence CO2 plasticization resistance up to 30

atm, while demonstrating 12% less aging effect compared to pristine PIM-1 membranes over 120

days. An investigation on molecular mobility was conducted to gain a deeper understanding of the

effects of PhEPOSS with respect to miscibility, phase behavior and free volume in the PIM-

1/PhEPOSS MMMs [162]. The developed MMMs indicated substantial improvement in gas

separation performance.

Tuning POSS functionality could improve the versatility of these fillers in MMMs. Kinoshita

functionalized the octaphenyl substituted POSS (OPS) with nitro and amino groups via a two steps

modification and embedded them in the PIM-1 matrix [140]. Octa nitrophenyl POSS (ONPS) was

obtained by subjecting OPS to nitration reaction, while further hydrogenate ONPS produced octa

42
aminophenyl POSS (OAPS). The introduction of OPS improved the CO2 permeability by 80%

with a comparable selectivity to pristine PIM-1, while OAPS (up to 10 wt% loading) contributed

a significant increase in the O2/N2, CO2/N2 and CO2/CH4 selectivities of MMMs, that is 85-120%.

Unfortunately, both the ONPS and OAPS filler failed to enhance the permeability. For example,

there was a 98% loss of CO2 permeability at a 20 wt% loading of OAPS. Likewise, the

incorporation of CO2-philic PEG-POSS in the PIM-1 matrix remarkably improved the gas-pair

selectivity [136]. At a lower loading of PEG-POSS (e.g., 1-2 wt%), CO2/CH4 and CO2/N2

selectivities were enhanced by 8-33% and 5-16%, respectively, whereas at higher loading of 5-10

wt %, their selectivities improved 75-150% and 37-63%, respectively. The loss of CO2

permeability in MMMs was up to 65% at higher loading, which can be attributed to the

rigidification of the polymer chain that retarded the gas transportation.

3.4.3. Carbon-based fillers

The embedment of carbon-based 2D multilayer nanosheets in the PIM-1 matrix started with the

introduction of graphene oxide (GO) in PIM-1 membranes for adsorption by Gonciaruk et al. [163].

Later, one manometer thick GO was used in high-performance gas separation [164]. A superior

CO2/N2 selectivity in PIM-1/GO MMMs was reported, attributed to two reasons: (1) polar groups

in GO (e.g., –COOH and –OH) that have specific adsorption affinity toward CO2 which facilitate

the CO2 sorption across membranes, and (2) the occurrence of the microphase segregation in GO

filled membranes which created additional gas transport channels. Functionalized GO could be an

efficient strategy to retard aging behaviour in GO-filled PIM membranes. For instance, GO

functionalized with octylamine (e.g., GO-OA, rGO-OA) [165], octadecylamine (e.g., GO-ODA,

rGO-ODA) [165], PIM-1 (e.g., PGO) [137], (3-aminopropyl) triethoxysilane (e.g., GO-APTS)

[137], POSS (e.g., GO-POSS) [16], as well as holey GO (HGO) functionalized with ODA (e.g.,

43
HGO-ODA) [15], tris(4-aminophenyl)amine (e.g., rHGO-TAPA) [141] and N,N,N’,N’-tetrakis(4-

aminophenyl)-1,4-phenylenediamine (tetrakis) (e.g., rHGO- tetrakis) [141] are some examples of

functionalized-GO that demonstrated promising capability in suppressing the aging phenomenon

in PIM/GO membranes. The remarkable performance of GO materials in MMMs has inspired the

continuous investigation of other versatile carbon-based fillers in fabricating high-performance

membranes. These include graphitic carbon nitride (g-C3N4) [143, 166], boron nitride nanosheets

(BNNS) [167], single-wall carbon nanotubes (SWCNTs) [168, 169], and multi-walled carbon

nanotubes (MWCNTs) [75, 170, 171]. For the first time, Tian et al. fabricated a C2N-type gas

separation membrane for CO2 capture by combining N/O para-doped noble carbons, C2NxO1-x with

PIM-1 [172]. The MMMs achieved superior CO2 permeability of 22110 Barrer with a comparable

CO2/N2 selectivity of 15, while the PEG activation further boosted the permeability to 37272

Barrer. The attractive performance in this ultrapermeable MMMs was attributed to the numerous

polar channels in C2NxO1-x that simultaneously enhanced the solubility and diffusivity of CO2, as

well as the introduction of PEG as plasticizer and liquid porogen which further expanded the

porosity of the membranes. In a recent study, MXene laden with rich polar functionalities (e.g.,

‒OH, ‒O, ‒F) and interlayer spacing of ~3.5 Å was integrated into the PIM-1 matrix to prepare

MMMs with both synergistic effects of solution-diffusion and molecular sieving mechanism [173].

The resultant MMMs presented CO2/N2 selectivity of 32.7 along with excellent CO2 permeability

of 12475 Barrer. The presence of polar groups in MXene augmented CO2 uptake through

preferential adsorption of CO2 over N2, while the presence of precise molecular diffusion channels

provided by the MXene for the transportation of CO2 and N2.

44
3.4.4. Metal-organic frameworks (MOFs)

MOFs are a type of porous organic-inorganic hybrid material that displayed great potential in gas

separation. They are made up of a cluster of metal ions linked through organic linker units to yield

crystalline 1-, 2-, or 3D structures with organic surface and porosity. Through different

combinations of metal precursor and organic ligands, MOFs can be specially designed with tunable

pore structures for efficient separation and good selectivity. MOFs from the subclasses, namely

zeolite imidazolate frameworks (ZIFs) and Universitetet i Oslo (UiO) have been extensively

reported for PIM-1 based MMMs with MOFs nanoparticles. ZIFs are constructed of imidazolate

(im) anionic organic ligands tetrahedrally coordinated by transition metals (M), usually zinc (Zn)

or cobalt (Co). Owing to their ease of preparation, high stability and small pore aperture, ZIFs have

been an attractive candidate for MMMs for CO2 separation from larger gas through the molecular

sieving mechanism. The idea of incorporating ZIF-8 in the PIM-1 membrane was first proposed

by Bushell et al. [151]. They prepared nanosized ZIF-8 at 20 – 60 nm and studied the separation

performance of PIM-1/ZIF-8 MMMs. Inspired by this study [151], Yahia et al. prepared four

different sizes of ZIF-8 nanoparticles with particle sizes of 45 nm, 120 nm, 250 nm, and 450 nm

to study the optimum filler size in fabricating a high-performance PIM-1 MMMs [174]. The

crystallinity and particle size were controlled by the drying conditions of crystals and the

concentration of sodium formate. Higher formate concentration retarded the growth of ZIF-8

crystals as formate acted as a competing ligand for Zn2+ ions. The PIM-1/ZIF-8 MMMs with ZIF-

8 size of 120 nm exhibited the highest CO2 permeability of 9667 Barrer with CO2/CH4 selectivity

of 11.4, lying closely to the Robeson 2008 upper bound. Recently, He et al. introduced a symbiosis-

inspired de novo synthesis method to prepare in-situ nanosized ZIF-8 assembled within the PIM-

1 matrix (Figure 6a) [155]. The interactions between -CN groups of PIM-1 and uncoordinated

45
imidazole groups of ZIF-8 brought excellent MOF dispersion and good PIM-1/ZIF-8 compatibility,

realizing a defect-free interface MMMs with unprecedented ZIF-8 loading of 67.2 wt%. These

symbiosis-inspired solubility-selective MMMs exhibited promising separation performance while

maintaining high long-term stability and great resistance to physical aging and plasticization. ZIF-

7/PIM-1 MMMs and ZIF-67/PIM-1 MMMs were also prepared using a similar synthesis approach

in this work. Notably, the MOFs content in these two MMMs reached a next level of 71.9 wt%

and 82.5 wt%, respectively, demonstrated the universality of this MMMs synthesis method. Zhu

et al. introduced a molecular soldering strategy to engineer the polymer-filler adhesion as well as

gas transport properties of PIM-1/ZIF-8 MMMs, featuring both highly permeable PIM-1 and

highly porous ZIF-8 mediated via multifaceted polyphenol [146]. Polyphenol tannic acid was

employed as a surface modifier to increase the functional groups on the surface of ZIF-8, and acted

as an etching agent to create transfer channels within the PIM-1 matrix (Figure 6b). The

exceptional adhesion nature of polyphenol led to densely packed and stiffed polymer chains,

strengthening the selectivity, while the hollow architecture on ZIF-8 and the protecting-etching

strategy induced by tannic acid reduced the mass transfer resistance and improved the gas

permeability. Synchronous improvement in both permeability (e.g., up to 36%) and selectivity

(e.g., up to 28%) was observed, attributed to the structural advantages that surpassed the trade-off

line.

ZIF-67 is a Co-substituted ZIF-8 that possesses a smaller effective pore aperture as a result of a

stiffer Co-N bond than the Zn-N bond in ZIF-8. This particular characteristic would further

improve the separation performance of MMMs compared to ZIF-8 fillers. This hypothesis was

confirmed by Wu et al. via their investigation of PIM-1/ZIF-67 MMMs and their corresponded

gas separation performance [175]. With the embedment of ZIF-67, the CO2/CH4 selectivity of

46
MMMs showed an apparent improvement over PIM-1/ZIF-8 MMMs. In the proceeding study, the

embedment of hollow ZIF-67 further enhanced the CO2 permeability and CO2/CH4 selectivity of

the membrane by 58% and 31%, respectively [149]. The enhanced permeability was attributed to

the hollow cavity that created an extra volume for gas diffusion, while the improved selectivity

was provided by its molecular-sieving shell. Knowing that particle size played a crucial role in the

polymer-filler interfacial compatibility and gas pair selectivity, Ye et al. introduced nanosized ZIF-

67 (e.g., 25-35 nm) into PIM-1/ZIF-67 MMMs to fortify the PIM-1-ZIF-8 interaction thereby

alleviating the formation of non-selective void [150]. The resultant MMMs exhibited a 69.4%

enhancement in CO2/CH4 selectivity (e.g., from 12.5 to 21.1) compared to the pristine PIM-1

membrane. A modified Maxwell model was also proposed to describe the gas transport behaviors,

justifying that the enhancement of selectivity of PIM-1/ZIF-67 MMMs was attributed to the

rigidification of polymer chains.

Designing fillers with functional groups makes it possible to improve the interfacial interactions

of ZIFs with the PIM-1 matrix through the formation of hydrogen bonding with the polymer matrix,

thereby alleviating the formation of non-selective voids in the MMMs and thus enhancing

CO2/CH4 diffusion across the membrane. Wang et al. incorporated amino-functionalized ZIF-7

(e.g., NH2-ZIF-7) into PIM-1 membranes to prepare MMMs for biogas upgrading [142].

Compared to the narrower pore aperture of ZIF-7 (e.g., 3.0 Å), NH2-ZIF-7 featured a slightly larger

aperture that favored CO2. The PIM-1/ZIF-7 MMMs achieved a CO2/CH4 separation enhancement

of 36% compared to the pristine PIM-1 membranes. Interestingly, the incorporation of amino-

functionalized filler further elevated the enhancement to 65%, attributed to the intrinsic properties

of fillers and synchronous rigidification of polymer chains at the interface. A modified Maxwell

model was also established as illustrated in Figure 6c.

47
The isoreticulars of UiO MOFs are another particular interest of PIM-1 based MMMs. It is made

up of terephthalic acid-based ligands coordinated to [Zr6O4(OH)4] metal clusters. The studies on

UiO MOFs, specifically UiO-66 were first started since its introduction by Lillerud et al. [176].

Khdhayyer et al. investigated the effect of linker functionalization on gas transport by preparing

PIM-1/UiO MMMs loaded with three UiO isoreticular, namely UiO-66, UiO-66-NH2, and UiO-

66-(COOH)2 [177]. The presence of different linkers was proven to tailor the affinity with the

polymer matrix, hence enhancing the gas transport properties of resultant MMMs. All the PIM-

1/UiO MMMs obeyed the Maxwell model, indicating good dispersibility of MOFs in the MMMs.

With increasing filler loading, the MMMs displayed increasing permeability, owing to the

enhanced diffusion pathway offered by the MOFs or polymer-MOFs interface. Likewise, Ghalei

et al. utilized water-modulated synthesis method to prepare UiO isoreticular (e.g., UiO-66, UiO-

66-NH2, and UiO-66-Br) at nanosized and found corollary improvement on the selectivity of

resultant MMMs [152]. Geng et al. designed a defective UiO-66-FA to act as a pillar in PIM-

1/UiO-66-FA MMMs in preventing the collapse of micropores in PIM-1 membranes which further

retarding its physical aging phenomenon [178]. Benefiting from the hydrogen-bond networks

formed between the exposed –OH and –C=O groups in the defective UiO-66 with –C≡N and –

CH3 groups in PIM-1 (Figure 6d), the MMMs demonstrated excellent interfacial compatibility

with fast CO2 diffusion, achieving CO2 permeability of 16591 Barrer. Upon 160 days of prolonged

operation under mixed gas conditions, the membranes experienced only a 25% reduction in

permeability. Zhou et al. introduced a hydrophobic trifluoromethyl functionalized UiO-66 (e.g.,

UiO-66-(CF3)2) into the PIM-1 matrix for effective CO2/N2 separation under humid conditions

[179]. The ‒CF3 groups improved the stability of MOFs in the presence of water, enhanced the

CO2 separation performance through intermolecular force acting on CO2 molecules, as well as

48
reduced the pore size of MOFs for effective gas seiving. An 8% loading of the functionalized

MOFs in the MMMs resulted in 60% and 40% increment in CO2 permeability and CO2/N2

selectivity, respectively. Under a simulated flue gas test, the PIM-1/UiO-66-(CF3)2 composite

membrane exhibited a permeability of 1111 GPU and a competitive CO2/N2 separation factor of

43.8. Long-term permeability test also proved the stability of the membranes in an actual flue gas

environment. Additionally, PIM-MOF chemical cross-linking strategy was also proposed by the

mean of in-situ cross-linked the PIM-1 with UiO-66-NH2 to improve the interfacial compatibility

(Figure 6e) [132]. The covalently linked PIM-1/UiO-66-NH2 MMMs demonstrated a high gas

separation performance and high aging resistance. Apart from this, Su et al. reported on a

polymer/MOF hybridization concept to prepare highly compatible MMMs for CO2 separation

[138]. The membrane system consisted of three components, namely, (i) polymer matrix (e.g.,

PIM-1), (ii) UiO-66 as filler for the membranes and as host for pore confinement in constructing

PIM/MOF nanohybrids (PhUiO-66), and (iii) polymer guests in UiO-66 (e.g., cPIM-1). The

presence of cPIM-1 not only reduced the size and altered the microporosity properties of UiO-66-

NH2, but also promoted the formation of linker-analogous and matrix analogous segments that

coordinated the fillers in the polymer matrix which served as interfacial binders for strengthened

polymer-filler interaction and filler dispersibility. The membranes demonstrated superior CO2

separation performance with CO2 permeability of 9532 Barrer and CO2/N2 selectivity of 63.2 even

after six months, indicating the effectiveness of PhUiO-66 embedment in retarding the physical

aging.

49
a) b)

c) d)

e)

Figure 6. (a) Schematic of PIM-1/ZIF-8 MMMs via symbiosis-inspired de novo synthesis

method [155]. (b) Schematic on the synthetic process and mechanism for tannic acid-tailored

hollow ZIF-8 [146]. (c) Maxwell model predicted for PIM-1/NH2-ZIF-7 MMMs [142]. (d)

Schematic of hydrogen-bond networks form between the exposed –OH and –C=O groups in

defective UiO-66-FA with –C≡N and –CH3 groups in PIM-1 [178]. (e) Schematic of in-situ

polymerized PIM-1 with UiO-66-NH2 [132]. Copyright 2018, 2020, 2022 and 2023. Reproduced

with permission from American Chemical Society, Elsevier, National Academy of Science, and

Nature Portfolio.

Post-synthetic modification (PSM) on MOFs has been proposed to further increase its structural

property and adsorption capacity in MMMs, including metal-based PSM and ligand-based PSM

[180, 181]. For instance, ligand-based PSM was performed to convert UiO-66-NH2 MOFs to a

50
higher CO2-philicity UiO-66-CN. Yu et al. presented PIM-1/UiO-66-CN MMMs with superior

CO2 separation performance with CO2 permeability of 16121 Barrer and CO2/N2 selectivity of 27

[144]. The excellent performance of the MMMs not only attributed to the molecular selectivity

mediated by the UiO-66-CN, but also the reinforced polymer chain which was mutually linked to

UiO-66-CN crystals with rendered freeways for CO2 transport. Gao et al. prepared UiO-66-AO

fillers via a two-step PSM from UiO-66-Br (Figure 7a) and incorporated them in the PIM-1 matrix

for CO2 capture [182]. Benefits from the rich hydrogen bond network formed between the amino

and hydroxyl groups on UiO-66-AO with N and O atoms in the PIM-1 chain, the resultant MMMs

demonstrated excellent polymer-filler compatibility and anti-aging behavior, with CO2

permeability of 7536 Barrer along with CO2/N2 selectivity of 26.9, closing to the 2019 Robeson

upper bound.

To ameliorate the performance of PIMs polymer blend membranes (e.g., PIM-1/MEEP80 blend

membranes), UiO-66-NH2 was incorporated in the MMMs [183]. The developed MMMs led to a

90% enhancement in the CO2 permeability as well as a superior anti-aging performance for up to

300 days. As aforementioned, the AO-PIM-1 membranes despite demonstrating high CO2

selectivity and suppressed aging phenomenon, experienced significant loss in permeability due to

an increase in polymer chain interaction that decreased the d-spacing. A combination of molecular-

sieving materials with functionalized PIMs would be an optimal way to overcome the drawbacks

of these membranes. The incorporation of UiO-66-NH2 into AO-PIM-1 was proven to improve

the CO2 permeability by 190% without compromising the gas-pair selectivity while maintaining

two-thirds of the permeability after 80 days [145]. The promising result was attributed to the strong

hydrogen bonding interactions between –NH2 groups of UiO-66-NH2 and –OH group of AO-PIM-

1 that enhanced the interfacial compatibility by providing a defect-free interface. Carja et al. further

51
evidenced the good MOF/polymer interfacial interactions between AO-PIM-1 and UiO-66-NH2

through a computational and experiment study [184].

Materials of Institute Lavoisier (MILs) MOFs, specifically MIL-101, is a chromium-based MOFs

framework [Cr3O(F/OH)(H2O)2(O2CC6H4CO2)3]) consist of two permanent cages in diameter of

2.9 nm and 3.4 nm. These cages have pentagonal- and hexagonal-ring openings at 1.2 nm and 1.5

nm x 1.6 nm diameters which favor gas transportation [185]. Moreover, the Cr metal clusters

possess higher CO2 and O2 adsorption affinity compared to other metal sites such as Fe, Co, Ni,

and Cu [186]. A series of MIL-101 MOFs was reported to be embedded into PIM-1 to study the

effect of particle size and the effect of the chemical nature of amine- or ethylenediamine-

functionalized MIL-101 on gas separation performance of resultant MMMs [153]. Herein, MIL-

101 (ca. 0.2 µm), NanoMIL-101 (ca. 50 nm), ethylenediamine functionalized MIL-101 (EDMIL-

101), and MIL-101 synthesized by the ligand 2-aminoterephthalic acid (MIL-101-NH2) were

chosen. It was found that both the MIL-101 with amine groups barely improved the separation

performance of the MMMs owing to the initial decrease in selectivity at low MOF loading. On the

other hand, MIL-101 and NanoMIL-101 created a dramatic improvement in the CO2 permeability

of PIM-1/MIL-101 membranes, achieving superior CO2 permeability of 35600 Barrer after an

alcohol treatment. Notably, the MMMs demonstrated excellent CO2 permeability of 3500–3800

Barrer and CO2/N2 selectivity of 25–27; CO2/CH4 selectivity of 21–24 even after 7 years of aging.

3.4.5. Advanced MOFs

A series of advanced MOFs have also been reported for PIM-1-based MMMs in CO2 capture. Mg-

MOF-74, having 1-D hexagonal channels of 11 Å built from the ligand, 2,5-dihydroxyterephthalic

acid and Mg2+ metal ion, is a potential candidate to chemically cross-linked with PIM-1 via the

hydroxyl groups exposed on the outer surface and fluoride chain-ends in the fillers and polymers,

52
respectively, which created a defect-free MMM with superior CO2 permeability of >20000 Barrer

[187]. Meanwhile, a Zr-based CO2-philic framework that possesses a similar topology as UiO-66,

namely MOF-801, demonstrated attractive performance in both selective transport of CO2 and

anti-aging behavior in PIM-1-based MMMs [18]. The resultant membranes showed CO2

permeability and CO2/N2 selectivity improvement of 130% and 35%, respectively, with only a 30%

reduction in permeability after 90 days. Shen et al. incorporated a pyrazole-based MOF-303

(Al(OH)HPDC), containing pyrazoledicarboxylic acid (H3PDC), as a nanofiller blended into PIM-

1 [188]. MOF-303, with imino group from H3PDC ligand, has higher reactivity (i.e. polarizability,

deprotonation and alkaline) as compared to other alkyl-based imino group as a result of irregular

distribution of electron cloud density by N=N double bonds in pyrazole rings. The polar

characteristic of H3PDC linkers was also found to enhance the affinity towards CO2 via dipole-

quadrupolar interaction. The best performing 10% MOF-303/PIM-1 reached a CO2 permeability

of 6602 Barrer and 6199 Barrer for both pure- and mixed gas test respectively, surpassing the 2008

Robeson line with CO2/N2 selectivity up to 25.6.

Numerous advanced MOFs named after the founded university have also shown promising

performance in membrane-based CO2 separation, for instance, King Abdullah University of

Science and Technology-7 (KAUST-7) [17], Massey University Framework-15 (MUF-15) [189],

Hong Kong University of Science and Technology-1 (HKUST-1) [190] and National University

of Singapore-8 (NUS-8) [191-194]. KAUST-7 is a type of fluorinated MOF synthesized from Ni-

pyrazine (4,4’) square grid and (NbOF5)2- inorganic pillars (Figure 7b) with superior sorption

selectivity of CO2 over CH4. The promising characteristics of KAUST-7 enhanced the CO2

permeability of PIM-1/KAUST-7 MMMs while maintaining 91% of the permeability over 150

days owing to the good interfacial compatibility that restricted the movement of polymer chains

53
[17]. Back in 2019, Shane Tefler’s group from Massey University introduced a new type of

framework, namely MUF-15, constructed from a hexanuclear cobalt cluster connected by 10

isophthalate (ipa) struts (Figure 7c) [195]. Benefiting from the exceptional CO2 adsorption

capacity, that is 72 mL/g STP at 293 K and 1 bar, the same group of researchers adopted MUF-15

to fabricate MMMs for membrane-based CO2 separation [189]. Featuring the layered crystalline

structure in MUF-15 (Figure 7d-f), the exfoliated MUF-15 nanosheets/PIM-1 MMMs brought

about significant enhancement in both CO2 permeability and CO2/CH4 selectivity that surpassed

the 2008 Robeson upper bound, even at a low filler loading of 5 wt%. HKUST-1, also known as

Cu-BTC or MOF-199, is a paddlewheel unit copper-based MOFs connected by benzene-1,3,5-

tricarboxylate linker molecules. Fuoco et al. prepared a series of MOF/sandwich membranes by

growing the HKUST-1 or ZIF-8 onto the neat PIM-1, amide(AMD)-modified PIM-1 and

hexamethylenediamine (HMDA)-modified PIM-1 membranes [190]. The present of terminal

amino group of HMDA was found to facilitate efficient nucleation of ZIF-8 or HKUST-1, forming

a uniform and continuous film of MOFs on the polymer surface. Despite, additional growth cycle

compromised the permeability of membranes, mainly due to the decrease in diffusion coefficient

of gases through the membranes.

54
a)

b)

c) d) e) f)

Figure 7. (a) UiO-66-AO fillers synthesized via a two-step PSM from UiO-66-Br [182]. (b)

KAUST-7 synthesized from Ni-pyrazine (4,4’) square grid and (NbOF5)2- inorganic pillars [17].

(c) MUF-15 MOFs [189]. (d-e) SEM images of MUF-15-Bulk particles (10 x 25 μm) and (f)

TEM images of MUF-15 nanosheet (thickness 20 nm) [189]. Copyright 2020, 2022 and 2023.

Reproduced with permission from Elsevier and MDPI.

3.4.6. Photo-responsive MOFs

Stimuli-responsive MOFs have special flexible structures, open active sites, and reversible

physiochemical characteristics which makes them exhibit a larger magnitude of response when

triggered by external stimuli. These stimulants can be pressure, pH, magnetic field, temperature,

or light. Specifically, light-responsive MOFs are categorized based on three generations whereby

in Generation-1, the light-responsive guest molecules were introduced inside the pores. Both

55
Generation-2 and Generation-3 used light-responsive moiety to synthesize MOFs, nonetheless, the

light-responsive ligands do not protrude into the MOFs pores in Generation-3 as of Generation-2

[196]. Light-responsive MOFs have been recently applied in smart membranes [196-198] and low-

energy CO2 capture [199, 200]. They were also found to experience high CO2/N2 selectivity and

efficient CO2 dynamic photoswitching that are applicable for post-combustion CO2 capture [196,

201]. Prasetya & Ladewig synthesized and incorporated a light-responsive MOF, Azo-DMOF-1

(Figure 8a,b), inspired by Dabco MOF or DMOF-1 (Zn(bdc)(dabco)0.5 as filler in PIM-1 MMMs

for effective low-energy CO2 capture [202]. Herein, Zn was used as the metal source, while a

mixed-ligand strategy with 2-phenyldiazenyl terephthalic acid and 1,4-diazabicyclo(2.2.2)octane

was employed to control the light-responsive ligand in the MOFs framework. The azobenzene

functional group from the ligand was found to exhibit a photo-responsive CO2 adsorption in both

static and dynamic states, thus improving the CO2 affinity. The MMMs of Azo-DMOF-1-PIM-1

with 5 wt% of filler loading enhanced the permeability of PIM-1 from 4000 to 6700 Barrer with

CO2/N2 selectivity of around 20. Upon higher filler loading, the reduction of CO2 selectivity was

observed, which possibly attributed to the particle agglomeration and microcracks formation.

Likewise, the same group of researchers adopted a similar methodology to synthesize a light-

responsive UiO-66, namely Azo-UiO-66 for similar applications [203]. The composition of

ligands, that is 2-phenyldiazenyl terephthalic acid and terephthalic acid, was carefully studied to

examine the effect of CO2 light-responsive properties and CO2/N2 selectivity in the resultant

MMMs. It was revealed that Azo-UiO-66 MOFs with a moderate azobenzene linker (e.g.,

Azo(16.7)-UiO-66 and Azo(33.3)-UiO-66) demonstrated a better performance as low-energy CO2

adsorbents, due to the larger pore volume and higher CO2 adsorption capacity (Figure 8c,d).

56
Meanwhile, those of higher azobenzene content (e.g., Azo(66.7)-UiO-66 and Azo(100)-UiO-66)

were suitable for MMMs to improve the separation performance in term of CO2/N2 selectivity.

a) b)

c) d)
UiO-66 0.51

Azo(16.7)-UiO-66 0.41

Azo(33.3)-UiO-66 0.35
MOFs

Azo(66.7)-UiO-66 0.29

Azo(100)-UiO-66 0.18

0 0.2 0.4 0.6


Pore volume (cm3/g)

Figure 8. (a) Hypothetical building unit of Azo-DMOF-1 [202]. (b) Azo-DMOF-1 appeared as

rod-shape crystals with a size extending up to 5 μm [202]. (c) Pore volume and (d) CO2

adsorption of UiO-66 and Azo(X)-UiO-66 at 273 K [203]. Copyright 2018 and 2019.

Reproduced with permission from American Chemical Society and Royal Society of Chemistry.

3.4.7. Task-specific ionic liquid (TSILs) modified MOFs

ILs compatibilization is a strategy proposed to enhance polymer-filler compatibility in MMMs.

The IL-modified MOF (IL/MOF) composites are prepared by impregnating ILs into the pores of

57
MOF, acting as the cavity occupants which alter the physiochemical properties and gas affinities

of MOFs [204]. In addition, ILs can also act as a wetting agent in MMMs to promote the

compatibility between polymer and MOFs. Figure 9a illustrates a few examples of IL. Han et al.

successfully tuned ZIF-67 using two different TSILs and implemented them into PIM-1 MMM to

enhance CO2 permeability and membrane stability [147]. Two different TSILs,

tetramethylgunidinium imidazole (TMGHIM) and tetramethylgunidinium phenol (TMGHPhO)

were selected based on the high solubility and reversible interaction with CO2. The good

compatibility of ZIF-67 and PIM-1 with the TSILs was proven by the SEM images, indicating no

obvious interfacial voids or defects in the membranes (Figure 9b-d). The highest gas separation

performances were obtained by both 10 wt% TMGHPhO@ZIF-67/PIM-1 and TMGHIM@ZIF-

67/PIM-1 with an excellent CO2 permeability up to 12850 Barrer, exceeding the 2008 Robeson

upper bound. In the next study, they adopted a superbase IL, namely (1,8-

diazabicyclo[5,4,0]undec-7-ene imidazole ([HDBU][Im])) to modify the ZIF-67 [148].

[HDBU][Im] displayed higher affinity towards ZIF-67 and PIM-1 as compared to the two TSILs

aforementioned.

IL [Bmim][NTf2] was found to be a promising regulator in MMMs for the construction of defect-

free membranes [205, 206]. Similar to other ILs, the presence of a hydrophobic anion, herein, the

Tf2N-, could promote CO2 affinity and act as a CO2 transport carrier which enhanced the CO2

permeability and selectivity in MMMs [205]. For instance, [Bmim][Tf2N-]@UiO-66-NH2 [207]

and [Bmim][Tf2N-]@MOF-801 [208] both contributed significantly to the construction of MMMs

with good interfacial compatibility and stability, exhibiting CO2 permeability of 8400 Barrer and

9420 Barrer, along with CO2/N2 selectivity of 23 and 29, respectively. In a later study, Geng et al.

employed amino-functionalized IL, e.g., 1-aminopropyl-3-methylimidazole

58
bis(trifluoromethanesulfonyl)imide salt ([NH2PMIM][Tf2N]) in encapsulating UiO-66 to improve

the stability and mechanical strength of the UiO-66 [209]. The resulting PIM-

1/[NH2PMIM][Tf2N]@UiO-66 MMMs demonstrated excellent CO2 permeability of 13779 Barrer

and CO2/N2 selectivity of 35.2 under humid conditions, evidenced by the importance of Tf2N-

anion and amino group in ameliorating the separation performance of MMMs.

a) O - O
N
S S +
-
O + CF3 CF3 N N
+ H 2N N
-
- O O
H 2N N N H
N
N O - O
+ N
N N S S
N N N N N
+
CF3 H CF3
O O
H
TMGHPhO TMGHIM [HDBU][Im] [Bmim][Tf2N]

TMGHPhO@ZIF-67/PIM-1 MMM TMGHIM@ZIF-67/PIM-1 MMM [HDBU][Im]@ZIF-67/PIM-1 MMM

b) c) d)

Figure 9. (a) Chemical structure of ILs TMGHPhO, TMGHIM, [HDBU][Im], and

[Bmim][Tf2N]. Good compatibility of ZIF-67 fillers tuned with (b) TMGHPhO, (c) TMGHIM,

and (d) [HDBU][Im] in PIM-1 MMMs [147, 148]. Copyright 2021 and 2022. Reproduced with

permission from American Chemical Society and Elsevier.

3.4.8. Two-dimensional (2D) MOFs

2D fillers with distinctive structures have become a topical subject for MMM fabrication.

Benefiting from the large specific surface area and high aspect ratio, these nanosheets efficiently

alter the polymer chain packing and enhance the polymer crystallinity, thereby distorting the

diffusion path and creating a tortuous pathway on the gas transportation, eventually resulting in

higher gas-pair selectivity. Cheng et al. first proposed the incorporation of bottom-up-strategy-

59
synthesized NUS-8 MOF nanosheet into PIM-1 membranes (Figure 10a,c) [191]. NUS-8 are made

up of zirconium clusters and 1,3,5-benzenetribenzoate (BTB), and possess high chemical

resistance toward boiling water and acidic conditions [210]. Strong hydrogen bonding and π–π

interactions between PIM-1 and NUS-8 (Figure 10f) have contributed to the enhanced interfacial

compatibility of resultant MMMs. Notably, a mere loading of NUS nanosheets (e.g., 2 wt%) in

MMMs has significantly enhanced the CO2/CH4 selectivity to 30. Following that, a series of

functionalized NUS-8 (e.g., NUS-8-NH2 [192] (Figure 10a,d) and NUS-8-COOH [193, 194])

(Figure 10b,e) were proposed for the application of PIM-1 based membranes in CO2 capture. The

computational method was also performed as a preliminary step to assess the membrane system at

a molecular level. These functional groups significantly contributed to the enhancement of

membrane performance, verifying the importance of hydrogen bonding and π–π interactions

between the PIM-1 and the functionalized MOFs. Sun et al. synthesized basic cobalt carbonated

supported ZIF-67 (BCoC-ZIF) nanoplates via facile template conversion method and introduced

the composite into PIM-1 matrix [211]. BCoC as the CO2-philic basic carbonates was proven to

increase the CO2 adsorption capacity, while the BCoC-ZIF with 12 h reaction time in 2-imidazolate

aqueous solution demonstrated great potential in boosting the separation performance of PIM

MMMs to a CO2 permeability of 7326 Barrer, along with CO2/N2 selectivity of 32.5, surpassing

the 2019 Robeson upper bound.

60
a) b)

c) d) e)

f)

Figure 10. Synthesis of (a) NUS-8, NUS-8-NH2 [192], and (b) NUS-8-COOH [194] nanosheets

from the building blocks of Zr cluster and BTB ligand. The morphology of (c) NUS-8 [192], (d)

NUS-8-NH2 [192], and (e) NUS-8-COOH [194] nanosheets using FESEM. (f) Illustration on the

alignment of PIM-1 and NUS-8 nanosheets as a result of strong π–π interactions and hydrogen

bonding [191]. Copyright 2018 and 2022. Reproduced with permission from American Chemical

Society and Elsevier.

61
3.4.9. Microporous organic polymers (MOPs) materials

MOPs are a family of covalently bonded organic components with robust structure, rigid pore

chemistry, and excellent chemical stability. Benefiting from the formation of π–π aromatic group

stacking that contributed great affinity to the polymer matrix, these materials have superior

compatibility with polymer, constructing high-performance MMMs for gas separation [212].

Figure 11 illustrates several examples of MOPs that have been reported in the fabrication of PIM-

1 MMMs.

COFs are the current research hot spot in many applications such as energy storage and conversion,

catalysis, drug delivery, and gas separation. They are synthesized with a versatile combination of

organic building units assembled by covalent bonds, which endow them with orderly arranged

pores, tunable pore size, large surface area, as well as excellent thermal and chemical stability. The

implementation of COFs in MMMs is limited, particularly ascribed to the poor polymer-filler

interfacial compatibility, non-uniform filler dispersion, limited functional groups and relatively

large pore size [213]. Despite this, some advancement has been achieved in PIM/COF MMMs

recently. Schiff base network (SNW-1) synthesized via Schiff base chemistry, held the potential

to enhance the carbon capture performance of PIMs membranes [214]. The formation of hydrogen

bonds between the –C≡N and Ar-O-Ar groups of PIM-1 with –N-H groups in SNW-1 endowed

good polymer-filler compatibility within the matrix. Moreover, the uniform dispersion of SNW-1

in the PIM-1 matrix created more FFV and offered extra 0.5 nm gas transport pathways, resulting

in an enhancement of CO2 permeability of 116%. Similar good polymer-filler interaction was also

found on PIM-1/triphenyltriazine-based β-ketoenamine-linked COF (TpTta-COF) MMMs with

2.4 times improvement in CO2 permeability [215]. Yu et al. presented COFs synthesized from

triangle melamine with linear 1, 4-piperazinedicarboxaldehyde via microwave-assisted Schiff base

62
reaction to derive nanoscale fillers for PIM-1 based MMMs on selective CO2 separation [216].

The nanosized (e.g., ~42 nm), high porosity (e.g., ~1 Å) and molecular affinity of MAPDA have

contributed to the enhancement of the CO2 separation performance of resultant MMMs. Besides,

the incorporation of COFs containing λ5-phosphinine, denoted as CPSF-EtO in PIM-1 matrix also

demonstrated 50% and 27% enhancement on CO2 permeability and CO2/N2 selectivity,

respectively [217].

The covalent triazine framework (CTF-1) is a type of POPs proposed by Thomas et al. in 2008

[218]. CTFs possess several attractive characteristics as fillers in MMMs, such as ordered structure,

hydrophobic nature, amorphous microporous structure, displaying good dispersibility in an

organic solvent, excellent chemical and thermal stability, as well as good sorption with CO2.

Inspired by the ultramicropore structure and strong CO2 affinity (e.g., CO2 uptake 1.76 mmol/g)

of perfluorinated CTF-1 (FCTF-1) [219], Guiver et al. incorporated the FCTF-1 into PIM-1 to

prepare MMM for CO2 capture [220]. The resultant MMMs demonstrated promising separation

performance, ascribed to the microporosity of FCTF-1 and the presence of polar functionalities

such as triazine rings and fluorine atoms that contributed to the diffusivity selectivity and solubility

selectivity, respectively. Wang et al. presented two amine-modified triptyene-based POPs as fillers

to prepare PIM-1 MMMs [221]. The POP fillers were synthesized from triptyene and

formaldehyde dimethyl acetal (FDA), catalyzed by FeCl3, and later subjected to amine

functionalization to yield two amine-modified POPs (e.g., TPFC-CH2NH2 and TPFC-CH2PEI).

Density functional theory (DFT) analysis proved the formation of defect-free PIM-1/TPFC-

CH2NH2 MMMs with parallel planes distance of only 3.8 Å and 3.2 Å. The ideal separation

performance was found on PIM-1/TPFC-CH2NH2-15 MMM, with CO2/N2 selectivity 2.6 times

63
higher than PIM-1 membrane. The membrane also demonstrated fair aging behavior with only

22.4% loss of CO2 permeability over 210 days.

PAFs, first described in 2010 [222], are tetraphenylenemethane building units with tetrahedral

carbon sites connected to biphenyl linkers. PAF-1 possesses a surface area of 5600 m2/g and pore

volume ranging 0.89–1.44 cm3/g with excellent physical and chemical stability [223]. Previous

studies evidenced the presence of PAF-1 in suppressing physical aging that is associated with loss

of permeability in high FFV glassy polymers such as PTMSP and PIM-1 [224, 225]. In addition,

a mere loading of 10 wt% PAF-1 was significantly to improve the H2, N2, CH4 and CO2 gas

permeability of phenazine-containing triptycene ladder polymer membranes (e.g., TPIM-2) by

130–200% [226]. Recent investigation on PIM-1/PAF-1 MMMs revealed superior enhancement

in H2 permeability after incorporation of PAF-1, further validating the H2-selective properties of

PAF-1 in MMMs [227].

HCPs are also prepared via Friedel-Crafts chemistry by self-cross-linking or “FDA-knitting” of

aromatic hydrocarbon monomers. HCPs are distinctive from PAFs from their structural flexibility

and visibly swell in the presence of certain gases and solvents especially under high pressure [228,

229]. HCPs have been a promising porous additive in MMMs to improve permeability and resist

aging [230, 231]. Hou et al. synthesized simple, low-cost and scalable HCPs, namely poly-

dichloroxylene (pDCX) and hydroxylated pDCX for PIM-1 MMMs [231]. On top of the polymer-

additive interaction that influences the membrane performance, the choice of casting solvents (e.g.,

THF, DMF/mesitylene) was also found to improve the gas-pair selectivity and tune the physical

aging rates.

Porous polymer networks (PPNs) are prepared via acid-catalyzed trimerization of three acetyl

groups and possess large surface area (e.g., 400 m2/g) and rigid backbones. Han et al. reported a

64
one-step facile preparation of cPIM-1/PPNs composite membranes for enhanced CO2 capture

[101]. In this study, methane sulfonic acid (MSA) was employed to simultaneously hydrolyze and

cross-link PIM-1 to cPIM-1 as well as in situ synthesis of PPNs in the PIM-1 matrix. The resultant

cPIM-1/PPNs membranes showed a significant enhancement of 200% in CO2 permeability, which

is ascribed to the large pore size of PPNs (e.g., 20.1 Å) and penetration of MSA molecules into the

PIM-1 matrix that protonated the nitrile groups and facilitated the polymer chain movement.

A tightly cross-linked polymer network was introduced in the pursuit of developing practical

organic fillers for MMMs. Tetrafunctional monomer dicyanooctafluoro biphenyl (DCOB) was

used to replace the TFTPN monomer and synthesized with TTSBI to yield network-PIM-1 via a

step-growth polymerization. Results showed that the nanosheet structure of network PIM-1,

aroused from the innate anisotropy in the monomers that extended in the direction perpendicular

to the orientation of nitrile groups during the reaction [232]. A noticeable improvement in

CO2/CH4 separation was found at a low loading (e.g., 0.5 wt%) in the PIM-1 matrix, analogous to

that of PIM-1 MMMs with high aspect ratio fillers [232]. As a continuation of the above research,

the same group of researchers proposed a low cross-linked density (LCD) network PIM-1 using a

combination of TTSBI with tetrafluoro- and octofluoro-monomers [233]. The LCD-network-PIM-

1 was further grafted with PIM-1 chains, yielding a spheroidal polymerized mixed matrix (PMM)

composite that constrained the PIM-1 mobility and enhanced polymer-network compatibility. The

membrane with grafted-LCD-network-PIM-1 after 24 h achieved a 2.8 times increment in CO2

permeability and well-controlled aging behavior over 160 days (i.e., 29% drop of initial CO2

permeability).

POFs such as cyclodextrin (e.g., β-CD [234]) and calixarene (e.g., 4-tert-butylcalix[4]arene [235]),

are attractive fillers in PIM-1 based MMMs. The unique bulky-bowl conformation provides

65
additional transport pathways or size-selective ability to the polymer matrix, thereby enhancing

the separation performance of resultant MMMs. However, the addition of POFs via physical

blends faced the challenge of agglomeration of the particles. The free volume of the membrane

can be also affected due to the grafting of POFs on the side chain of the polymer. Containing

reactive hydroxyl groups at the rim of the bowl structure, POFs readily react with PIMs via

aromatic nucleophilic polycondensation reaction to achieve a restructured copolymer that

mitigates the problem aforementioned. Moreover, this approach regulates the polymer chain and

produces a membrane with larger micropores and higher FFV for improved gas permeability, as

well as suppressed the occurrence of physical aging as a result of chain movement and collapse of

micropores [234, 235].

TpTta-COF SNW-1

MAPDA
PAF-1

pDCX OH-pDCX MOPs


β-CD 4-tert-butylcalix[4]arene

POPs

PPNs (4,4′-diacetylbiphenyl)

TTSBI, TFTPN, DCOB

FCTF-1

Figure 11. Common MOPs used in PIM-1 MMMs for CO2 capture.

66
Table 4. Recent state-of-the-art PIMs-based MMMs for membrane-based gas separation.

Loading T/P Permeability (Barrer)a Selectivity


Membrane ID Filler Ref.
(wt%) (°C/atm) H2 O2 N2 CH4 CO2 O2/N2 CO2/N2 CO2/CH4
Silica-based fillers
PIM-1b Fumed silica 0 23/4.5 3320 1340 405 - 6000 3.3 15.0 - [156]
(Cabosil TS 530)
PIM-1/FS-13vol%b 13 23/4.5 5060 2330 880 - 10100 2.7 12.0 - [156]
PIM-1/DMBA-FS-0wt% DMBA-FS 0 35/1 - - 63 - 1250 - 19.8 - [139]
PIM-1/DMBA-FS-25wt% 25 35/1 - - 472 - 7930 - 16.8 - [139]
PIM-1 SO1440 POSS 0 35/3.5 - 950 300 450 4780 3.2 15.9 10.6 [161]
PIM-1/SO1440-0.5wt% 0.5 35/3.5 - 1550 550 850 7950 2.8 14.5 9.4 [161]
PIM-1/SO1440-2wt% 2 35/3.5 - 1300 400 600 6730 3.3 16.8 11.2 [161]
PIM-1/PhE-POSS-0wt% PhE-POSS 0 35/1 - 275 75 90 1112 3.7 14.8 12.4 [162]
PIM-1/PhE-POSS-1wt% 1 35/1 - 900 280 400 4920 3.2 17.6 12.3 [162]
PIM-1 0 25/4 2504 966 256 335 4087 3.8 16.0 12.2 [140]
PIM-1/OPS-10wt% OPS 10 25/4 3811 1488 487 615 6589 3.1 13.5 10.7 [140]
PIM-1/OPS-20wt% 20 25/4 4103 1707 576 715 7420 3.0 12.9 10.4 [140]
PIM-1/ONPS-10wt% ONPS 10 25/4 1895 677 165 195 3002 4.1 18.2 15.4 [140]
PIM-1/ONPS-20wt% 20 25/4 1582 482 115 147 2416 4.2 21.0 16.4 [140]
PIM-1/OAPS-10wt% OAPS 10 25/4 1025 124 18 20 542 6.8 29.6 26.8 [140]
PIM-1/OAPS-20wt% 20 25/4 330 17 2 2 82 10.6 50.7 41.7 [140]
PIM-1 0 30/1 - - 198 315 3795 - 19.2 12.0 [136]
PIM-1/PEG-POSS-1wt% PEG-POSS 1 30/1 - - 188 255 3360 - 17.9 13.2 [136]
PIM-1/PEG-POSS-2wt% 2 30/1 - - 154 208 3381 - 22.0 16.3 [136]
PIM-1/PEG-POSS-5wt% 5 30/1 - - 74 91 1875 - 25.3 20.6 [136]
PIM-1/PEG-POSS-10wt% 10 30/1 - - 43 44 1309 - 30.4 29.8 [136]
Carbon-based fillers
PIM-1 GO 0 30/4 - - 230 - 3277 - 14.2 - [164]
PIM-1/GO-0.2wt% 0.2 30/4 - - 50 - 6169 - 123.4 - [164]
PIM-1b,c 0 25/2 - - - 330 6400 - - 20.3 [165]
(Aged 155 days)b,c 0 25/2 - - - 70 2000 - - 30.0 [165]
PIM-1/GO-ODA-0.05wt%b,c GO-ODA 0.05 25/2 - - - 320 5500 - - 17.6 [165]
(Aged 155 days)b,c GO-ODA 0.05 25/2 - - - 130 3000 - - 23.7 [165]
PIM-1/rGO-ODA-0.05wt%b,c rGO-ODA 0.05 25/2 - - - 230 4700 - - 21.0 [165]
(Aged 155 days)b,c rGO-ODA 0.05 25/2 - - - 80 2400 - - 29.1 [165]
PIM-1/rGO-OA-0.05wt%b,c rGO-OA 0.05 25/2 - - - 360 5700 - - 17.4 [165]
(Aged 155 days)b,c rGO-OA 0.05 25/2 - - - 140 3500 - - 22.9 [165]
PIM-1b,c 0 25/1.5 - - - 543 6190 - - 11.7 [137]
(Aged 150 days)b,c 0 25/1.5 - - - 219 3283 - - 15.6 [137]
PIM-1/P-GO24-1wt%b,c PIM-1-functionalized GO 1 25/1.5 - - - 437 5421 - - 12.6 [137]
+ unattached PIM-1
(Aged 150 days)b,c 1 25/1.5 - - - 187 3120 - - 16.9 [137]

67
PIM-1/P-A24-10wt%b,c PIM-1-functionalized GO- 10 25/1.5 - - - 434 5027 - - 11.7 [137]
APTS
+ unattached PIM-1
(Aged 150 days)b,c 10 25/1.5 - - - 193 3458 - - 18.4 [137]
PIM-1/f-P-A24-10wt%b,c 100 wt% (PIM-1)- 10 25/1.5 - - - 176 2073 - - 11.8 [137]
functionalized GO-APTS
(Aged 150 days)b,c 10 25/1.5 - - - 110 1763 - - 16.0 [137]
PIM-1b - 0 25/1 - - - 387 7323 - - 18.9 [16]
PIM-1b,c - 0 25/1 - - - 590 7195 - - 12.3 [16]
(Aged 160 days)c - 0 25/1 - - - 186 3114 - - 20.9 [16]
(Aged 160 days)b,c - 0 25/1 - - - 208 3048 - - 14.6 [16]
PIM-1/GO-POSS48-0.05wt%b GO-POSS48 0.05 25/1 - - - 415 8250 - - 19.9 [16]
PIM-1/GO-POSS48-0.05 wt%b,c 0.05 25/1 - - - 613 8026 - - 13.3 [16]
(Aged 160 days)b 0.05 25/1 - - - 189 3940 - - 20.8 [16]
(Aged 160 days)b,c 0.05 25/1 - - - 181 3524 - - 19.4 [16]
PIM-1/GO-POSS72-0.05wt%b GO-POSS72 0.05 25/1 - - - 567 11682 - - 20.6 [16]
PIM-1/GO-POSS72-0.05 wt%b,c 0.05 25/1 - - - 1017 12185 - - 12.0 [16]
(Aged 160 days)b 0.05 25/1 - - - 279 5962 - - 21.3 [16]
(Aged 160 days)b,c 0.05 25/1 - - - 434 5908 - - 13.6 [16]
PIM-1b,c - 0 25/1.5 - - - 543 6190 - - 11.7 [15]
(Aged 150 days)b,c - 0 25/1.5 - - - 219 3283 - - 15.6 [15]
PIM-1/ODA-HGO-4h-0.1wt%b,c HGO-ODA-4h 0.1 25/1.5 - - - 527 6146 - - 11.8 [15]
(Aged 150 days)b,c 0.1 25/1.5 - - - 263 3763 - - 14.4 [15]
PIM-1/P-H24-1wt%b,c PIM-1 functioanalized 1 25/1.5 - - - 480 5675 - - [15]
12.2
HGO-OFA-4h
(Aged 150 days)b,c 1 25/1.5 - - - 240 3707 - - 15.8 [15]
PIM-1/P-H24-10wt%b,c 10 25/1.5 - - - 360 4727 - - 13.6 [15]
(Aged 150 days)b,c 10 25/1.5 - - - 232 3283 - - 16.1 [15]
PIM-1b,c - 0 25/1 - - - 611 6887 - - 11.5 [141]
(Aged 611 days)b,c - 0 25/1 - - - 129 2309 - - 19.4 [141]
PIM-1/rHGO-TAPA-0.1wt%b,c rHGO-TAPA 0.1 25/1 - - - 856 11077 - - 13.0 [141]
(Aged 611 days)b,c 0.1 25/1 - - - 268 3573 - - 13.3 [141]
PIM-1/rHGO-Tetrakis-0.1wt%b,c rHGO-Tetrakis 0.1 25/1 - - - 816 9760 - - 12.2 [141]
(Aged 611 days)b,c 0.1 25/1 - - - 257 3770 - - 15.3 [141]
PIM-1 g-C3N4 0 30/2 1936 - 148 229 3425 - 23.1 15.0 [166]
PIM-1b 0 30/2 3018 - 328 572 6501 - 19.8 11.4 [166]
PIM-1/gCN-1wt% 1 30/2 3830 - 354 503 5785 - 16.3 11.5 [166]
PIM-1/gCN-1wt%b 1 30/2 5720 - 765 1270 10528 - 13.8 8.3 [166]
PIM-1/gCN-SO3-0wt% sulfonated g-C3N4 0 35/3.5 2201 813 236 343 3930 3.5 16.7 11.5 [143]
PIM-1/gCN-SO3-1wt% 1 35/3.5 2018 660 189 303 3740 3.5 19.8 12.4 [143]
PIM-1b,c BNNS 0 25/2 - - - 758 6578 - - 8.7 [167]
(Aged 414 days)b,c 0 25/2 - - - 320 2758 - - 10.8 [167]
PIM-1/BNNS-0.5wt%b,c 0.5 25/2 - - - 644 5940 - - 9.3 [167]
(Aged 414 days)b,c 0.5 25/2 - - - 387 4647 - - 12.8 [167]
ITTB/COOH-CNT-0wt% COOH-CNTs 0 35/2 3426 504 91 134 2052 5.5 22.6 15.3 [169]

68
ITTB/COOH-CNT-1wt% 1 35/2 3886 701 121 165 2620 5.8 21.7 15.9 [169]
PIM-1d EDA-MWCNTs 0 25/1 - - 290 - 4400 - 15.2 - [75]
Cardo-PIM-1d 0 25/1 - - 890 - 13000 - 14.6 - [75]
Cardo-PIM-1/EDA-MWCNTs- 7.5 25/1 - - 1200 - 29000 - 24.2 - [75]
7.5wt%d
PIM-1b,c,e C2NxO1-x 0 27/1.3 - - 627 779 9968 - 15.9 12.8 [172]
PIM-1/C2NxO1-x-10wt%b,c,e 10 27/1.3 - - 1193 1511 20391 - 17.1 13.5 [172]
PIM-1/PEG-C2NxO1-x-10wt%b,c,e PEG-C2NxO1-x 10 27/1.3 - - 2570 3388 37272 - 14.5 11.0 [172]
PIM-1f MXene 0 25/3 - - 344 - 6474 - 18.8 - [173]
PIM-1/MXene-0.5wt%f 0.5 25/3 - - 382 - 12475 - 32.7 - [173]
(Aged 120 days)f 0.5 25/3 - - 173 - 6400 - 37.0 - [173]

Metal-organic Frameworks (MOFs)


PIM-1/ZIF-8-32.4wt% ZIF-8 32.4 RT/1 5745 1640 380 510 6820 4.3 17.9 13.4 [151]
PIM-1/ZIF-8-5wt% ZIF-8 5 RT/4 - - - 852 9667 - - 11.4 [174]
(Aged 40 days) - - - 283 3923 - - 13.9 [174]
PIM-1/ZIF-8-Sym67.2wt% ZIF-8 67.2 35/3.5 2860 - 260 336 6338 - 24.4 18.9 [155]
(Aged 300 days) - - 189 226 4817 - 25.5 21.3 [155]
PIM-1/ZIF-7-Sym71.9wt% ZIF-7 71.9 35/3.5 1800 - 93 120 1900 - 20.4 15.8 [155]
PIM-1/ZIF-67-Sym82.5wt% ZIF-67 82.5 35/3.5 2300 - 95 300 3300 - 34.5 11.0 [155]
PIM-1 0 35/3.5 - - 309 416 6065 - 19.6 14.6 [146]
PIM-1/TA-ZIF-8-5:0.5 TA-ZIF-8 0.5t 35/3.5 - - 382 512 6917 - 18.1 13.5 [146]
PIM-1/TA-ZIF-8-5:1 1t 35/3.5 - - 822 1141 13564 - 16.5 11.9 [146]
PIM-1/TA-ZIF-8-5:3 3t 35/3.5 - - 2012 2687 22046 - 11.0 8.2 [146]
PIM-1/TA-HZIF-8-5:0.5 TA-HZIF-8 0.5t 35/3.5 - - 201 276 5442 - 27.1 19.7 [146]
PIM-1/TA-HZIF-8-5:1 1t 35/3.5 - - 309 441 8268 - 25.1 18.7 [146]
PIM-1/TA-HZIF-8-5:1f 1t 35/2 - - 292 - 6495 - 21.9 - [146]
PIM-1/TA-HZIF-8-5:1c 1t 35/2 - - - 398 6352 - - 16.2 [146]
PIM-1/TA-HZIF-8-5:3 3t 35/3.5 - - 381 635 9024 - 23.7 16.7 [146]
PIM-1/TA-HZIF-8-5:5 5t 35/3.5 - - 568 1059 11057 - 19.5 13.4 [146]
PIM-1/ZIF-67-20wt% ZIF-67 20 30/2 3542 - 215 310 5206 - 24.2 16.8 [175]
PIM-1/ZIF-67-20wt%o ZIF-67 20 30/2 - - - 343 5104 - - 14.9 [175]
PIM-1 HZIF-67 0 30/2 - - - 362 4521 - - 12.5 [149]
PIM-1/HZIF-67-28vol% 28u 30/2 - - - 317 5204 - - 16.4 [149]
PIM-1/SZIF-67-15wt% SZIF-67 15 30/2 - - - 133 2805 - - 21.1 [150]
PIM-1/SZIF-67-15wt%o SZIF-67 15 30/2 - - - 129 2567 - - 19.9 [150]
PIM-1/ZIF-7-20wt% ZIF-7 20 30/2 - - - 157 2663 - - 17.0 [142]
PIM-1/ZIF-7-NH2-20wt% ZIF-7-NH2 20 30/2 - - - 143 2953 - - 20.6 [142]
PIM-1/ZIF-7-NH2-20wt%o - - - 143 2832 - - 19.8 [142]
PIM-1 0 25/1 2710 875 219 286 4770 4.0 21.8 16.7 [177]
PIM-1b 0 25/1 5040 1580 387 522 8210 4.1 21.2 15.7 [177]
PIM-1/UiO-66-9.1wt% UiO-66 9.1 25/1 3080 1010 256 371 5940 3.9 23.2 16.0 [177]
PIM-1/UiO-66-16.6wt% 16.6 25/1 4560 1360 368 527 4560 3.7 12.4 8.7 [177]
PIM-1/UiO-66-28.6wt% 28.6 25/1 7530 1070 364 440 4940 2.9 13.6 11.2 [177]
PIM-1/UiO-66-NH2-9.1wt% UiO-66-NH2 9.1 25/1 2970 802 216 291 4810 3.7 22.3 16.5 [177]

69
PIM-1/UiO-66-NH2-9.1wt%b 9.1 25/1 4760 1660 397 595 8740 4.2 22.0 14.7 [177]
(Aged 3 months)b,g 9.1 25/1 - 754 171 - 4835 4.4 28.3 - [177]
(Aged 3 months)b,h 9.1 25/1 - - - 193 4337 - - 22.5 [177]
PIM-1/UiO-66-NH2-16.6wt% 16.6 25/1 3130 1090 303 425 6340 3.6 20.9 14.9 [177]
PIM-1/UiO-66-NH2-23.1wt% 23.1 25/1 3210 910 252 345 5070 3.6 20.1 14.7 [177]
PIM-1/UiO-66-NH2-28.6wt% 28.6 25/1 3000 1050 293 474 6310 3.6 21.5 13.3 [177]
PIM-1/UiO-66-(COOH)2-9.1wt% UiO-66-(COOH)2 9.1 25/1 2070 748 220 326 4600 3.4 20.9 14.1 [177]
PIM-1/UiO-66-(COOH)2-16.6wt% 16.6 25/1 2370 860 254 392 5190 3.4 20.4 13.2 [177]
PIM-1/UiO-66-(COOH)2-23.1wt% 23.1 25/1 2390 873 266 410 5300 3.3 19.9 12.9 [177]
PIM-1/UiO-66-(COOH)2-28.6wt% 28.6 25/1 3740 1220 296 401 6090 4.1 20.6 15.2 [177]
PIM-1/UiO-66-ref-10wt% UiO-66-ref 10 25/4 2560 1205 315 381 5210 3.8 16.5 13.7 [152]
PIM-1/UiO-66-10wt% UiO-66 10 25/4 2985 626 112 140 2631 5.6 23.5 18.8 [152]
PIM-1/UiO-66-NH2-10wt% UiO-66-NH2 10 25/4 2641 658 104 102 2869 6.3 27.5 28.3 [152]
PIM-1/UiO-66-Br-10wt% UiO-66-Br 10 25/4 2275 595 132 166 2846 4.5 21.6 17.1 [152]
PIM-1/UiO-66-FA-20wt%d UiO-66-FA 20 30/1 - - 715 - 16519 - 23.1 - [178]
PIM-1 UiO-66-(CF3)2 0 60/1 - 135 - 3265 - 24.2 - [179]
PIM-1/UiO-66-(CF3)2 8 60/1 - - 155 - 5242 - 33.8 - [179]
UiO-66-PIM12h UiO-66-NH2 20 25/2 - 1942 555 828 15815 3.5 28.5 19.1 [132]
UiO-66-PIM72h 20 25/2 - 945 230 392 12498 4.1 54.2 31.9 [132]
PIM-1 0 25/1 - - 613 - 7975 - 12.1 - [138]
PIM-1/UiO-66 UiO-66 20 25/1 - - 895 - 10522 - 12.2 - [138]
PIM-1/PhUiO-662-20wt% PhUiO-662 20 25/1 - - 394 - 6040 - 23.2 - [138]
(Aged 6 months)d 20 25/1 - - 151 - 9532 - 63.2 - [138]
PIM-1/ PhUiO-662-30wt% PhUiO-662 30 25/1 - - 329 445 9477 - 29.0 21.5 [138]
PIM-1 UiO-66-CN 0 25/1.4 - - 113.9 - 3027.7 - 26.6 - [144]
PIM-1/UiO-66-CN-20wt% 20 25/1.4 - - 265 - 7070 - 26.7 - [144]
PIM-1/UiO-66-CN-20wt%d 20 25/1.4 - - 145 - 5087 - 35.0 - [144]
sPIM-1/UiO-66-NH2-20wt% UiO-66-NH2 20 25/1.4 - - 479 - 8620 - 18.0 - [144]
sPIM-1/UiO-66-CN-20wt% UiO-66-CN 20 25/1.4 - - 596 - 16121 - 27.0 - [144]
sPIM-1/UiO-66-CN-20wt%d 20 25/1.4 - - 226 - 12063 - 53.5 - [144]
sPIM-1/UiO-66-CN-20wt%e 20 25/1.4 - - 308 - 10389 - 33.7 - [144]
sPIM-1/UiO-66-CN-20wt%i 20 25/1.4 - - 277 - 9095 - 32.8 - [144]
sPIM-1/UiO-66-CN-20wt%j 20 25/1.4 - - 312 - 10354 - 33.2 - [144]
PIM-1 UiO-66-AO 0 25/2 - - 301 496 5573 - 18.5 11.2 [182]
PIM-1/UiO-66-AO-10wt% 10 25/2 - - 279 500 7536 - 27.0 15.1 [182]
PIM-1/25%MEEP80k UiO-66-NH2 0 22/1.3 - - - - 3140 - 25.4 - [183]
PIM-1-25%MEEP80/UiO-66-NH2- 10 22/1.3 - - - - 4968 - 22.5 - [183]
10wt%k
(Aged 306 days)k 10 22/1.3 - - - - 4905 - 26.0 - [183]
PIM-1-25%MEEP80/UiO-66-NH2- 30 22/1.3 - - - - 5970 - 22.5 - [183]
30wt%k
PIM-1-25%MEEP80/UiO-66-NH2- 10 22/1.5 - - - - 4783 - 27.1 - [183]
10wt%l
10 35/1.5 - - - - 4328 - 23.1 - [183]
10 45/1.5 - - - - 4127 - 20.2 - [183]

70
10 22/2.5 - - - - 4431 - 27.00 - [183]
10 22/3.5 - - - - 4380 - 26.7 - [183]
PIM-1 0 35/1 - - 103 121 2902 - 28.0 24.0 [145]
PIM-1/UiO-66-NH2-30wt% UiO-66-NH2 30 35/1 - - 603 768 9420 - 15.6 12.3 [145]
AO-PIM-1/UiO-66-30wt% UiO-66 30 35/1 - - 361 442 8126 - 22.5 18.4 [145]
AO-PIM-1/UiO-66-NH2-30wt% UiO-66-NH2 30 35/1 - - 306 366 8425 - 27.5 23.0 [145]
AO-PIM-1/UiO-66-NH2-30wt%m 30 35/1 - - 265 - 8351 - 31.5 - [145]
AO-PIM-1/UiO-66-NH2-30wt%d 30 35/1 - - 252 - 8541 - 33.9 - [145]
AO-PIM-1/UiO-66-NH2-30wt%n 30 35/1 - - 242 - 8446 - 34.9 - [145]
AO-PIM-1/UiO-66-NH2-30wt%o 30 35/1 - - - 311 8271 - - 26.6 [145]
AO-PIM-1/UiO-66-NH2-30wt%c 30 35/1 - - - 302 8329 - - 27.6 [145]
AO-PIM-1/UiO-66-NH2-30wt%p 30 35/1 - - - 292 8322 - - 28.5 [145]
(Aged 80 days) 30 35/1 - - 175 - 5630 - 32.0 - [145]
AO-PIM-1b UiO-66 0 35/2 914 150 33 31 986 4.5 29.9 31.8 [184]
AO-PIM-1/UiO-66-10wt%b 10 35/2 1069 206 51 55 1380 4.0 27.1 25.1 [184]
Advanced MOFs
PIM-1 Mg-MOF-74 0 25/2 3537 1072 351 536 6576 3.1 18.7 12.3 [187]
PIM-1/Mg-MOF-74-20wt% 20 25/2 11469 2251 742 1114 21269 3.0 28.7 19.1 [187]
PIM-1 MOF-801 0 35/4 - 844 211 - 4200 4.0 20.0 - [18]
PIM-1/MOF-801-5wt% 5 35/4 - 1752 362 - 9686 4.8 27.0 - [18]
PIM-1 MOF-303 0 35/3 - - 227.5 - 3842.3 - 16.9 - [188]
PIM-1q 0 35/3 - - 251.4 - 3253.1 - 12.9 - [188]
PIM-1/MOF-303-10wt% 10 35/3 - - 257.4 - 6602.8 - 25.6 - [188]
PIM-1/MOF-303-10wt%q 10 35/3 - - 280.9 - 6199.2 - 22.1 - [188]
PIM-1 KAUST-7 0 35/2 - - - 307 4145 - - 13.5 [17]
PIM-1o 0 35/2 - - - 316 3945 - - 12.5 [17]
PIM-1/KAUST-7 -30wt% 30 35/2 - - - 250 6230 - - 24.9 [17]
PIM-1/KAUST-7 -30wt%o 30 35/2 - - - 252 5928 - - 23.5 [17]
PIM-1b MUF-15-Bulk 0 20/1 4230 - 650 1490 11840 - 18.2 7.9 [189]
PIM-1/MUF-15-Bulk-15wt%b 15 20/1 8410 - 1530 3460 23380 - 15.3 6.8 [189]
PIM-1/MUF-15-NS-5wt%b MUF-15-NS 5 20/1 5670 - 810 1870 16120 - 19.8 8.6 [189]
PIM-1/MUF-15-NS-5wt%b,d 5 20/1 - - 670 - 13180 - 19.6 - [189]
(Aged 35 days) 5 20/1 4100 - 400 770 9620 - 24.2 12.6 [189]
HMDA-PIM-1 - 25/1 589 145 36 48 620 4.0 17.0 12.9 [190]
ZIF-8/HMDA-PIM-1/ZIF-8 ZIF-8 - 25/1 316 73.9 19.3 27.9 412 3.8 21.3 14.8 [190]
HKUST-1/HMDA-PIM-1/HKUST- HKUST-1 - 25/1 360 75 18 22 453 4.1 25.0 21.1 [190]
1
PIM-1 0 25/1 2020 907 314 541 5940 2.9 18.9 11.0 [153]
PIM-1r 0 25/1 4710 2130 773 1300 12800 2.8 16.6 9.8 [153]
PIM-1/MIL-101-28.6wt% MIL-101 28.6 25/1 - 3470 1430 2750 22000 2.4 15.4 8.0 [153]
PIM-1/MIL-101-28.6wt%r 28.6 25/1 - 6540 2320 3040 35600 2.8 15.3 11.7 [153]
PIM-1/NanoMIL-101-28.6wt% NanoMIL-101 28.6 25/1 4180 1700 549 894 10600 3.1 19.3 11.9 [153]
PIM-1/NanoMIL-101-28.6wt%r 28.6 25/1 10800 4260 1310 1850 23200 3.3 17.7 12.5 [153]
Photo-responsive MOFs
PIM-1 - 0 37/4 - - 297 - 3882 - 13.0 - [202]

71
PIM-1d - 0 37/4 - - 321 - 4683 - 14.5 - [202]
PIM-1/Azo-DMOF-1-5wt% Azo-DMOF-1 5 37/4 - - 361 - 6626 - 18.4 - [202]
PIM-1/Azo-DMOF-1-5wt%d 5 37/4 - - 376 - 7391 - 19.6 - [202]
PIM-1/Azo-DMOF-1-10wt% 10 37/4 - - 433 - 8095 - 18.7 - [202]
PIM-1/Azo-DMOF-1-10wt%d 10 37/4 - - 371 - 7500 - 20.2 - [202]
PIM-1b UiO-66 0 25/1.4 - - 536 - 7500 - 14.0 - [203]
PIM-1/UiO-66-10wt%b UiO-66 10 25/1.4 - - 959 - 13425 - 14.0 - [203]
PIM-1/Azo(16.7)-UiO-66-10wt%b Azo(16.7)-UiO-66 10 25/1.4 - - 789 - 12225 - 15.5 - [203]
PIM-1/Azo(33.3)-UiO-66-10wt%b Azo(33.3)-UiO-66 10 25/1.4 - - 735 - 11325 - 15.4 - [203]
PIM-1/Azo(66.7)-UiO-66-10wt%b Azo(66.7)-UiO-66 10 25/1.4 - - 639 - 11175 - 17.5 - [203]
PIM-1/Azo(100)-UiO-66-10wt%b Azo(100)-UiO-66 10 25/1.4 - - 574 - 11250 - 19.6 - [203]
Task-specific ionic liquid (TSILs) modified MOFs
PIM-1 - 0 25/1 - - 288 371 5181 - 18.0 13.9 [147]
PIM-1/ZIF-67-10w% ZIF-67 10 25/1 - - 1426 1797 13255 - 9.3 7.4 [147]
PIM-1/TMGHPhO@ZIF-67-10wt% TMGHPhO@ZIF-67 10 25/1 - - 509 822 7145 - 14.0 8.7 [147]
PIM-1/TMGHIM@ZIF-67-10wt% TMGHIM@ZIF-67 10 25/1 - - 936 1218 12848 - 13.7 10.5 [147]
PIM-1 [HDBU][Im]@ZIF-67 0 30/1 - - 520 450 4053 - 7.8 9.0 [148]
PIM-1/[HDBU][Im]@ZIF-67-5wt% 5 30/1 - - 459 567 5100 - 11.1 9.0 [148]
PIM-1 [Bmim][Tf2N-]@UiO-66- 0 20/1 - - 686 1121 7062 - 10.3 6.3 [207]
NH2
PIM-1/[Bmim][Tf2N-]@UiO-66- [Bmim][Tf2N-]@UiO-66- 10 20/1 - - 368 672 8268 - 22.5 12.3 [207]
NH2-10wt% NH2
PIM-1 [Bmim][Tf2N-]@MOF-801 0 35/4 - - 205 - 4110 - 20.0 - [208]
PIM-1/[Bmim][Tf2N-]@MOF-801- [Bmim][Tf2N-]@MOF-801 5 35/4 - - 325 - 9420 - 29.0 - [208]
5wt%
PIM-1d - 0 25/2 - - 124 - 3059 - 24.8 - [209]
PIM-1s 0 25/2 - - 90.1 - 2537 - 28.2 - [209]
PIM-1/UiO-66-20wt%d UiO-66 20 25/2 - - 408 - 8968 - 22.0 - [209]
PIM-1/UiO-66-20wt%s UiO-66 20 25/2 - - 290 - 6709 - 23.1 - [209]
PIM-1/[NH2PMIM][Tf2N]@UiO- [NH2PMIM][Tf2N]@UiO- 20 25/2 - - - - 11427 - 31.8 - [209]
66-20wt%d 66
PIM-1/[NH2PMIM][Tf2N]@UiO- [NH2PMIM][Tf2N]@UiO- 20 25/2 - - - - 13779 - 35.2 - [209]
66-20wt%s 66
2D-MOFs
PIM-1d NUS-8 0 25/1 - - 231 - 3895 - 15.6 - [191]
PIM-1/NUS-8-2wt%c 2 25/1 - - 252 - 6725 - 26.8 - [191]
PIM-1c 0 25/1 - - - 288 4020 - - 13.9 [191]
PIM-1/NUS-8-2wt%d 2 25/1 - - - 217 6462 - - 30.1 [191]
PIM-1b,q NUS-8-NH2 0 25/2 - - 475 - 7826 - 16.5 - [192]
PIM-1/NUS-8-NH2-10.4wt%b,q 10.4 25/2 - - 502 - 14638 - 29.2 - [192]
PIM-1/NUS-8-NH2-13wt%b,q 13 25/2 - - 594 - 14622 - 24.7 - [192]
PIM-1b,o 0 25/2 - - - 660 6325 - - 9.6 [192]
PIM-1/NUS-8-NH2-10.4wt%b,o 10.4 25/2 - - - 1505 10817 - - 7.2 [192]
PIM-1/NUS-8-NH2-13wt%b,o 13 25/2 - - - 1661 13180 - - 7.9 [192]
PIM-1c,d NUS-8-COOH 0 25/2 - - 350 - 7000 - 20.0 10.0 [194]

72
PIM-1/NUS-8-COOH-2wt%d 2 25/2 - - 372 - 11512 - 31.0 - [194]
PIM-1/NUS-8-COOH-5wt%d 5 25/2 - - - - 14652 - - 17.0 [194]
PIM-1d BCoC-ZIF-12h 0 25/2 - - - - 3000 - 23.0 - [211]
PIM-1/BCoC-ZIF-12h-10wt%d 10 25/2 - - - - 7325 - 32.5 - [211]
(Aged 70 days) 10 25/2 - - - - 5192 - 36.6 - [211]

Microporous organic polymers (MOP) materials


PIM-1 SNW-1 0 30/2 - - 223 346 3672 - 16.5 10.6 [214]
PIM-1/SNW-1-10wt% 10 30/2 - - 333 559 7553 - 22.7 13.5 [214]
PIM-1 TpTta-COF 0 25/3 - - 231 - 4018 - 17.4 - [215]
PIM-1/TpTta-COF-6wt% 6 25/3 - - 363 - 9534 - 26.3 - [215]
PIM-1 MAPDA 0 25/3 - - 196 - 3695 - 18.9 - [216]
PIM-1/MAPDA-15wt% 15 25/3 - - 329 - 7862 - 23.9 - [216]
PIM-1 CPSF-EtO 0 35/5 - 566 252 236 2340 2.2 9.3 6.2 [217]
PIM-1/CPSF-EtO-7wt% 7 35/5 - 742 286 380 3235 2.6 11.3 8.5 [217]
PIM-1b FCTF-1 0 35/1 - - - - 5774 - - 11.5 [220]
PIM-1/FCTF-1-2wt%b 2 35/1 - - - - 7287 - - 16.6 [220]
PIM-1 - 0 25/6 - - 166 183 2989 - 18.0 16.3 [221]
PIM-1/TPFC-15wt% TPFC- 15 25/6 - - 181 214 4700 - 26.0 22.0 [221]
PIM-1/TPFC-CH2NH2-15wt% TPFC-CH2NH2 15 25/6 - - 168 212 7730 - 45.9 36.4 [221]
PIM-1/TPFC-CH2PEI-15wt% TPFC-CH2PEI 15 25/6 - - 144 208 5000 - 34.8 24.0 [221]
PIM-1 PAF-1 0 25/1 5343 - 623 991 12355 - - 12.5 [227]
PIM-1/PAF-1-10wt% 10 25/1 7066 - 638 902 12354 - - 13.7 [227]
PIM-1THFb - 0 -/2 4500 1300 330 410 7460 4.0 22.9 18.1 [231]
PIM-1THF/OH-pDCX-5wt%b OH-pDCX 5 -/2 5230 1470 300 380 8510 4.8 28.0 22.4 [231]
PIM-1D/Mb - 0 -/2 4900 2020 650 1020 11430 3.1 17.5 11.2 [231]
PIM-1D/M/pDCX-5wt%b pDCX 5 -/2 9710 3800 1130 1650 20550 3.4 18.2 12.4 [231]
cPIM-1 PPNs prepared from 4,4′- 0 25/2 3185 284 107 135 3739 2.7 27.7 34.9 [101]
diacetylbiphenyl
cPIM-1/PPN-1wt% 1 25/2 6310 1421 311 314 5481 4.6 17.0 17.0 [101]
cPIM-1/PPN-3wt% 3 25/2 5503 2008 474 519 11511 4.2 22.0 24.0 [101]
PIM-1c Network PIM-1 0 25/2 - - - 165 2100 - - 12.7 [232]
PIM-1b,c 0 25/2 - - - 452 5917 - - 13.1 [232]
PIM-1/networkPIM-1-0.5wt%c 0.5 25/2 - - - 486 6470 - - 13.3 [232]
PIM-1/networkPIM-1-0.5wt%b,c 0.5 25/2 - - - 679 9778 - - 14.4 [232]
PIM-1/networkPIM-1-10wt%c 10 25/2 - - - 200 3700 - - 18.5 [232]
PIM-1/networkPIM-1-10wt%b,c 10 25/2 - - - 214 4600 - - 21.5 [232]
PIM-1b - - 25/3 - - - 452 5920 - - 13.1 [233]
grafted-LCD-network-PIM-1b - - 25/3 - - - 1051 12510 - - 11.9 [233]
PIM-1 β-CD 0 35/3.5 1784 590 182 218 3364 3.2 18.5 15.4 [234]
PIM-1/β-CD-0.5wt% 0.5 35/3.5 2994 994 214 315 5435 4.6 25.4 17.2 [234]
PIM-1 4-tert-butylcalix[4]arene, 0 35/3.5 1738 510 149 198 3127 3.4 21.0 15.8 [235]
PIM-1/CA-2wt% CA 2 35/3.5 2375 701 181 232 4469 3.9 24.8 19.3 [235]
a 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. b Methanol treated membranes. c Mixed gas composition CO2:CH4 50:50. d Mixed gas composition CO2:N2 50:50. e Mixed gas composition CO2:N2 15:85.
f Mixed gas composition CO2:N2 10:90. g Mixed gas composition N2:CO2:O2 80:10:10. h Mixed gas composition CO2:CH4 52.1:47.9. i Mixed gas composition CO2:N2:H2O 14:81:5. j Mixed gas composition

73
CO2:N2:SO2 15:85:800ppm. k Mixed gas composition CO2:N2:Ar 20:20:60. l Mixed gas composition CO2:N2 14:86. m Mixed gas composition CO2:N2 30:70. n Mixed gas composition CO2:N2 70:30. o
Mixed gas composition CO2:CH4 30:70. p Mixed gas composition CO2:N2 70:30. q Mixed gas composition CO2:N2 20:80. r Ethanol treated membranes. s Mixed gas composition CO2:N2 50:50, steam
content 3.1 mol%. t loading ratio. u vol%.

74
3.5. Post-modification on membranes

Post-modification on membranes is another strategy for tuning the membrane properties. Table 5

shows the post-modification method on PIMs membranes, alongside its corresponding separation

performance. For example, thermal cross-linking of PIM-1 membranes at elevated temperature

(e.g., 300-450 °C) with prolonged treatment duration (e.g., 24-48 h) formed stable bulky triazine

rings with enhanced gas separation performance (Figure 12a) [236-238]. Another work on thermal

oxidative cross-linked PIM-1 membranes at 385 °C for 24 h demonstrated a superior 5.5-fold

improvement in CO2/CH4 selectivity [237]. He et al. presented a short-duration (e.g., 310 min)

thermal-induced cross-linking of PIM-1, in which the resultant membranes demonstrated

promising CO2/CH4 selectivity of 197 [238]. In another study, Zhang et al. synthesized PIM-TB

copolymer (e.g., CA-PIM-1 and CA-PIM-2) with different amounts of carboxylic acid side groups

using two diamine triptycenes monomers, i.e., 2,6-diaminotriptycene (DAT) and 2,6-

diaminotriptycene-14-carboxylic acid (DATCA), and subjected the CA-PIM to decarboxylation

cross-linking [239]. A more open polymer structure was obtained after cross-linking, resulting in

a remarkable enhancement of permeability for all gases. It was also found that the lowest thermal

treatment temperature was 350 °C to suppress CO2 plasticization. Post cross-linking strategy with

electro-donating polycyclic aromatic hydrocarbons, namely pyrene or 1-aminopyrene was

performed to modify the gas sorption properties and physical characteristics of the PIM-1

membrane [240].

Amino-decorated membranes are favorable to regulate the gas transport property of membranes

through the enhanced preferential sorption with CO2, thereby approaching efficient carbon capture.

Amination of PIM-1 membranes can be achieved via simple chemical vapor deposition at mild

conditions [241]. Nucleophilic substitution occurs at this stage and initiates the reaction of amines

75
towards ether and halogen groups by grafting, ring-opening, chain-scission, terminal replacement,

and cross-linking of the polymer. In a recent work, a surface amination strategy by diazonium-

induced anchoring process was introduced to coat a layer of polyaminophenylene (PAP) on PIM-

1 membranes to improve the surface sorption selectivity of the membranes [242]. Prolong

modification duration was found to increase the mass transport resistance and surface non-

uniformity due to a high degree of amination with extensive attachment of noncovalently grafted

aniline groups on the membrane surface. Further treatment with methanol at a short duration (e.g.,

10 s) can clean the redundant polyaminophenylene and improve the homogeneity on the membrane

surface. At optimal modification parameters, the resultant membrane demonstrated superior

separation performance that surpassed the 2019 Robeson upper bound, with CO2 permeability of

11421 Barrer and CO2/N2 selectivity of 34.1.

Ultraviolet (UV) surface modification has been introduced since the early start of polymer

membrane separation back in 1980s [243]. Li et al. reported a molecular rearrangement method

on the spiro-carbon center of PIM-1 polymer by UV light radiating the membranes at λ = 254 nm

[244]. Upon UV radiation, the PIM-1 polymer chains underwent a 1,2-migration reaction

(homolytic cleavage) and transformed into a close-to-planar like rearranged structure. Prolonged

exposure to UV treatment significantly reduced the fractional free volume and size of micro-pores,

as a consequence of the destructed spiro-carbon center, which resulted in enhanced selectivity of

resultant membranes. Song et al. proved the molecular transformation of PIM-1 after photo-

oxidative surface modification, in which oxidative chain scission and local densification at the

membrane surface occurred upon short wavelength UV irradiation [245]. They also revealed that

membranes sandwiched in quartz glass during UV irradiation limited the source of O2 at the

interface which affected the reaction kinetics and reduced the gas transport performance. In both

76
cases, the PIM-1 materials with high FFV behaved as nanoreactors attacked by the shallowly

penetrating UV light. The successful UV rearrangements can be noticed with the reduction in PIM-

1 membrane molecular weight [227], micropore size, and fractional free volume [246], as well as

the appearance of a new carbonyl or hydroxyl functional groups in the polymer matrix, which

favor interaction with CO2 molecules [245]. The reduction in free volume up to a certain extent

also contributed to suppressing the physical aging of the PIM-1 membranes [247].

Supercritical carbon dioxide (scCO2) can be used to treat finished membranes to alter the gas

permeation performance. scCO2 can solvate the polymer chains which enhances the chain mobility

and rearrangement [248]. Swelling of membranes in scCO2 occurred at high temperature and

pressure (e.g., 130 °C, 450 atm [249]), and a subsequent depressurization resulted in membranes

with different densities and free volumes. Scholes et al. investigated the effect of scCO2 exposure

duration (e.g., 2 or 8 h) and depressurization rates (e.g., 1.7 atm/min or 118 atm/min) on the gas

separation and aging properties of PIM-1 membranes [248]. It was found that all scCO2 treated

membranes transformed to denser membrane morphology and experienced a reduction in gas

permeability as a result of increased density. In contrast to N2, CH4 and O2 permeabilities, the

permeabilities of He and CO2 were noticed to decrease over extended periods as a result of micro-

cavities contraction within PIM-1. Starannikova et al. also conducted scCO2 treatment on PIM-1

membranes, and a similar trend on the permeability values was observed [249]. Additionally, they

reported a higher reduction of permeability on strain aging (e.g., edge of membranes fixed with O-

rings and subjected to annealing), as a result of radial mechanical strains appearing in the shrinking

film upon annealing.

Alcohol treatment is a sustainable regeneration method to restore the aged membrane. The physical

aging effect in PIM-1 films can be reversed by methanol wash as observed for PTMSP [250]. With

77
alcohol treatment, it released the stresses of the polymer chain to a more relaxed state, restoring

the high free fraction volume and subsequently permeability of the membranes [251]. Detailed

long-term sorption kinetics and permeation studies on aged PIM-1 membranes revealed that the

decrease of permeability during aging was hardly affected by solubility, but was attributed to the

diffusion coefficient of permanent gases [252].

Recently, atomic layer deposition (ALD) technique was introduced on PIM-1 to tune the

microporosity and gas separation properties of resultant membranes at the atomic level [253, 254].

An ALD cycle typically utilized short vapor pulses to tailor the microporosity of polymers,

followed by water as the second precursor loaded at similar procedures of metalorganic precursor

with pulse, exposure and purge time of 0.1, 8 and 25 s, respectively (Figure 12b) [254]. Chen et

al. used trimethyl aluminum (TMA) as the metalorganic precursors for the deposition of Al2O3 on

the micropores and bulky PIM-1 membrane surface [253]. The Al2O3 distribution and

concentration on the membrane surface across different ALD cycles were confirmed by the Al and

N elemental Energy-dispersive X-ray spectroscopy (EDS) mapping (Figure 12c-f). Al2O3 cluster

concentration was observed to increase sharply upon six ALD cycles, while a further increase in

the number of ALD cycles (e.g., nine or twelve cycles) eventually increased the cluster domain

size, forming a bi-continuous phase of Al2O3 and PIM-1 (Figure 12f). The PIM-1 membrane that

underwent six ALD cycles (e.g., PIM-1-Al2O3-6) showed a uniform pore size of ~6.2 Å, achieving

CO2/CH4 selectivity improvement of 250%. Notably, both the permeability and selectivity

decreased sharply upon nine to twelve ALD cycles, which were attributed to the irregular pore size

distribution and severe blockage of micropores. Niu et al. further conducted a study to clarify the

relationships between microstructural properties and CO2 separation performance on Al2O3-, ZnO-

and TiO2-modified PIM-1 membranes [254]. Experimental studies and density functional theory

78
(DFT) calculations revealed that the loading amount, location of metal clusters as well as

interactions between metal clusters with CO2 molecules all contributed to the enhancement in CO2

separation performance. Both the Al2O3- or ZnO-ALD modified PIM-1 showed diffusion-

dominated permeation behavior, while TiO2-ALD modified PIM-1 was sorption-dominated. A

decrease in permeability was observed in most of the ALD-modified PIM-1 membranes. This

phenomenon was ascribed to the metal oxide nucleation-induced pore-filling effect which

narrowed the micropores of the membranes, thus increasing the transport resistance and tortuosity

of CO2 diffusion.

Post-synthetic reconstruction of MMMs is proposed to further enhance the separation of

membranes even after the incorporation of functional fillers. Hao et al. reported a photo-oxidation

treatment on PIM-1/ZIF-71 MMMs [255]. The presence of ZIF-71 enhanced the CO2 permeability

up to 2.5-fold, while the UV-treated MMMs demonstrated a 3-fold CO2/CH4 selectivity

improvement. The former enhancement was attributed to the intrinsic properties of ZIF-71 as filler,

while the latter was due to densification at the surface of the membrane after UV irradiation that

increased the gas transport resistance. The UV-treated PIM-1/ZIF-71-30wt% MMMs showed a

comparable CO2 permeability with pristine PIM-1 with CO2/N2 and CO2/CH4 selectivities

increased to 26.9 and 35.6, respectively. Hou et al. reported a post-UV irradiation treatment on

PIM-1/PAF-1 MMMs [227]. The UV treatment was carried out by exposing both sides of the

membranes to air under the UV lamp for 30-270 min. The resulting membranes that benefited

from the synergistic effect between highly porous PAF-1 and UV irradiation were found to have

~33% higher permeability compared to the UV-treated PIM-1 and achieved CO2/CH4 selectivity

of >40, near the 2019 Robeson line. Huang et al. reconstructed the PIM-1/ZIF-8 MMMs by in-situ

chemical vapor aminolysis to improve the polymer-filler compatibility, chemical property of

79
transport channel, and preferential sorption within the MMMs (Figure 12g) [256]. Aminolysis was

carried out under mild conditions, in which the amines can be easily substituted to the functional

groups in PIM-1 and chelated with the metal centers of MOFs. The resultant MMMs showed CO2

permeability of 6632 Barrer, along with CO2/N2 selectivity of 25.4. This study also demonstrated

the universality of the in-situ vapor aminolysis strategy in a series of MMMs.

a)

b)

c) d) g)

e) f)

80
Figure 12. (a) Possible structure variation of polymer chain after thermal treatment of PIM-1

under N2 atmosphere [238]. (b) Preparation of ALD-modified PIM-1 membranes [254]. Al and

N distribution on the surface of PIM-1 after c) 3, d) 6, e) 9, and f) 12 Al2O3 ALD cycles using

EDS mapping [253]. (g) In-situ chemical vapor aminolysis on PIM-1/ZIF-8 MMMs [256].

Copyright 2020, 2021, 2022 and 2023. Reproduced with permission from Wiley-VCH and

Elsevier.

Table 5. Gas separation performance of post-modifications of PIMs membranes.

T/P Permeability (Barrer)a Selectivity


Membrane ID Modification method Ref.
(°C/atm) H2 O2 N2 CH4 CO2 O2/N2 CO2/N2 CO2/CH4
PIM-1 Pristine 35/3.5 2696 712 166 204 3375 4.3 16.6 20.4 [236]
PIM-1-300°C-48h Thermal oxidative cross- 35/3.5 3872 582 96 73 4000 6.1 54.8 41.7 [236]
linking (300 °C, 48 h)
PIM-1 Pristine 22/4 3408 1135 356 397 5135 3.2 14.4 12.9 [237]
TOX-PIM-1-385°C-24h Thermal oxidative 22/4 1820 245 30 16 1100 8.1 37.0 69.0 [237]
crosslinking (385 °C, 1
mbar, 24 h)
PIM-1-450°C-310min Thermal crosslinking 35/3.5 234 - - 6 1239 - - 197.0 [238]
(450 °C, 310 min, N2)
CA-PIM-1 (DAT/DATCA = Pristine 35/1 6527 1327 365 410 6831 3.6 18.7 16.7 [239]
9:1)
CA-PIM-1-350°C-2h Decarboxylation cross- 35/1 8394 1693 497 606 8602 3.4 17.3 14.2 [239]
linking @ 350 °C, 2 h
CA-PIM-1-375°C-2h Decarboxylation cross- 35/1 9012 2326 685 828 9671 3.4 14.1 11.7 [239]
linking @ 375 °C, 2 h
CA-PIM-2 (DAT/DATCA = Pristine 35/1 5724 1185 318 327 5946 3.7 18.7 18.2 [239]
8:2)
CA-PIM-2-350°C-2h Decarboxylation cross- 35/1 7262 1552 474 534 7668 3.3 16.2 14.4 [239]
linking @ 350 °C, 2 h
CA-PIM-2-375°C-2h Decarboxylation cross- 35/1 8035 2128 596 661 8512 3.6 14.3 12.9 [239]
linking @ 375 °C, 2 h
PIM-1 Pristine 25/1 2631 - 578 832 7842 - 13.6 9.7 [241]
PIM-A-2 Amination with 25/1 1135 - 85 229 2590 - 30.8 13.4 [241]
ethylenediamine (2 h)
PIM-A-2.5 Amination with 25/1 662 - 32 86 1103 - 35.0 12.8 [241]
ethylenediamine (2.5 h)
PIM-1 Pristine 25/1 - - 321 - 4500 - 14.0 - [242]
PIM-1-PAP24 Amination with PAP (24 25/1 - - 335 - 11422 - 34.1 - [242]
h) + soaking in methanol
(10 s)
PIM-1 Pristine 35/3.5 3731 1172 309 431 6601 3.8 21.4 15.3 [244]
PIM-1-UV20min-quartz UV (λ = 254 nm) in air, 35/3.5 2818 416 74 62 1869 5.7 25.4 30.1 [244]
sandwiched in quartz, 20
min
PIM-1-UV20min-quartzb UV (λ = 254 nm) in air, 35/3.5 - - - 61 1554 - - 25.4 [244]
sandwiched in quartz, 20
min
PIM-1-UV20min-quartzc UV (λ = 254 nm) in air, 35/3.5 - - - 83 745 - - 9.1 [244]
sandwiched in quartz, 20
min
PIM-1-UV30min-quartz UV (λ = 254 nm) in air, 35/3.5 2247 189 28 23 724 6.8 26.1 31.1 [244]
sandwiched in quartz, 30
min

81
PIM-1-UV30min-quartzb UV (λ = 254 nm) in air, 35/3.5 - - - 20 724 - - 29.3 [244]
sandwiched in quartz, 30
min
PIM-1-UV30min-quartzc UV (λ = 254 nm) in air, 35/3.5 - - - 34 367 - - 10.9 [244]
sandwiched in quartz, 30
min
PIM-1 Pristine 22/4 3195 1089 325 418 5622 3.4 17.3 13.5 [245]
PIM-1-UV20min-air UV (λ = 254 nm) in air, 22/4 2103 736 198 228 4374 3.7 22.1 19.2 [245]
20 min
PIM-1-UV30min-air UV (λ = 254 nm) in air, 22/4 1666 317 58 62 1555 5.5 26.9 25.00 [245]
30 min
PIM-1-UV20min-quartz UV (λ = 254 nm) in air, 22/4 2609 839 146 161 3781 5.2 27.7 25.1 [245]
sandwiched in quartz, 20
min
PIM-1-UV30min-quartz UV (λ = 254 nm) in air, 22/4 2104 513 84 85 2394 6.1 28.4 28.1 [245]
sandwiched in quartz, 30
min
PIM-1 Pristine, MeOH treated, 35/8 - 1865 349 1307 7595 5.3 21.8 5.8 [248]
aged 7 days
PIM-1-scCO2-2h-rapid scCO2, 2 h, 35/8 - 1245 314 823 5118 4.0 16.3 6.2 [248]
depressurization rate 118
atm/min, MeOH treated,
aged 7 days
PIM-1-scCO2-2h-slow scCO2, 2 h, 35/8 - 1577 338 887 7068 4.7 20.9 8.0 [248]
depressurization rate 1.7
atm/min, MeOH treated,
aged 7 days
PIM-1-scCO2-8h-rapid scCO2, 8 h, 35/8 - 1444 338 788 6621 4.3 19.6 8.4 [248]
depressurization rate 118
atm/min, MeOH treated,
aged 7 days
PIM-1-scCO2-8h-slow scCO2, 8 h, 35/8 - 1513 338 857 6989 4.5 20.7 8.2 [248]
depressurization rate 1.7
atm/min, MeOH treated,
aged 7 days
PIM-1 Pristine 20/1 3280 1590 610 1130 11540 2.6 18.9 10.2 [249]
PIM-1-EtOH EtOH treated 2 days 20/1 4720 2465 1040 1890 17490 2.4 16.8 9.3 [249]
PIM-1-scCO2-4h-rapid scCO2, 4 h, 20/1 4530 2110 800 1370 14900 2.6 18.6 10.9 [249]
depressurization rate 450
atm/min
PIM-1-t,str strained aging, 100 °C, 4 20/1 2360 745 190 260 4850 3.9 25.5 18.7 [249]
h
PIM-1-t,fr free aging, 100 °C, 4 h 20/1 2930 1080 335 530 7190 3.2 21.5 13.6 [249]
PIM-1 Pristine 35/3.5 4381 1686 434 527 8287 3.9 19.1 15.7 [253]
PIM-1-Al2O3-3 Al2O3 ALD, 3 cycles 35/3.5 3480 1141 264 279 5577 4.3 21.1 20.0 [253]
PIM-1-Al2O3-6 Al2O3 ALD, 6 cycles 35/3.5 2492 188 21.3 11.1 624 8.8 29.3 56.2 [253]
PIM-1-Al2O3-9 Al2O3 ALD, 9 cycles 35/3.5 287.9 17 4 2 44 4.7 12.5 21.8 [253]
PIM-1-Al2O3-12 Al2O3 ALD, 12 cycles 35/3.5 16 2 0.6 0.4 6 3.0 10.5 15.8 [253]
PIM-1 Pristine 25/1.5 - - 410 550 6606 - 16.1 12.0 [254]
PIM-1-Al2O3-10 Al2O3 ALD, 10 cycles 25/1.5 - - 204 184 3245 - 15.9 17.6 [254]
PIM-1-Al2O3-20 Al2O3 ALD, 20 cycles 25/1.5 - - 2 1 35 - 14.0 25.9 [254]
PIM-1-Al2O3-30 Al2O3 ALD, 30 cycles 25/1.5 - - 1 1 3 - 1.9 2.7 [254]
PIM-1-ZnO-10 ZnO ALD, 10 cycles 25/1.5 - - 318 472 5946 - 18.7 12.6 [254]
PIM-1-ZnO-20 ZnO ALD, 20 cycles 25/1.5 - - 295 364 4690 - 15.9 12.9 [254]
PIM-1-ZnO-30 ZnO ALD, 20 cycles 25/1.5 - - 195 211 3019 - 15.5 14.3 [254]
PIM-1-TiO2-10 TiO2 ALD, 10 cycles 25/1.5 - - 485 644 7666 - 15.8 11.9 [254]
PIM-1-TiO2-20 TiO2 ALD, 20 cycles 25/1.5 - - 577 770 9005 - 15.6 11.7 [254]
PIM-1-TiO2-30 TiO2 ALD, 30 cycles 25/1.5 - - 688 943 10660 - 15.5 11.3 [254]
PIM-1 Pristine 35/3.5 - 562 164 322 3295 3.4 20.1 10.2 [255]
PIM-1-UV40min UV (λ = 254 nm), 40 min 35/3.5 - 277 41 36 1233 6.7 29.8 34.1 [255]
PIM-1/ZIF-71-30wt% 30 wt% ZIF-71 35/3.5 - 1602 457 750 8377 3.5 18.3 11.2 [255]
PIM-1/ZIF-71-30wt%- 30 wt% ZIF-71 + UV (λ 35/3.5 - 807 129 97 3459 6.3 26.9 35.6 [255]
UV40min = 254 nm), 40 min
PIM-1/ZIF-71-30wt%- 30 wt% ZIF-71 + UV (λ 35/7 - - - 105 3020 - - 28.8 [255]
UV40minb = 254 nm), 40 min
PIM-1/ZIF-71-30wt%- 30 wt% ZIF-71 + UV (λ 35/7 - - - 93 2629 - - 28.3 [255]
UV40mind = 254 nm), 40 min
PIM-1 Pristine 25/1 5343 - 623 991 12355 - - 12.5 [227]
PIM-1-UV180min UV (λ = 254 nm), 180 25/1 2040 - 33 28 986 - - 35.2 [227]
min

82
PIM-1-UV180minb UV (λ = 254 nm), 180 35/4 - - - 344 5079 - - 14.8 [227]
min
PIM-1/PAF-1-10wt% 10 wt% PAF-1 25/1 7066 - 638 902 12354 - - 13.7 [227]
PIM-1/PAF-1-10wt%- 10 wt% PAF-1 + UV (λ = 25/1 4769 - 68 53 2081 - - 29.3 [227]
UV180min 254 nm), 180 min
PIM-1/PAF-1-10wt%- 10 wt% PAF-1 + UV (λ = 35/4 - - - 474 6700 - - 14.1 [227]
UV180minb 254 nm), 180 min
PIM-1 Pristine 25/1 - - 577 - 7842 - 13.6 - [256]
PIM-1-Am In-situ chemical vapor 25/1 - - 219 - 4327 - 19.8 - [256]
aminolysis
PIM-1/ZIF-71-10wt% 10 wt% ZIF-71 25/1 - - 806 - 12084 - 15.0 - [256]
PIM-1/ZIF-71-10wt%-Am 10 wt% ZIF-71 + In-situ 25/1 - - 261 - 6632 - 25.4 - [256]
chemical vapor
aminolysis
PIM-1/ZIF-71-10wt%-Ame 10 wt% ZIF-71 + In-situ 25/1 - - 254 - 6689 - 26.3 - [256]
chemical vapor
aminolysis
a 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. b Mixed gas composition CO :CH 50:50. c Mixed gas composition CO :CH :H S 50:49.95:0.05.
2 4 2 4 2
d Mixed gas composition CO :CH 30:70. e Mixed gas composition CO :N 50:50.
2 4 2 2

3.6. Physical aging of ladder PIMs membranes

Glassy polymers have emerged as the predominant choice for gas separation membranes, owing

to their remarkable balance between permeability and selectivity. Nevertheless, their physical

aging issue impacts mechanical and optical characteristics, as well as gas separation performance,

such as increased stiffness and brittleness [257]. Physical aging refers to the gradual transition of

a non-equilibrium glassy state towards equilibrium, occurring at temperatures significantly below

the Tg. This transition arises from the excess free volume between polymer chains through local

segment chain motion, which ultimately leads to the reduction of the free volume of the polymer

and membrane permeability [258]. This phenomenon can be observed through changes in the

specific volume and enthalpy of the polymer. While external factors do not influence the onset of

physical aging, their rate is contingent upon variables such as temperature, gas environment, and

polymer structure. Upon aging, the mobility of segmental chains reduced the pores and FFV of the

PIMs matrix [24]. This structural evolution leads to a substantial decline in gas permeability, often

reaching up to 75% within 30-90 days. Consequently, the phenomenon of aging significantly

compromises the separation durability of PIMs, thereby impeding their commercial application

development. Positron annihilation lifetime spectroscopy (PALS) has been used for more precise

free-volume measurement.

83
Numerous strategies have been devised to tackle the challenge of physical aging, in which these

approaches are meant to effectively tighten the polymer chains and impede chain relaxation,

thereby suppressing aging on the membrane. These involve (i) polymer backbone design and

structural changes through the incorporation of highly rigid matrix that capable of providing

intermolecular free volume, accomplished by constructing the framework of PIM with shape-

persistent bridged bicyclic units (such as triptycenes and pentiptycenes), developing novel

monomers, or adding substituents or nitro and hydroxide groups [51, 53, 54, 79, 82, 91, 92, 259];

(ii) post-modification of PIMs structure by altering the polymer main or side chains that result in

a more rigid matrix through methods such as hydrogen bonding, cross-linking, or functionalization

[101, 104-107]; (iii) copolymerization with PIMs or blending of PIMs with other polymers of

lower free volume [119]; (iv) embedment of nanofillers into the polymer [15, 16, 24, 132, 137,

138, 141, 153, 155, 161, 165, 178, 182, 183, 221, 230, 231, 233-235]; and (v) post-modification

of PIMs membranes with thermal, chemical or UV treatment as well as scCO2 exposure, inducing

chain rearrangement or reduce free volume of the membranes [236-238, 246, 248, 249].

Furthermore, treating membranes with alcohol promote chain relaxation and rejuvenate aged

membranes [250, 251]. Each of these promising approaches and their corresponding impact on the

gas separation and aging performances of the membrane has been comprehensively discussed in

the previous section, which provides insights into the sustained research and development of PIMs-

based membrane separation technology for viable commercial products and processes.

3.7. Summary on the progress of ladder PIMs in CO2 capture

The excellent intrinsic properties of PIMs have flourished in their implementation in the

membrane-based separation process, especially on CO2 capture. Particularly, the micro-structure

architecture on ladder PIMs polymer is discussed from five different aspects, each of them

84
demonstrating fascinating separation performance in CO2 capture. The selection or incorporation

of rigid and bulky groups played crucial roles in determining the structural properties such as FFV

and the resultant performance of these membranes. Despite challenges such as physical aging that

caused a reduction in separation performance, recent advancements in the architecting of polymer

backbones and synthesis techniques offered exciting opportunities for further improving the gas

separation performance of these membranes.

Post-modification on PIMs polymer structures offers a versatile approach to tailor their properties

for enhanced gas separation capabilities based on tuning of polymer structure. Functionalization

techniques such as conversion of nitrile groups, incorporation of bulky side groups, and

introduction of functional moieties like amidoxime and carboxyl groups have been explored to

enhance selectivity, permeability, and mechanical properties of PIMs membranes. These

modifications enable fine-tuning of pore size, free volume, and surface chemistry, thereby

optimizing gas separation performance. Meanwhile, notable improvements in CO2 permeability,

selectivity, and mechanical properties have been witnessed through strategic combinations with

various polymers and copolymers. The miscibility between blend components, alongside

modifications like cross-linking and copolymerization, requires careful investigation in optimizing

the membrane performance.

MMMs combine the synergistic properties of both polymer and fillers revealing a potential

approach to ameliorating the gas separation performance. A total of eight types of nanofillers have

been explored and scrutinized, comprising the porous and non-porous materials. Non-porous

materials enhanced the membrane performance by creating a tortuous pathway for gas transport,

while porous materials aided the separation by providing precise molecular sieves, selective

adsorption and faster gas diffusion. Additionally, the integration of fillers into the PIMs matrix

85
adjusted the contorted molecular configuration and restrained the polymer chain rearrangement,

thus effectively suppressing physical aging in high FFV polymers. Based on a comprehensive

study of the state-of-the-art PIMs-based MMMs, most of the high separation performance of the

MMMs was found to be contributed by the incorporation of fillers with functional polar groups

(e.g., ‒NH2 or ‒CN groups). The presence of polar functionalities in fillers effectively enhanced

the interfacial compatibility of PIMs-filler through hydrogen bonding and dipole interactions,

while the incorporation of these functionalized fillers into the polymer matrix led to the membrane

with high affinity and strong dipole-quadrupole interactions with CO2 molecules. The 2D structure

(e.g., MXene) further created a tortuous pathway for the transport of gas molecules which

consequently improved the transport resistance and gas-pair selectivity. Meanwhile, nano-sized

MIL-101 and Mg-MOF-74 could be considered in regards to commercial application which

requires high CO2 permeability at comparable selectivity. All in all, the inherent properties of the

fillers and MMMs have value added to the commercialization of these materials.

Lastly, the investigation established that the improvised gas-pair selectivities can be achieved

through post-modification on PIMs membranes via thermal, UV, scCO2, and ALD treatment, in

which these modifications rearranged and restructured the microporous environment of PIMs and

thus resulted in enhanced transport resistance. Alcohol treatment, on the other hand, could improve

the gas transport performance by releasing the stresses in polymer chains and restoring the FFV.

Figure 13 summarizes the state-of-the-art pure gas performance of ladder PIMs membranes for

CO2/N2 and CO2/CH4 separation. All the data were extracted from Table 1-5. In general, the data

were accumulated towards the right side of the plot, indicating that PIMs-derived membranes were

of superior gas permeability with comparable selectivities. It can be observed that the PIMs-based

MMMs and post-modified PIMs membranes were outperformed in CO2/N2 separation, with a

86
number of PIMs membranes transcending the 2019 pure gas upper bound, being a potential

candidate for commercialization. A number of redesigned PIMs membranes (e.g., TMN-Trip,

HMI-Trip, and DTFM-Btrip) were also showed remarkable CO2/N2 performance with data lying

on the 2019 upper bound. PIM-1 MMMs incorporated with GO nanosheet were reported to pose

superior CO2/N2 selectivity as high as 123. The incorporation of GO nanosheets into PIMs

membranes not only provides specific adsorption affinity to facilitate the CO2 sorption across

membranes, the uniform assembles of GO nanosheets in the polymer matrix also created

hydrophilic/hydrophobic microphase segregation in the membrane which provides additional gas

transport channels. Both these properties have significantly contributed to the excellent

performance of PIM-1/GO MMMs that far surpassed the 2019 Robeson upper bound. Interestingly,

benzotriptycene-PIMs, PIM-TMN-Trip demonstrated the highest CO2 permeability reported thus

far (e.g., 52800 Barrer), in which the excellent permeability was contributed by the solubilizing

tetramethylnaphthalene groups on the polymer backbone. Meanwhile, the majority of the recently

published ladder PIMs membranes were located between 1991 and 2019 upper bound for CO2/CH4,

further demonstrating the great potential of MMMs and post-modified PIMs membranes in natural

gas sweetening applications.

a) 100 1000 10000 b) 100 1000 10000


1000 1000

100 100

100 100
CO₂/CH4 selectivity
CO₂/N₂ selectivity

10 10
10 10

1 1 1 1
100 1000 10000 100 1000 10000
CO₂ permeability (Barrer) CO₂ permeability (Barrer)

87
Figure 13. Pure gas (a) CO2/N2 and (b) CO2/CH4 separation performance of representative PIMs

membranes with 1991, 2008 and 2019 trade-off lines. Data extracted from Table 1–5. ()

represents pristine PIMs membranes, () represents backbone-architected PIMs membranes,

() represents post-structure-modified PIMs membranes, () represents PIMs-based blend

membranes, () represents PIMs-based MMMs, and () represents post-modified PIMs

membranes.

Figure 14 shows the gas separation performance of recently reported ladder PIMs membranes for

binary mixed gas CO2/CH4 separation. The 2018 CO2/CH4 mixed gas upper bound was developed

based on polyimide membranes (mostly 6FDA-derived PIM-PI) tested with a feed composition of

1:1 CO2/CH4 at CO2 partial pressure of 10 bar at 35 °C [7]. The data collected were of various

CO2/CH4 feed compositions and pressures as tabulated in Table 1-5. Notably, most of the recently

reported ladder PIMs membranes have transcended the upper bound, with PIMs-based MMMs

demonstrating the most distinctive separation performance. It was noted that in all cases, the mixed

gas selectivity values were lower than pure gas selectivity values due to competitive sorption. The

relative drop of selectivity is more apparent in highly CO2 permeable PIM materials, especially

under aggressive testing conditions.

88
100 1000 10000
1000 1000
CO₂/CH4 selectivity

100 100

10 10
2018 mixed gas upper bound

1 1
100 1000 10000
CO₂ permeability (Barrer)

Figure 14. Mixed gas CO2/CH4 (solid symbols) separation performance of representative PIMs

membranes with 2018 trade-off lines (10 bar CO2 partial pressure at 35 °C). Open symbols

represent pure gas data for comparison. Data extracted from Table 1–5. () represents pristine

PIMs membranes, () represents backbone-architected PIMs membranes, () represents post-

structure-modified PIMs membranes, () represents PIMs-based blend membranes, ()

represents PIMs-based MMMs, and () represents post-modified PIMs membranes.

4. PIM membrane in light hydrocarbon separation

Million tons of light olefins, such as ethylene (or ethane, C2H4) and propylene (or propane, C3H6)

are produced annually from petrochemical plants and refineries [260]. Daily consumables such as

plastics [261], rubber [262], and fibers [263] are some derivatives of light olefins. It is no

exaggeration to say that light olefins are the backbone of modern life and important raw material

for downstream chemical plants [264, 265]. The separation of olefins (i.e., alkenes) from paraffin

89
(i.e., alkanes) typically relies on high energy-intensive cryogenic distillation, operated under a

pressure of 7-28 bar and temperature of 183-258 K [260]. The energy consumption for the

purification of these ethylene and propylene is substantial, being roughly equivalent to the annual

energy consumption of Singapore, which is 0.3% of global energy use [265, 266]. Nevertheless,

owing to the close properties, molecular sizes, and volatilities (Table 6) [267, 268], their separation

is challenging and thus requires an extreme operation condition. Therefore, alternative technology,

such as membrane separation technology has been proposed for light olefin separation.

Specifically, for the past 30 years, a tremendous amount of studies on paraffin/olefin separation

can be found based on membranes using polymers such as polysulfone (PSf) [269-271],

polyethersulfone (PES) [272, 273], polyimide [263, 274-276], polyphenylene oxide (PPO) [277],

polycarbonate [278], cellulose acetate [279], ethylcellulose [280, 281], polypyrrolone [282], poly

(amide-b-ethylene oxide) (Pebax) [283-285], generating scalable, cost- and energy-efficient

separation technology.

An innovative material of linear PIMs, PIM-PI has emerged as a strategic solution for C2 or C3

separation applications. By modifying the diamine and/or dianhydride components or replacing

one contorted center with another, it becomes possible to enhance gas permeability. The relatively

inflexible and contorted backbones of PIM-PI result in inefficient packing, consequently yielding

a high free volume. This characteristic facilitates improved gas separation efficiency [286].

Compared with PIMs, polyimide endeavors to bolster both inter- and intra-chain interactions via

the incorporation of flexible diamines or hydroxyl functionalization, thereby disrupting chain

mobility [287]. It is noteworthy that the majority of current literature focuses on these low-free

volume materials [288, 289].

90
PIMs materials pose challenges in separating gases with small differences in kinetic diameter in

the separation of paraffin/olefin due to their broad pore size distribution (PSD). Heat-treated

membranes like CMS or TR membranes are extensively employed in this domain to enhance gas

separation performance. Heat treatment facilitates the attainment of the necessary PSD and

stability for light hydrocarbon separation. This is attributed to various reactions including cross-

linking, oxidation, dehydrogenation, and cyclization occurring during the thermal breakdown of

polymer precursor structures, leading to the creation of ultrafine pores. Specifically, heating above

the degradation temperature of the polymer can densify the micropores within the membrane [288,

289]. The pore structure model introduced by Steel and Koros offers insight into the variations

observed in the critical size (dc) of ultramicro pores within the molecular sieving mechanism of

CMS membranes [290].

CMS membranes derived from PIM or PIM-PI have attracted widespread attention due to their

high flux, excellent molecular sieving ability, and stability [22, 291]. The porous nature of CMS

membranes leads to high permeability and superior molecular sieve efficiency [291]. Besides, the

extremely high “entropy selectivity” of the CMS membrane enables “slimmer” ethylene molecules

to better penetrate the rigid diffusion-limiting of the CMS ultramicropores [292]. Furthermore,

CMS membranes demonstrate chemical stability in low-pressure environments with prolonged

exposure to hydrocarbons, owing to their carbonization attributes [293].

Table 6. Key physical and chemical properties of C2 and C3. Data extracted from [260, 267, 268,

294].

Light-Hydrocarbon Ethylene Ethane Propylene Propane


(C2H4) (C2H6) (C3H6) (C3H8)

91
Normal Boiling Point (℃) -103.7 -89 -47.6 -42.0
Normal Melting Point (℃) -169 -184 -185 -188
Kinetic Diameter (Å) 4.16 4.44 4.68 4.30-5.12
Polarizability (10-25 cm3) 42.52 44.3-44.7 62.6 62.9-63.7
Quadupole Moment (10-26 esu cm2) 0 0 0.366 0.084

4.1. Ethylene/ethane separation

Ethylene is one of the most productive products in the chemical industry and is widely utilized as

raw feedstock to produce a variety of organic compounds, such as polyethylene [295].

Nevertheless, limited reports were found on the performance of PIM and its derived materials in

C2H4/C2H6 separation. Table 7 summarizes the performance of PIMs-based membranes for

ethylene/ethane separation.

Salinas et al. prepared CMS membranes derived from a diffusion-selective hydroxyl-

functionalized PIM-PI polymer (e.g., PIM-6FDA-OH) for C2H4/C2H6 separation [288]. Partial

carbonization structure was observed when the membranes were treated at 500 ℃, leading to pore

enlargement that increased the C2H4 permeability by a factor of 50. Further heating to 800 ℃

resulted in pore densification which sintered the dimension of ultramicropores in the carbon matrix,

thereby enhancing the sieving ability of the resultant membrane, achieving C2H4/C2H6 selectivity

to 17.5. However, in the mixed gas test, the selectivity of C2H4/C2H6 declined to 8.3, which may

be related to the competitive adsorption effect and CMS expansion.

For the first time, PIM-1 based CMS membrane was proposed for ethylene/ethane separation

through properly controlling the pyrolysis process to densify the pore size distribution of PIM-1

[296]. Likewise, these PIM-1 derived CMS membranes achieved a superior improvement of 600%

in C2H4/C2H6 selectivity. In a follow-up study, the same group of researchers proposed

spirobisindane-based polyimide of intrinsic microporosity as the precursor to prepare CMS

92
membranes [289]. The membrane displayed a high value of C2H4/C2H6 selectivity of 16 under a

mixed gas test at 20 bar.

In addition, combining metal ions with cPIM is improved to improve the selectivity of olefin

separation, and the resulting metal-carboxyl complex can improve separation performance by

enhancing gas adsorption or promoting gas molecule transport, depending on the type of metal

ion. In particular, boron can affect the pore properties of carbonized membranes, thereby

controlling the microstructure and separation performance of cPIM membranes [297, 298]. Liao

et al. prepared cPIM-1 doped with boron of different molecular sizes as the precursor for CMS

membranes [298]. Results showed that with the addition of boron compounds such as borax,

benzene-1,4-diboronic acid (BDA), 9,9-dihexylfluorene-2,7-diboronic acid (DHFDA), the C2H4

permeability and the C2H4/C2H6 selectivity increased in the order of CB700-DHFDA > CB700-

BDA > CB700-borax > CB700-cPIM, while in the case of CB700-borax, the permeability of C2H6

was the highest at 1.76 Barrer. This can be attributed to the impermeable aromatic units, which

provide additional diffusion pathways for the transport of C2H6 with larger molecular sizes.

Combining the pore size of CMS membranes modified with different borons, it was found that the

C2H4 permeability and C2H4/C2H6 selectivity increase with the increase of molecular size, as the

CMS membrane can further adjust its tortuosity and pore characteristics after being chemically

modified with larger molecular size degradable boron compounds. Unlike other membranes,

CB700-DHFDA and CB700-BDA in mixed gas testing may cause greater diffusion competition

for C2H4 and C2H6 due to their larger pore sizes, resulting in lower permeability and selectivity

than in pure gases.

Recently, Hazazi et al. reported on the synthesis of high-performance and plasticization resistance

CMS membranes from two three-dimensional triptycene-based ladder PIMs precursors, namely

93
Trip(Me2)-TB and EA(Me2)-TB for C2H4/C2H6 separation [299]. The CMS membrane derived

from Trip(Me2)-TB displayed bulkier and superior thermal stability owing to the presence of 3D

triptycene units (Figure 15a). This characteristic led to a looser structure with larger interlayer pore

sizes, consequently resulting in higher C2H4 permeability. Upon heat treatment exceeding 600 oC,

both membranes exhibited a substantial increase in gas selectivity, attributed to the enhanced

conjugated electronic and ordered structure in the CMS membranes. Moreover, the 900 oC

pyrolyzed Trip(Me2)-TB membrane demonstrated a remarkable C2H4/C2H6 selectivity of up to 57

under long-term testing conditions with a high-pressure mixed gas environment of 20 bar.

Numerous MOFs were incorporated with highly aromatic or aliphatic components, exhibiting

multiple weak interactions with C2H6, but are favorable in ethane-preferential adsorption [300-

302]. Despite this, there is a scarcity of research focusing on crystalline MOF membrane-based

C2H6/C2H4 separation. Sun and colleagues introduced a fluorinated MOF-based MMM, marking

the inaugural exploration of crystalline MOF-based MMMs in ethylene separation [303].

Fluorinated carboxylic acids (FCA) with different C-F chain lengths were incorporated into the

MOF-808 framework to augment the C2H6 affinity. This integration facilitated the creation of a

selective transport channel tailored for C2H6, while concurrently reducing the pore size, thereby

amplifying its molecular sieving capability. The HFBA-MOF/PIM-1 membrane derived from the

heptafluorobutyric acid (HFBA), exhibited optimal separation performance at 253 K and 6 bar.

This achievement was primarily attributed to the increase in the length of the FCA chain which

increased the concentration of ‒F sites within the pores, thereby promoting the diffusion and

transport of C2H6. Additionally, further elongation of the chain length led to a reduction in the size

of the gas channel, which could impede diffusion. Furthermore, lowering the temperature or

increasing the pressure enhanced the interaction between C2H6 and the pore channels.

94
Table 7. Ladder PIMs-based membranes for ethylene/ethane separation.

Thermal Permeability
Thermal Selectivity
treatment T/P (Barrer)a
Membrane ID treatment
temperature (°C/atm) Ref.
method C2H4 C2H6 C2H4/C2H6
(°C)
PIM-6FDA-OH-250C Pyrolysis 250 35/2 5.5 1.4 4.0 [288]
PIM-6FDA-OH-500C 500 35/2 276 95.2 2.9 [288]
PIM-6FDA-OH-600C 600 35/2 66.1 10.7 6.2 [288]
PIM-6FDA-OH-800C 800 35/2 10.4 0.6 17.5 [288]
800 35/4 10.6 0.8 14.0 [288]
800 35/8 10.7 0.9 12.5 [288]
800 35/12 10.5 1.0 11.0 [288]
800 35/20 9.6 1.2 8.3 [288]
PIM-1-250C Pyrolysis 250 35/2 1640 906 1.8 [296]
PIM-1-400C 400 35/2 1082 439 2.4 [296]
PIM-1-600C 600 35/2 44 7 6.3 [296]
PIM-1-800C 800 35/2 1.3 0.1 13.0 [296]
PIM-6FDA-250C TR- 250 35/2 7.9 2.2 3.6 [289]
PIM-6FDA-CMS500C Pyrolysis 500 35/2 328 156 2.1 [289]
PIM-6FDA-CMS600C 600 35/2 77 18.8 4.1 [289]
PIM-6FDA-CMS800C 800 35/2 3.0 0.1 25.0 [289]
PIM-6FDA-CMS800C 800 35/4 3.0 0.2 17.9 [289]
800 35/8 3.1 0.2 17.2 [289]
800 35/12 3.1 0.2 16.7 [289]
800 35/20 3.2 0.2 15.6 [289]
CB700-cPIM Pyrolysis 700 35/2 11.8 1.7 7.0 [298]
CB700-cPIMb 700 35/2 7.5 0.8 9.3 [298]
CB700-borax 700 35/2 13.5 1.8 7.7 [298]
CB700-boraxb 700 35/2 9.1 1.0 9.4 [298]
CB700-BDA 700 35/2 14.1 1.4 10.3 [298]
CB700-BDAb 700 35/2 9.2 1.0 9.3 [298]
CB700-DHFDA 700 35/2 17.9 1.4 13.1 [298]
CB700-DHFDAb 700 35/2 13.7 1.4 9.7 [298]
Trip(Me2)-TB-120 Pyrolysis 120 35/2 401 207 1.9 [299]
Trip(Me2)-TB-600 600 35/2 160 32 5 [299]
Trip(Me2)-TB-800 800 35/2 14 0.9 17 [299]
Trip(Me2)-TB-850 850 35/10 4.6 0.1 47 [299]
Trip(Me2)-TB-900 900 35/10 2 0.02 96 [299]
EA(Me2)-TB-120 120 35/2 75 29.5 3 [299]
EA(Me2)-TB-600 600 35/2 12.9 1.4 9 [299]
EA(Me2)-TB-800 800 35/10 0.4 0.005 77 [299]
Trip(Me2)-TB-850b 850 35/10 2.7 0.1 33 [299]
Trip(Me2)-TB-850b 850 35/15 2.4 0.1 33 [299]
Trip(Me2)-TB-850b 850 35/20 2.1 0.1 33 [299]
Trip(Me2)-TB-900b 900 35/10 1.6 0.02 100 [299]
Trip(Me2)-TB-900b 900 35/15 1.3 0.02 72 [299]
Trip(Me2)-TB-900b 900 35/20 1.1 0.02 57 [299]

95
HFBA-MOF/PIM-1 - - 1.2/30 7569 5526 0.7 [303]
HFBA-MOF/PIM-1 2/30 6244 5433 0.9 [303]
HFBA-MOF/PIM-1 4/30 4975 5224 1.1 [303]
HFBA-MOF/PIM-1 6/30 3070 5005 1.6 [303]
HFBA-MOF/PIM-1 6/-20 1830 4246 2.3 [303]
HFBA-MOF/PIM-1 6/-10 2130 4538 2.1 [303]
HFBA-MOF/PIM-1 6/0 2472 4797 1.9 [303]
HFBA-MOF/PIM-1 6/50 5233 5547 1.1 [303]
a 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. c Mixed gas composition C2H4:C2H6 50:50

4.2. Propylene/propane separation

Propylene is the second most commonly used olefin in the world after ethylene. It is an attractive

raw material for the production of propylene derivatives such as polypropylene and propylene

oxide [304]. Propane, also known as liquid petroleum gas (LPG), is also a fuel in industry and

households. Various membranes have been developed and discussed on the separation of

propylene/propane, including PIM and its derived ion-doped membranes, CMS membranes, TR

polymer membranes, and MMMs (Table 8).

Liao et al. proposed two methods to modify PIMs membranes, e.g., hydrolysis and doping with

metal ions (Zn2+, Mg2+, and Ag+) for improving C3H6/C3H8 separation [297]. The hydrolyzed PIM-

1 (e.g., cPIM-1) has the smallest free volume, which was responsible for the high C3H6/C3H8

selectivity (Figure 15b). It was found that the membrane modified by metal ions has higher

selectivity than cPIM, which is related to the characteristics of metal-carbonyl complex. Notably,

PIMs membrane modified with Zn2+ ions demonstrated better plasticization resistance, achieving

C3H6 permeability of 837 Barrer. While for the case of Mg2+ ions doping, the presence of Mg2+

enhanced the affinity between the membrane to π-orbital of propane molecules, creating an

enhancement of four-fold in C3H6/C3H8 selectivity.

To solve the low gas selectivity in PIM membranes due to inefficient chain stacking, Ren et al.

proposed a coordination-driven reconstruction (CDR) strategy by adding metal ions to AO-PIM

96
for in situ generate coordination cross-linked networks of size-sieving capability (Figure 15c)

[305]. Metal ions, especially alkali metals (e.g., Na+ and K+) or alkaline earth metals (e.g., Ca2+),

possess the ability to form weak coordination bonds with amidoxime, subsequently enlarging the

interchain spacing of AO-PIM-1 for effective separation of C3H6. Among AO-PIM-Na, AO-PIM-

K and AO-PIM-CA membranes, the AO-PIM-K membrane demonstrated promising C3H6

permeability of 147 Barrer along with C3H6/C3H8 selectivity of 50. In comparison, the introduction

of transition metals (e.g., Fe, Zn, Ag) created strong coordination bonds with amidoxime groups

thus compromising their separation performance. Particularly, AO-PIM-Ag membrane showed a

high selectivity of 31, which was attributed to the presence of Ag+ that promotes facilitated

transport of C3H6 as well as the smaller pore size of AO-PIM-Ag membranes that inhibited the

transport of C3H8 molecules. The coordinated cross-linking structure also significantly improved

the stability and plasticization resistance of the membrane under high pressure.

Likewise, CMS membranes have been an intense research area in developing membranes for

propylene/propane separation. Inspired by the previous work [234], for the first time, Liu et al.

thermally treated PIM-CD to covalently cross-linked polymer networks with large gas transport

free volume and tighter structure for the separation of C3H6/C3H8 [306]. Heat treatment

decomposed the thermally unstable molecule β-CD, releasing C2/C3 compounds with gases of

similar size to C3H6 and C3H8, accelerated the self-crosslinking reaction of nitrile, and

subsequently forming a 3D cross-linking network conducive to the separation of C3H6/C3H8. They

also found that an increase in the thermal treatment temperature and CD loading of the PIM-CD

CMS membranes improved the selectivity.

TR and CMS membranes are both prepared via carbonization of polymer precursors at elevated

temperatures. The former is produced at an oxygen-containing furnace of lower temperature (e.g.,

97
400 °C) while the latter is at a higher temperature (e.g., 600 °C) under oxygen-free conditions [67].

TR membranes generally possess excellent thermal, chemical, and plasticization resistance.

However, there is only limited progress reported for PIMs based TR membranes on C3H6/C3H8

separation, with precursors focused mostly on PIM-PIs. Swaidan et al. reported the synthesis of

CMS membrane and TR membrane from precursor PIM-6FDA-OH [307]. The high C3H6/C3H8

selectivity in the CMS membranes was attributed to the larger fraction of ultramicropores in CMS

membranes (Figure 15d). Yerzhankyzy et al. fabricated PBO membrane prepared from

hydroxyfunctionalized PIM-PI precursor, 6FDA-DAT1-OH (Figure 15e) [308]. The resultant

membrane showed an improvement in mechanical properties. It is worth noting that the

permeability of all gases, especially C3H8 achieved 190-fold higher than the 6FDA-DAT1-OH

membrane after thermal treatment which is due to the significant increment in membrane porosity.

Shen et al. prepared high-performance MMMs from SIFSIX-3-Zn nanoparticles and PIM-1 in

which the fluorinated MOF nanoparticles created new transport channels and facilitated the

transport of propylene [309]. Figure 16a depicts the transportation characteristics of C3H6 and

C3H8 in membrane system. As the loading of SIFSIX-3-Zn increased to 10%, the C3H6

permeability of the resultant membrane increased to 2300 Barrer, while the C3H6/C3H8 selectivity

enhanced nearly one-fold. This is inseparable from the fact that SIFSIX-3-Zn nanoparticles acted

as chain stiffener in providing additional channels for the gas transport. Peng et al. developed a

new ZIF-8 composite material functionalized by silver complex (e.g., Ag3pz3) [310]. The

incorporation of the functionalized ZIF-8 into PIM-1 membrane promotes gas transport through π

complexation. The PIM-1/ZIF-8@Ag3pz3-1 membrane with 10 wt% fillers achieved highest C3H6

permeability with 84% higher selectivity than that of PIM-1/ZIF-8 membrane. Moreover, the

membrane demonstrated good aging resistance and long-term stability towards the separation of

98
C3H6/C3H8. Similarly, Cong et al. used Ag3pz3 fillers to prepare PIM-1/Ag3pz3 MMMs. It was

found that the π-acids of Ag+ in Ag3pz3 were the key to favor C3H6/C3H8 separation, while Ag3pz3

complexes were able to overcome the long-term stability of Ag+ during the separation process

[311]. Figure 16b shows the synthesis route of Ag3pz3 complexes. Similarly, PIM-1/Ag3pz3-10 wt%

has the highest C3H6/C3H8 selectivity of ~200. Further research on the effects of pressure and

temperature revealed that an increase in both parameters reduced the C3H6 permeability

significantly. The former was due to the occurrence of plasticization in the membrane, while the

latter was attributed to the acid-base complexation is an exothermic process. Ma et al. reported

PIM-6FDA-OH/ZIF-8 MMMs with excellent C3H6 plasticization resistance up to 7 bar [312].

Good miscibility between ZIF-8 and PIM-6FDA-OH was observed, with a good distribution of

ZIF-8 in the membrane matrix at 52 wt% (Figure 16c). Although agglomeration was visible at

ZIF-8 loading of 65 wt%, there were no interfacial voids observed in the TEM images. The PIM-

6FDA-OH/ZIF-8-65 wt% maintained excellent separation performance in both pure- and mixed

gas tests, and the C3H6/C3H8 selectivity was improved by 43%.

Wang et al. proposed a covalent-linking strategy to reinforce the interfacial compatibility of ZIF-

8/PIM-1 membranes [313]. The mixed-ligands ZIF-8-CN was covalently linked to PIM-1 via the

cyano groups of both entitles by thermal reaction at 300 °C, which strengthened the polymer-filler

connection of resultant membranes. The C3H6/C3H8 selectivity of tPIM-1@ZIF-886-CN14 was 27.6,

significantly higher than that of untreated PIM-1/ZIF-886-CN14 and non-functionalized tPIM-

1@ZIF-8. This may be attributed to the thermal treatment which created a covalent linkage of

PIM-1 polymer chain with the surfaces of ZIF-886-CN14 crystal as well as the elimination of non-

selective voids for effective transportation of C3H6. Figure 16d illustrates the proposed permeation

99
path of C3H6 and C3H8, whereby the separation was dominated by pore diffusion as evidenced by

the sorption analyses.

Table 8. Ladder PIMs-based membranes for propylene/propane separation.

Permeability
T/P Selectivity
Membrane ID Modification method (Barrer)a Ref.
(°C/atm)
C3H6 C3H8 C3H6/C3H8
PIM-1 - 35/3.5 2534 851 3.0 [297]
cPIM Hydrolysis 35/3.5 421 44.7 9.4 [297]
ZncPIM Hydrolysis + Zn2+ ion 35/3.5 837 80.9 10.4 [297]
treated
AgcPIM Hydrolysis + Ag+ ion 35/3.5 759 64.9 11.7 [297]
treated
MgcPIM Hydrolysis + Mg2+ ion 35/3.5 568 37.3 15.2 [297]
treated
PIM-1 - 25/1.5 1298 209.4 6.2 [305]
PIM-1b 50/1.1 1100 157.1 7.0 [305]
AO-PIM Amidoxime- 25/1.5 55 2.4 23.0 [305]
AO-PIMb functionalized 50/1.1 55 6.0 9.1 [305]
AO-PIM-K Amidoxime- 25/1.5 147 2.9 50.0 [305]
AO-PIM-Kb functionalized + K+ion 50/1.1 170 12.1 14.0 [305]
treated
AO-PIM-Ag Amidoxime- 25/1.5 40 1.3 31.8 [305]
AO-PIM-Agb functionalized + Ag+ 50/1.1 52 2.5 21.0 [305]
ion treated
Pristine PIM Thermal annealing 35/3.5 2861 1022 2.8 [306]
PIM-600 35/3.5 27 4 6.8 [306]
PIM-CD-1% 35/3.5 4315 1725 2.5 [306]
PIM-CD-2% 35/3.5 5032 2097 2.4 [306]
PIM-CD-2%-600 35/3.5 42 5 8.4 [306]
TR-PBO (PIM-6FDA-OH) Thermal rearrangement 35/2 14 0.9 15.0 [307]
CMS (PIM-6FDA-OH) Pyrolysis 35/2 45 1.4 33.0 [307]
TRIP-TR-460-30 (6FDA-DAT1-OH) Thermal rearrangement 35/2 39 1.7 23.0 [308]
PIM-1/SIFSIX-3-Zn-10 wt% MMMs 35/1 4000 500 7.9 [309]
PIM-1/ZIF-8-10 wt% MMMs 30/1 3300 600 6.4 [310]
PIM-1/ZIF-8@Ag3pz3-1-10 wt% 30/1 3767 280 11.7 [310]
PIM-1/Ag3pz3-10 wt% MMMs 30/1 306 1.6 194.0 [311]
PIM-6FDA-OH/ZIF-8-65 wt% MMMs 35/2 38 0.9 43.0 [312]
PIM-6FDA-OH/ZIF-8-65 wt%b 35/2 35 1.1 31.2 [312]
tPIM-1@ZIF-8-CN MMMs + thermal 45/1.5 369 13.4 27.6 [313]
treatment @ 300 °C
a 1 Barrer = 1 x 10-10 cm3(STP)cm/cm2scmHg. b Mixed gas composition C3H6:C3H8 50:50

100
a)

b) c)

d) e)

6FDA-DAT1-OH

TR-6FDA-DAT1-OH

Figure 15. (a) Potential alterations in pore structure for CMS membranes derived from

Trip(Me2)-TB and EA(Me2)-TB precursors [299]. (b) Free volume distribution in PIM-1, cPIM,

and cPIM membranes treated with metal ions [297]. (c) Schematic of the coordination driven

structure reconstruction of PIM-1 membrane [305]. (d) Micropores formation and sintering of

ultramicropores in transition of TR-PBO (440 °C) to CMS (600 °C) [307]. (e) Thermal

rearrangement process from 6FDA-DAT1-OH polyimide to TR-6FDA-DAT1-OH [308].

Copyright 2016, 2020, 2022 and 2023. Reproduced with permission from Cell Press, Elsevier

and Nature.

101
a) b)

c) d)
d)

Figure 16. (a) C3H6 and C3H8 in SIFSIX-3-Zn/PIM-1MMM exhibit different transport

characteristics [309]. (b) Synthesis route of Ag3pz3 complex [311]. (c) TEM images of PIM-

6FDA-OH based MMM containing 13, 33, 41, 52, and 65 wt% ZIF-8 [312]. (d) The permeation

mechanism of C3H6 and C3H8 through ZIF-886-CN14@tPIM-1 membranes [313]. Copyright

2018, 2019, 2022 and 2023. Reproduced with permission from American Chemical Society,

Elsevier and Wiley-VCH.

4.3. Summary on the development of PIMs membranes in light hydrocarbon separation

Light hydrocarbon separation is indispensable across diverse industrial sectors. Compared to

conventional separation techniques, a growing interest in exploring membrane separation

technologies has been witnessed owing to the cost- and efficiency of membrane separation. Among

the developed membrane materials, PIMs stand out due to the large FFV of finer pores, making

them promising for separating gases with minute-size differentials. However, the wide PSD of

PIM-1 leads to sub-optimal separation efficacy. Consequently, the focal point has shifted towards

102
investigating PIMs-derived membranes or modifying existing PIMs membranes to enhance

separation performance.

The heat treatment of membranes based on PIMs demonstrates remarkable stability and excellent

separation efficiency, attributed to the densification of micropores and carbonization

characteristics resulting from the pyrolysis of precursor structures. This can be achieved through

precise control of pyrolysis temperature, the utilization of functional polymers or the incorporation

of metal ions/hydrolysis for modifying the PIM membranes. Furthermore, MOFs-based MMMs

offer innovative avenues for enhancing selectivity and permeability, particularly through covalent

grafting strategies that address the poor interfacial compatibility issues of MMMs. Notably, the

CDR strategy holds promise in ameliorating inefficient chain stacking, thereby enhancing stability

and separation performance even in demanding environments. Collectively, these investigations,

spanning from polymer material selection to membrane fabrication, present diverse strategies for

overcoming challenges in light hydrocarbon separation. Such advancements significantly

contribute to the progression of industrial gas separation applications.

5. PIMs in other advanced industrial applications

The macromolecular structure of PIM provides a unique performance combination for this organic

material, including interconnected microporous networks, good solubility, high free volume, and

structural diversity [28, 314]. This led to its wide application in environmental remediation (e.g.,

water treatment, oil clean-up, air filtration), catalysis, electrochemical energy storage and

conversion, sensing, as well as 3D printing. PIM can be produced in the form of powder, films,

membranes, and nanofibrous depending on the practical applications.

103
5.1. Environmental applications

Water pollution has become a global concern, with harmful substances such as heavy metals,

aromatic compounds and dyes constantly accumulating, threatening the environment. Various

wastewater treatment methods such as adsorption [315], oxidation [316, 317],

coagulation/flocculation [318], and membrane separation [319] have been studied. Adsorption is

one of the commonly applied methods in wastewater treatment, owing to its simplicity and

efficiency. Guo et al. reported an outstanding PIM adsorbent for the adsorption of phenolic

pollutants (e.g., phenols, and phenolic derivatives) in water, with high pH selectivity and low

pollutant residue ability over different phenols [320]. The PIM-1 adsorbent achieved a high phenol

selectivity of 22.4 over p-nitrophenol (PNP) and phenol removal efficiency of 99.2% for a

phenol/PNP mixture, meeting the safety requirement of EPA for drinking water. Notably, the

recover efficiency of PIM-1 adsorbent maintained above 95% after five cycles of regeneration by

dissolve-separation-precipitation (DSP) strategy.

Moreover, PIMs as adsorbents were applied for the adsorption of ions from wastewater. A borane

dimethyl sulfide complex modified PIM-1 fibrous membranes (AM-PIM-FM) was found to

demonstrate a high adsorption capacity for anionic dyes (e.g., methyl orange), with an adsorption

capacity of 312.5 mg/g according to the Langmuir isotherm model [321]. The same group of

researchers further utilized AO-PIM-1 fibrous membranes for the extraction of uranyl (VI) ions

from aqueous systems via column sorption [322]. The resultant membranes achieved a 20 times

increment in uranium adsorption efficiency (e.g., 590 μg/g) compared to the non-functionalized

PIM-1. Additionally, the membrane form AO-PIM-1 demonstrated higher feasibility in the column

adsorption process compared to the powder form AO-PIM-1, despite the powder form showing

higher adsorption performance in the batch adsorption process. In a following study, AO-PIM-1

104
powder was explored in the removal of ionic dyes (e.g., methylene blue and methyl orange) from

an aqueous system [323]. The AO-PIM-1 was found to simultaneously remove anionic and

cationic dye mixtures by changing the pH of the solution, in which low pH (e.g., pH 3) favored

the adsorption of anionic dye, and vice versa. The methyl orange and methylene blue adsorption

capacities of this material were reported at 86.7 mg g-1 and 81.3 mg g-1, respectively at pH 6 and

25 oC.

Furthermore, PIM-1-based membranes were used in liquid or water treatment such as desalination

[324], and pervaporation [124, 325, 326]. Self-standing PIM-2 fibrous membranes were utilized

in clean-up of oil spills, and benefited from the considerable amount of fluorine in the polymer

backbone which provides the polymer with significant hydrophobicity [327]. Notably, PIMs

materials are attractive in nanofiltration (NF), especially organic solvent nanofiltration (OSN)

owing to the nature of porosity delivered by the rigid polymer backbone and excellent solution

processability [328, 329]. However, the direct application of free-standing PIMs membrane in NF

or OSN is limited, particularly attributed to the ease of overly swelling of polymer chains in polar

solvent (e.g., methanol) during fabrication or separation which consequently impairs its filtration

properties [46, 330, 331]. Recently, PIMs were successfully engineered to prepare TFC

membranes by introducing rigid monomers during the fabrication of selective layers for NF and

OSN.

Seeking into NF application, Agarwal et al. adopted a layer-by-layer technique, coated the

oppositely charged Troger’s base polymer such as hydroxy functionalized partially quaternized

Troger’s base PIM and sulfonic acid functionalized Troger’s base PIM alternately on a substrate

to prepare TFC NF membranes with less defectivity and improved rejection performance [332].

While this technique is viable for the preparation of TFC membrane with improving separation

105
performance, however, a long duration for multilayer fabrication was required. In subsequent work,

they reported a time-efficient and low-defectivity method to prepare a single layer TFC membrane

with hydroxy functional branched Troger’s base PIMs [333]. The membrane stability was further

improved by chemical cross-linking with a tri-functional epoxy cross-linker. Additionally, the

introduction of PIMs monomer into the aqueous phase of interfacial polymerization is vital to

enhance the microporosity and improve the microstructure of the selective layer of TFC NF

membranes. Jiang et al. utilized rigidly-contorted bisphenols (e.g., two phenols containing

spirobisindane and Trӧger’s base structure, respectively) and piperazine as co-monomer in the

aqueous phase for interfacial polymerization with trimesoyl chloride (TMC) in the organic phase

to prepare the TFC selective layer [334]. The resultant NF membranes exhibited a 2.7-fold pure

water flux enhancement and Na2SO4 rejection of 99.2%. Likewise, Chen et al. prepared NF

membrane with “corolla-like” structure, high porosity and a thin selective layer of <55 nm from

Troger’s base 2,8-diamino-4,10-dimethyl-6H,12H-5,11-methanodibenzo [1,5]-diazocine (TBDA)

diamine and piperazine co-monomer [335]. The NF membrane showed excellent water flux of

87.6 LMH (14.6 LMH/bar), high Na2SO4 rejection of 99.3%, superior anion selectivity as well as

high acid resistance. Meanwhile, integration of rigidly-contorted TTSBI as co-monomer was also

found to endow thinner selective layers with undulating structures which influenced the monomer

diffusion behaviours at the interface during the interfacial polymerization process, subsequently

enhanced the pure water flux up to 9.9 LMH/bar [336]. To tailor the stacking density of as-form

polyarylate (PAT) network, Dou et al. investigated the full substitution of TTSBI via interfacial

polymerization with three types of acyl chloride monomers, namely, rigid TMC, short-chain

glutaryl chloride and long-chain sebacoyl dichloride (SDC) in replacing piperazine for fabricating

PAT NF membranes [337]. The developed membranes exhibited high permeability and tunable

106
size-selectivity, with TTSBI-TMC membranes demonstrating the highest water flux of about 480.5

LMH and TTSBI-SDC membranes showing the highest dye rejection (e.g., Congo Red

rejection >95%).

The development of OSN membranes poses a greater challenge compared to that of NF

membranes due to the presence of a large amount of organic solvent in the treatment process.

Fritsch and Livingston’s groups reported a series of PIM-1 based TFC membranes with superior

performance than the commercial StarmemTM 240 OSN membranes, especially for alkane

solvents treatment [330, 331]. Unfortunately, the PIM-1 selective layer was easily swelled and

detached in polar solvents, which compromised the lifespan of the membrane and severely

disrupted its separation performance. Thermal or chemical cross-linking techniques were proposed

to address this issue. Gao et al. spin-coated a layer of thioamide-functionalized PIM-1 on a cross-

linked P84 substrate which demonstrated excellent rejection of dyes and polar solvents (e.g.,

acetone and ethanol) [338]. Further cross-linked the membrane with TMC significantly tightened

the pore size and free volume of the selective layer, leading to better chemical stability and

rejection but with a reduction in solvent permeance. Similarly, improved stability and organic

solvent separation performance were noticed on TFC membranes prepared by interfacial cross-

linked the acyl chloride-functionalized PIM (PIM-COCl) with amine-functionalized PAN

substrate [339]. Tian et al. reported an enhanced separation selectivity and organic solvent

resistance in PIM-1 TFC OSN membranes prepared via a simple spin coating and thermal

treatment method, whereby better chemical stability was contributed by the formation of stable

triazine rings on the selective layer by the nitrile group of PIM-1 after thermal treatment [340].

Besides, the solvent vapor annealing process was also introduced to effectively control the solvent

107
evaporation rate and eliminated the generation of small pits on PIM-1 selective layer, thereby

enhancing the interfacial bonding and stability of the resultant PIM-1 TFC OSN membranes [341].

Enhancement of organic solvent resistance and separation selectivity often leads to a trade-off in

solvent permeance. Numerous attempts have been reported to simultaneously optimize the

filtration performance and physiochemical stability of PIMs OSN membranes, for instance by

solvent activation [342] or incorporation of nanomaterials [343-345]. Jin et al. reported on a spin

coating coupled with a solvent activation method to prepare AOPIM-1 TFC membranes [342].

The fabricated membrane showed excellent ethanol permeance of 15.5 LMH/bar with a good

rejection of dye (e.g., MWCO > 800 Da) and vitamin B12 in ethanol was achieved after ethanol

activation, particularly attributed to the enhanced free volume of AOPIM-1 layer that reduced

solvent transport resistance and promoted the solvent permeance. Meanwhile, the presence of

porous nanomaterials including metal oxides (e.g., ZnOx, TiOx, AlOx) [343], COFs [344] and

HCPs [345] were found to restrain the polymer chain mobility and offer extra pathways for the

transportation of molecular solvent across the membranes, leading to an improvement in solvent

permeance. The reported thin film nanocomposite (TFN) membranes were prepared via either

vapor phase infiltration, interfacial polymerization or surface coating for efficient OSN application.

A number of interesting works were reported on developing outstanding TFC membranes for the

treatment of organic solvents through nanofiltration. For example, co-monomers of TTSBI and m-

phenylenediamine were used to prepare the selective layer on TFC via interfacial polymerization,

in which the resultant membranes demonstrated high permeance on polar solvents and high

rhodamine B rejection of 99%, as well as maintained superior solvent resistance after immersion

in DMF at 80 °C for 63 days [346]. Abdulhamid et al. presented a simple approach to prepare OSN

membranes with honeycomb-like surface structure by regulating the molecular weight of the

108
polymer [347]. High and low molecular weights of spirobisindane-based intrinsically microporous

poly(ether-ether-ketone) (iPEEK-SBI) homopolymers were synthesized. The low molecular

weight polymers produced OSN membranes with flexible and honeycomb-like structures of

various honeycomb cell sizes, while the high molecular weight polymer yielded membranes with

flat surfaces. The honeycomb-like membranes displayed significantly higher solvent permeance

(e.g., 18-26%) and stable performance over seven days of cross-flow filtration and a six-month

aging period. Additionally, a polymer blend of PIM-1 and Matrimid was used in fabricating the

selective layer for TFC OSN membranes [348]. The presence of Matrimid endowed a more

compact chain packing and narrower pore size distribution on the selective layer compared to the

pristine PIM-1 layer. Other than PIM-1, two variants of PIMs such as PIM-7 and PIM-8, for the

first time, have been utilized to prepare OSN membranes via roll-to-roll dip coating [46]. It was

revealed that the composite membranes coated on cross-linked Ultem 1000 substrate possessed

higher stability. Compared to PIM-1 and PIM-7, the PIM-8 membranes demonstrated the highest

separation factor between n-C16 and i-C16 alkanes, suggesting the development of sub-nanometer

pores. Besides, the molecular interactions between PIM-1 membrane with toluene, methanol and

their mixtures in OSN were studied using Molecular Dynamics (MD) simulation [349]. Besides,

there are also a number of studies reported on TFC OSN hollow fiber membranes [343, 350, 351].

Antibiotics can be used in animal farming and aquaculture activities to prevent disease, treat

diseases caused by bacterial infections, and promote growth [352]. Nevertheless, the extensive use

of antibiotics has raised environmental pollution, as well as potentially increased the antibiotic

resistance of microorganisms. Traditional technologies such as biological treatment and membrane

filtration, despite being commercialized, could only partially remove the traces of antibiotics

present in wastewater. In a recent study, PIMs used for antibiotic adsorbents have been studied

109
[353]. In this study, four commonly used antibiotics, e.g., doxycycline, ciprofloxacin, penicillin

G, and amoxicillin were evaluated, whereby, the adsorbent was able to remove nearly 80% of

antibiotics from aqueous solutions in a short duration. The Freundlich model fitted the isothermal

data well, with the PIM-1 adsorbent demonstrating highest and lowest adsorption capacity on

penicillin G (275 mg/g) and ciprofloxacin (33.10 mg/g), respectively.

Particulate matters (PMs) are one of the major air pollutants and are in liquid or solid forms with

different aerodynamic diameters, such as nitrates, sulfates, mineral dust, and water [354]. Air

filtration has been proposed to remove this particle pollution for environmentally benign. Highly

selective, lightweight, and breathable protective clothing is requisite for first responders or the

military who work in an environment of high particulate concentrations over a long period. PIM-

1 was proposed to be used as the matrix material for the fabrication of aerosol-protective layered

fabric. Particularly, the electrospun PIM- 1 fiber web acted as the barrier to toxic gas penetration

while providing pathways for air vapor molecules [355]. Utilizing layer-by-layer electrospun-

deposition, PAN nanofibers were integrated with PIM-1 to obtain PIM/PAN materials with high

tensile strength and PM2.5 removal efficiency of up to 99.8%. Further addition of MOF particles

(e.g, UiO-66-NH2) to the PIM/PAN/MOF composite fiber mat enhanced the sorption capacity and

endowed remarkable catalytic degradation properties towards chemical warfare agents stimulant

(e.g., dimethyl 4-nitrophenyl phosphate (DMNP)) (Figure 17a).

Hydrogen sulfide (H2S) is corrosive, explosive (4.3-45% by volume in air), flammable (ignition

temperature 260 °C), and highly toxic gas [356]. Natural gas comprising of high H2S concentration

(e.g., > 4 ppm by volume under STP) is considered to be sour gas, in which the natural gas has to

be sweetened to meet the specification for transport and processing using a membrane system or

hybrid membrane-amine sorption system. In addition to the promising performance in CO2 capture,

110
the AO-PIM-1 membrane was also reported to demonstrate unprecedented separation performance

in sour gas purification [357]. Interestingly, the AO-PIM-1 membrane exhibited ultrahigh H2S

permeability >4000 Barrer with extraordinary H2S/CH4 selectivity of 75 under a ternary feed

mixture (H2S:CO2:CH4 = 20:20:60 vol%) under a challenging feed pressures up to 77 bar,

demonstrating performance of two to three orders of magnitude higher than commercially

available glassy polymeric membranes. The promising performance was ascribed to the strong

H2S interaction with the polymer backbone and enhanced solubility selectivity of the membranes.

5.2. Catalysis

Catalysts are used in a reaction or catalysis to increase the rate of reaction. Metallic catalyst is

mostly applied in catalysis owing to their high surface area and promising performance [27], while

non-metallic catalysts are desirable due to their low cost and ecological viability [358]. With the

continuous development of industrial processes, cheap, effective and environmentally friendly

metal or non-metallic catalyst alternatives are desirable. In this regard, knowing that the ability to

induce the formation of C-C bonds and condensation reaction are two important criteria for a

catalysis reaction, high porosity, functionality, high surface area and stability of PIMs made them

attractive candidates as efficient catalytic materials. Thus far, PIMs have been reported for

catalysis applications such as CO2 conversion and utilization [359-362], oxidation [363], reduction

of nitro compounds [364], photocatalysis [365-367], and electrocatalysis [47, 368-372].

Interestingly, Jung et al. anchored Keggin-type polyoxometalates (POMs), namely H3PW12O40

with tunable acidity and rich redox properties with AO-PIM-1 as matrix, forming PIM-POM

composite materials for the catalytic oxidation of a sulfur mustard simulant, 2-chloroethyl ethyl

sulfide (CEES) (Figure 17b) [373]. It was found that the selectivity of PW12@PIM-1-AO for single

oxide sulfoxide achieved 86% while maintaining the majority of their catalytic activity over

111
repeated cycles. The anchoring of PIMs to POMs (or other homogeneous catalysts) forming

heterogeneous catalysts compensated the challenges of aggregation and instability of POMs in

solvents, as well as favored the recovery and recycling of the catalyst as compared to homogeneous

catalysts.

5.3. Electrochemical energy storage and conversion

Being a versatile class of molecular polymer materials, PIMs have gained great interest in

electrochemical energy storage and conversion recently. Delving into the application in

electrochemical energy storage, PIMs were found to be able to suppress the dendrite growth

phenomenon which frequently happened on lithium anodes caused by lithium ions that approached

the electrolyte solution to the surface of anodes, favoring the growth of sharp surface features. For

instance, PIM-1 was introduced as an artificial solid-electrolyte-interphase (SEI) for batteries to

protect the lithium metal anode against dendrite growth by preventing the direct contact of lithium

anode with solvent molecules [374]. Likewise, PIM-DMBP-TB (DMBP: dimethylbiphenyl) with

high Young’s modulus and mechanically flexible was introduced as an interfacial layer to suppress

lithium dendritic growth (Figure 17c) [375]. PIM-1 was also employed to maintain uniform ion

flux in the lithium plating/stripping process with high Columbic efficiency and good cycling

performance [376, 377].

Separators that are placed between cathode and anode in batteries are critical in determining rate

capacity, charge-discharge rate, and cycling rate. Separators should possess characteristics such as

high mechanical rigidity, chemical and thermal stability, electric-insulated, wettability and ion

conductivity [26]. PIMs such as PIM-1 and PIM-7, have been proposed as separators in lithium-

ion batteries [378], lithium-sulfur batteries [379-381], Li/Na-H2O2 batteries [382], and sodium-

sulfur batteries [383], owing to the good Li+ transportation properties contributed by polar ether

112
bonds and cyano groups on the polymer backbone of PIMs. Nevertheless, PIM-1 and PIM-7 were

found to suppress the polysulfide “shuttle effect” in lithium-sulfur batteries, being a drawback that

degrades the efficiency of the batteries.

Supercapacitors, either electrochemical double-layer capacitors (EDLCs) or pseudocapacitors, are

energy storage devices that store and deliver energy at high charge-discharge rates. Porous carbon-

based materials can be one of the electrode materials, owing to the high surface area which favor

capacitor performance [384, 385]. For instance, carbonised PIM-1 material was found to have a

BET surface area of 546 m2/g, while further steam treatment increased the surface area to 1162

m2/g, demonstrating capacitance of 120 F/g at 10 mV/s scan rate in 1,3-ethylmethylimidizaolium

bis(trifluoromethanesulfonyl) imide ionic liquid electrolyte [386].

Meanwhile, PIMs were also applied as membranes for redox flow cells and fuel cells in

electrochemical energy conversion systems. PIMs can be incorporated as a catalyst binder, a gas

diffusion electrode, as well as an ion-exchange membrane. It is necessitated for the membranes

applied in fuel cells to have high chemical robustness, high cation-anion mobility and high

retention of water at elevated temperature, whereas that of redox flow cells have to be highly

selective to prevent cross-mixing of two redox-active species across the chamber. A few studies

revealed that PIM-1 has considerably improved the efficiency of lithium-sulfur flow batteries

[387], vanadium redox flow batteries [388, 389] and non-aqueous all-organic redox flow batteries

[390, 391].

A high-efficient proton exchange membrane (PEM) is often evaluated based on the proton mobility

and ion exchange capacity (IEC) in fuel cells, according to the classic Nafion membrane as

performance benchmark (proton conductivity: 70 mS/cm; IEC: 0.9 meq/g [392]). For instance, a

PEM fabricated from PIM-1, PBI and phosphoric acid obtained a proton conductivity of 313

113
mS/cm at 200 °C [120]. Noteworthy, despite acid doping could induce excellent proton

conductivity and thermal stability, control of acid doping level and retention ability within the

membranes are also essential to produce desirable PEMs [393].

The development of anion exchange membranes (AEMs) is more challenging due to the chemical

conditions in the hydroxide environment. Hu et al. reported good dimensional stability and alkaline

resistance multi-cation cross-linked Tröger's base polymer AEM with monomers of 4,4′-

diaminodiphenylmethane and 4,4′-diamino-3,3′-dimethyl-biphenyl and cross-linker of

dibromohexane [394]. The membrane demonstrated a hydroxide conductivity of 103.9 mS/cm at

80 °C and IEC of 1.67 meq/g. Besides, Gong et al. reported a robust polymer-blended AEM of

PIM and polysulfones [395]. The membrane containing 30% PIM showed hydroxide conductivity

of 16.9 mS/cm at 30 °C, and yielded a H2/O2 fuel cell powder density of 163 mW/cm2 at 60 °C. In

a following study, AEM was prepared using fluoro-terminated polysulfone and hydroxyl-

terminated PIM, in which the hydroxide conductivity improved to 52.6 mS/cm at 80 °C [396].

5.4. Sensing materials

PIMs having excellent solution processability are promising candidates for coating or deposition

on electroanalytic sensors for both gas and liquid phase sensing applications. The presence of

microporous environment in PIMs not only favor the permeation of analyte molecules but also

enhances the adsorption affinity, achieving rapid response and low concentration detection. Rakow

et al. pioneered employing PIM-1 as a visual indicator of volatile organic compounds (VOCs) by

casting PIM-1 on Ni-coated polyethylene terephthalate substrate [397]. Later, they developed a

PIM-1 based chemosensor for the detection of nitroaromatic explosive vapors [398]. Other than

these, PIMs have also demonstrated their potential as actuators in drug delivery applications [399].

Recently, a PIM-1 fluorescent film sensor was developed to monitor the radioactive gas iodine (I2)

114
[400]. The fluorescence spectroscopy results demonstrated a high quenching efficiency of up to

99.99% and achieved a low detection limit of 1.24×10−9 mol/L. Notably, the quenching process

was found to be I2 selective and not much influenced by H2O, H2S, or CO2.

The application of PIMs materials in liquid phase sensors is limited by the slow permeation and

transport rate as a result of the microporous filling of solvent molecules. Rapid sensor responses

can be achieved through the fabrication of thin PIMs membranes for sensing aqueous media. PIMs

have been reported to display good responses in ionic diode membrane sensors [401-403],

cholesterol detection [404], glucose and carbohydrate detection [405, 406]. In addition, PIM-1

nanoparticles coupled with tin-doped indium oxide were reported to produce

electrochemiluminescence sensors for potential reaction oxygen species detectors [407]. Likewise,

Afshary et al. prepared the same type of sensor with nitrogen-doped carbon quantum dots for the

selective determination of citrate ions as a prostate cancer biomarker [408]. It is worth noting that

the sensor was also found to be useful for the detection of citrate levels in human serum and urine.

5.5. 3D printing

Additive manufacturing techniques, for example, 3D printing are attractive in the medical [409,

410], electrochemical [411], environmental [412], food [413], and fashion industries [414] due to

their ability to create complex structures. However, the choice of polymers remains limited despite

the remarkable progress of additive manufacturing techniques thus far. Many of the commercially

available polymers, such as polyimides and microporous ladder polymers of significant

technological interest, are incompatible with stereolithography and fused deposition modelling

[415], which the former requires polymers that can be photopolymerized or photo-crosslinked,

while the latter requires polymers with moderate melting points. These polymers are currently

processed into functional materials such as flexible electronic devices [416] and membranes [417]

115
using solution-based approaches [418-420]. Zhang et al. introduced a solution-based direct-write

3D printing technique using a ternary ink formulation to 3D-print a high-efficiency mass transfer

contactor (Figure 17d) [421]. To eliminate shrinkage, prevent filament deformation and introduce

hierarchical porosity throughout the 3D-printed structure, they selected 3D printing ink consisting

of PIM-1, THF, and DMAc. With proper optimization of ternary polymer solution and printing

atmosphere, the 3D-printed scaffold was able to purify 1000 ppm of toluene vapor in N2 gas for

1.7 h, which was six times longer than the original PIM-1. This solution-based additive

manufacturing method lays the foundation for the subsequent development of 3D-printable

materials over a wide range of polymers.

5.6. Overview of the advancement of PIMs membranes in arduous industrial applications

An extensive exploration of PIMs materials in various forms has showcased their versatility in

addressing the arduous industrial application. Thus far, the exploitation in these applications

remains limited, even though PIMs feature solution-processability, tunable porosity and

functionality. A few drawbacks impede the commercial implementation of PIMs such as fast aging

properties, relatively high cost of materials and purification processes, and lack of pilot- or

commercial-scale production, necessitating perpetual efforts in this regard. Nevertheless, PIMs, in

the form of powder, film, membranes and fibrous, have been witnessed to demonstrate remarkable

potential, especially in water treatment and electrochemical energy storage and conversion.

Leveraging their high surface area and adsorption capacity, PIMs have been applied in the

treatment of phenolic pollutants, ionic materials, and antibiotics from wastewater. Meanwhile, the

inherent rigidity of PIMs monomer has signified their utilization in tailoring the porosity of the

selective layer of TFC membranes for the separation of organic solvents. In addition, PIMs offer

versatile platforms for electrochemical energy storage and conversion, with potential applications

116
ranging from suppression of lithium dendritic growth and separators to redox flow cells and fuel

cells. It is anticipated that the wide implementation of PIMs materials in various applications can

be witnessed in the near future through ceaseless innovation and development.

a) b)

c) d)

h)

Figure 17. (a) Schematic of highly selective permeable fiber mat and fiber mat structure of PIM-

1, PIM/PAN, and PIM/PAN/UiO-66-NH2 [355]. (b) Morphology of PW12@PIM-1-AO fibers

and illustration of CEES oxidation to CEESO in the presence of PW12@PIM-1-AO [373]. (c)

Illustration of lithium dendrite growth on Cu electrode upon Li deposition and suppressed

lithium dendrite growth by guided Li deposition on PIM-DMBP-TB-coated Cu electrode [375].

(d) 3D-printed PIM-1 scaffolds created from PIM-1/THF/DMAc ternary ink formulation (The

117
inset scale bars are 1 mm) [421]. Copyright 2018, 2019, 2020, and 2022. Reproduced with

permission from American Chemical Society, Elsevier and Wiley-VCH.

6. Conclusion and Future Outlook

In summary, this review accentuated the recent progress on ladder PIMs, and fundamentally

scrutinized the intrinsic microporous structure of PIMs in applying to multiple fields, including

carbon capture, olefin/paraffin separation, water treatment, catalysis, sensing and energy

applications. Ladder PIMs with varying bulky and sterically hindered building blocks such as

spirocyclic, bridged bicyclic moieties, Tröger’s base, and CANAL, possess extraordinary

microporosity and chain-rigidity with tunable inter-chain distance for various applications. They

also exhibited excellent solubility properties and compatibility with different organic molecules,

benefiting from the rich functional groups and active sites of the polymer matrix. A thorough

investigation of the current research trend on PIMs has showcased the dominant usage of this

material in membrane technology or any separation or adsorption application that requires well-

defined porosity. Precisely, this review has elucidated the design of microstructure on PIMs from

five aspects, namely (1) polymer backbone constituent modification with bulky building blocks,

(2) post-functionalization on contortion unit, (3) blends with different polymers, (4) fabrication of

mixed matrix membrane through the incorporation of nanoparticles such as MOFs and MOPs, and

lastly, (5) revamping membranes with a series of post-treatment. Thereafter, the aforementioned

intrinsic architecture has remarkably advanced the structural properties as well as ameliorated the

performance of PIMs membranes in carbon capture. Correspondingly, the exploration of PIMs for

light hydrocarbons (e.g., C2 and C3) separations demonstrated positive demand, in which the

significance of ethylene and propylene in many chemical industries has driven the research

endeavors to develop an energy-efficient technology to recover these precious substances.

118
Considering the outstanding stability and efficiency, PIMs-based adsorbents were found

applicable to deal with different substances, even under extreme operating conditions. The PIMs-

based membranes also demonstrated attractive interest and performances in water treatment.

Noteworthy, remarkable performances of PIMs in the application of catalysis, energy, and sensing

were outlined, displaying the versatility of this polymer.

While the impressive use of PIMs has been observed over the past decades, it is crucial to persist

in the development of more practical and feasible PIMs to meet the demands of real-time industrial

applications. Consequently, the subsequent perspectives are delineated for future research and

studies in this direction.

(1) PIMs are synthesized through three proven routes, namely dibenzodioxin, Tröger’s base, and

imide linkages. To date, most of the advancements in PIM materials were found on spirocyclic-

based (e.g., spirobisindane) and bridged bicyclic-based (e.g., triptycene) ladder PIMs. Oppositely,

less amount of studies were focused on the polymerization of Tröger’s base and CANAL PIMs.

Structural tuning of PIMs is essential to take full advantage of these high-performing materials in

boosting their implementation in various fields.

(2) There are a few drawbacks that impede the commercial implementation of PIMs as promising

polymer materials in the industry such as fast aging properties, relatively high cost of materials,

and lack of true-scale industrial production compared to the other commercially available

polymers. In this regard, the costs of precursors and purification processes in PIMs production

were found to be dominant over the other costs, which shall be reduced through pilot-scale

production. A holistic approach is pivotal to render a lower production cost of PIMs while

maintaining their intrinsic characteristics for industrial application.

119
(3) The adoption of green substitutes represents an evolutionary move towards achieving

sustainability objectives. However, despite the environmentally friendly approach, these non-toxic

solvents are relatively costly for polymer manufacturing, primarily due to the lack of true-scale

industrial production. To truly realize green remediation, it is imperative to scrutinize the entire

production process, from the beginning stages (e.g., raw materials production and purification) to

the final steps (e.g., washing and drying) in producing greener PIMs.

(4) In addition to the practical synthesis and experiments, implementing more machine learning

on PIMs is a promising technology to provide functional models of polymer chemistry based on

quantitative data. This approach will be useful for screening and predicting the properties of PIMs,

such as gas diffusivity and solubility, free volume, hydrophobicity and conductivity.

(5) Permeability/selectivity trade-off, aging and CO2 plasticization in PIMs-based membranes are

the main constraints on the commercialization of these membranes despite the excellent gas

separation performances, especially in carbon capture. Likewise, these constraints also limited the

variety of commercial membrane materials used in membrane-based gas separation, with only 10

to 15 different types of membrane materials developed over the past 40 years. Architectural design

in membrane properties such as post-modification, co-polymerization and incorporation of

functional materials may effectively suppress the inherent chain relaxation in PIMs as a result of

insufficient packing of the polymer chain. In addition to the structurally tailored fillers either on

reduced particle size or with functional groups to enhance permeability, or by creating high

transport resistance to improve selectivity as well as addressing physical aging issue, next

generation of PIMs-based MMMs could be directed to gel membranes with remarkable chain

flexibility and free volume, as well as smart membranes such as temperature- or pressure-

responsive membranes with interchangeable properties tuned by an applied stimulus for

120
groundbreaking CO2 separation performance. However, the underlying mechanism to surmount

these limiting factors in diverse applications of PIMs-based membranes devotes intensive and

fundamental investigation. Thereafter, PIMs membrane development could be expanded to

effective chemoselective adsorption and separations.

(6) Membrane separation emerged as one of the highly energy-efficient and environmentally

benign technologies for carbon capture. The application of membrane technology in the post-

combustion process is challenging, owing to the low CO2 partial pressure (e.g., 0.03-0.2 atm) and

concentration (e.g., 3-20%) in the post-combustion flue gases. On the other hand, membrane

technology is more favourable for the pre-combustion process with CO2 concentration of 15-60%

(dry basis) at a total pressure of 20-70 atm. Hence, it is necessary to innovate on robust materials

that can withstand harsh operating conditions in pre-combustion flue gas treatment such as high

pressure or moisture environment, while adhering to the stringent purity requirements enforced by

the authorities. Meanwhile, the integration of membrane-based carbon capture systems into

existing facilities without disrupting the operation can be challenging. Careful planning and

coordination are necessary to retrofit the existing infrastructures with newly developed membrane

modules. A hybrid capture system by integrating two or more carbon capture technologies (e.g.,

adsorption, membrane separation and cryogenic distillation) could also be realized to address the

challenge of involvement of high investment cost of new equipment by combining the existing

facilities.

(7) The close physiochemical properties of light olefin/paraffin molecules pose significant

challenges for employing PIMs in hydrocarbon separation. Even though polymeric membranes

have been commercialized for large-scale gas separation applications, it is only limited to olefin

recovery in hydrocarbon separation. The implementation of polymeric membranes is limited by

121
the low thermal and chemical stability, inherent permeability/selectivity trade-off and

plasticization at high pressure. Thus, novel strategies are essential to exploit the feasibility and

robustness of PIMs membranes in real industrial conditions. The next-generation PIMs membranes

shall be robust and practical to withstand the feed streams containing impurities and high operating

pressure, substituting the energy-intensive hydrocarbon separation technology.

(8) While demonstrating impressive implementation in filtration and adsorption applications, only

a handful of studies were found using PIMs materials in other applications. Recent studies have

highlighted the viability of PIMs materials in the fields of catalysis, energy and sensing. Hence,

continuous research studies are essential to harness its potential and integration into these sectors.

Moreover, a number of promising studies were witnessed on the exploration of carbonized PIMs

membranes in hydrocarbon separation and electrocatalysis application, benefitting from the

porosity, conductivity and flexibility of this material. The further functionalization of these

conductive materials with noble metal opens up innovative possibilities for their utilization in

electrocatalytic processes, including but not limited to water splitting and hydrogen evolution.

As in the overview, the work synopsizes the universality of PIMs in addressing industrial

challenges. Undeniably, the structural versatility of PIMs could favor the continuous research

progress with various modification directions as an efficient and feasible material. Through

ceaseless efforts and fostered collaboration between academia and industry, it is prospected that

future innovation could drive the production of PIMs-based novel functional materials for value-

added application in energy and environmental benign.

Declaration of Competing Interest

122
The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

The authors would like to acknowledge the financial support provided by National Natural Science

Foundation of China (grant number: 22108118), State Key Laboratory of Physical Chemistry of

Solid Surfaces, Xiamen University (grant number: 2024X32), Xiamen University Malaysia

Research Fund (grant number: XMUMRF/2021-C8/IENG/0042), and Hengyuan International Sdn.

Bhd. (grant number: EENG/0003).

References

[1] T. Masuda, E. Isobe, T. Higashimura, K. Takada, Poly[1-(trimethylsilyl)-1-propyne]: A new


high polymer synthesized with transition-metal catalysts and characterized by extremely high gas
permeability, J. Am. Chem. Soc. 105 (1983) 7473-7474. https://doi.org/10.1021/ja00363a061
[2] I. Pinnau, L.G. Toy, Transport of organic vapors through poly(1-trimethylsilyl-1-propyne), J.
Membr. Sci. 116 (1996) 199-209. https://doi.org/10.1016/0376-7388(96)00041-5
[3] K. Nagai, T. Masuda, T. Nakagawa, B.D. Freeman, I. Pinnau, Poly[1-(trimethylsilyl)-1-
propyne] and related polymers: Synthesis, properties and functions, Prog. Polym. Sci. 26 (2001)
721-798. https://doi.org/10.1016/S0079-6700(01)00008-9
[4] P.M. Budd, E.S. Elabas, B.S. Ghanem, S. Makhseed, N.B. McKeown, K.J. Msayib, C.E.
Tattershall, D. Wang, Solution-processed, organophilic membrane derived from a polymer of
intrinsic microporosity, Adv. Mater. 16 (2004) 456-459. https://doi.org/10.1002/adma.200306053
[5] Z.X. Low, P.M. Budd, N.B. McKeown, D.A. Patterson, Gas permeation properties, physical
aging, and its mitigation in high free volume glassy polymers, Chem. Rev. 118 (2018) 5871-5911.
https://doi.org/10.1021/acs.chemrev.7b00629
[6] M. Balcik, Y. Wang, I. Pinnau, Exploring the effect of intra-chain rigidity on mixed-gas
separation performance of a Triptycene-Tröger's base ladder polymer (PIM-Trip-TB) by atomistic
simulations, J. Membr. Sci. 677 (2023) 121614. https://doi.org/10.1016/j.memsci.2023.121614
[7] Y. Wang, X. Ma, B.S. Ghanem, F. Alghunaimi, I. Pinnau, Y. Han, Polymers of intrinsic
microporosity for energy-intensive membrane-based gas separations, Mater. Today Nano 3 (2018)
69-95. https://doi.org/10.1016/j.mtnano.2018.11.003
[8] N.B. McKeown, P.M. Budd, Polymers of intrinsic microporosity (PIMs): Organic materials
for membrane separations, heterogeneous catalysis and hydrogen storage, Chem. Soc. Rev. 35
(2006) 675-683. https://doi.org/10.1039/b600349d
[9] N.B. McKeown, B. Gahnem, K.J. Msayib, P.M. Budd, C.E. Tattershall, K. Mahmood, S. Tan,
D. Book, H.W. Langmi, A. Walton, Towards polymer ‐ based hydrogen storage materials:

123
Engineering ultramicroporous cavities within polymers of intrinsic microporosity, Angew. Chem.
Int. Ed. 118 (2006) 1836-1839. https://doi.org/10.1002/anie.200504241
[10] W.L. Song, Y.X. Tang, M.X. Zhang, D.G. Yu, Current developments of hypercrosslinked
polymers as green carbon resources, Curr. Opin. Green Sustain. Chem. 5 (2022) 100335.
https://doi.org/10.1016/j/crgsc.2022.100335
[11] J. Xiao, J. Chen, J. Liu, H. Ihara, H.D. Qiu, Synthesis strategies of covalent organic
frameworks: An overview from nonconventional heating methods and reaction media, Green
Energy Environ. 8 (2023) 1695-1618. https://doi.org/10.1016/j.gee.2022.05.003
[12] W. Bai, Y. Fan, F. Wang, P. Mu, H. Sun, Z. Zhu, W. Liang, A. Li, Facile synthesis of porous
organic polymers (POPs) membrane via click chemistry for efficient PM2.5 capture, Sep. Purif.
Technol. 258 (2021) 118049. https://doi.org/10.1016/j.seppur.2020.118049
[13] P. Budd, N. McKeown, B. Ghanem, K. Msayib, D. Fritsch, L. Starannikova, N. Belov, O.
Sanfirova, Y. Yampolskii, V. Shantarovich, Gas permeation parameters and other
physicochemical properties of a polymer of intrinsic microporosity: Polybenzodioxane PIM-1, J.
Membr. Sci. 325 (2008) 851-860. https://doi.org/10.1016/j.memsci.2008.09.010
[14] N.Y. Du, G.P. Robertson, J.S. 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. https://doi.org/10.1021/ma801858d
[15] J.M. Luque-Alled, M. Tamaddondar, A.B. Foster, P.M. Budd, P. Gorgojo, PIM-1/holey
graphene oxide mixed matrix membranes for gas separation: Unveiling the role of holes, ACS
Appl. Mater. Interfaces 13 (2021) 55517-55533. https://doi.org/10.1021/acsami.1c15640
[16] S. Mohsenpour, A.W. Ameen, S. Leaper, C. Skuse, F. Almansour, P.M. Budd, P. Gorgojo,
PIM-1 membranes containing POSS - graphene oxide for CO2 separation, Sep. Purif. Technol. 298
(2022) 121447. https://doi.org/10.1016/j.seppur.2022.121447
[17] K. Chen, L. Ni, H. Zhang, C. Xiao, L. Li, X. Guo, J. Qi, C. Wang, X. Sun, J. Li, Incorporating
KAUST-7 into PIM-1 towards mixed matrix membranes with long-term stable CO2/CH4
separation performance, J. Membr. Sci. 661 (2022) 120848.
https://doi.org/10.1016/j.memsci.2022.120848
[18] W. Chen, Z. Zhang, L. Hou, C. Yang, H. Shen, K. Yang, Z. Wang, Metal-organic framework
MOF-801/PIM-1 mixed-matrix membranes for enhanced CO2/N2 separation performance, Sep.
Purif. Technol. 250 (2020) 117198. https://doi.org/10.1016/j.seppur.2020.117198
[19] A.K. Sekizkardes, S. Budhathoki, L. Zhu, V. Kusuma, Z. Tong, J.S. McNally, J.A. Steckel,
S. Yi, D. Hopkinson, Molecular design and fabrication of PIM-1/polyphosphazene blend
membranes with high performance for CO2/N2 separation, J. Membr. Sci. 640 (2021) 119764.
https://doi.org/10.1016/j.memsci.2021.119764
[20] W.F. Yong, T.-S. Chung, Miscible blends of carboxylated polymers of intrinsic microporosity
(cPIM-1) and Matrimid, Polymer 59 (2015) 290-297.
https://doi.org/10.1016/j.polymer.2015.01.013
[21] X. Zou, G. Zhu, Microporous organic materials for membrane-based gas separation, Adv.
Mater. 30 (2018) 1700750. https://doi.org/10.1002/adma.201700750
[22] Y. Wang, B.S. Ghanem, Z. Ali, K. Hazazi, Y. Han, I. Pinnau, Recent progress on polymers
of intrinsic microporosity and thermally modified analogue materials for membrane‐based fluid
separations, Small Struct. 2 (2021) 2100049. https://doi.org/10.1002/sstr.202100049
[23] Y. Wang, B.S. Ghanem, Y. Han, I. Pinnau, State-of-the-art polymers of intrinsic
microporosity for high-performance gas separation membranes, Curr. Opin. Chem. Eng. 35 (2022)
100755. https://doi.org/10.1016/j.coche.2021.100755

124
[24] M.Z. Ahmad, R. Castro-Munoz, P.M. Budd, Boosting gas separation performance and
suppressing the physical aging of polymers of intrinsic microporosity (PIM-1) by nanomaterial
blending, Nanoscale 12 (2020) 23333-23370. https://doi.org/10.1039/d0nr07042d
[25] S. He, B. Zhu, S. Li, Y. Zhang, X. Jiang, C.H. Lau, L. Shao, Recent progress in PIM-1 based
membranes for sustainable CO2 separations: Polymer structure manipulation and mixed matrix
membrane design, Sep. Purif. Technol. 284 (2022) 120277.
https://doi.org/10.1016/j.seppur.2021.120277
[26] L. Wang, Y. Zhao, B. Fan, M. Carta, R. Malpass-Evans, N.B. McKeown, F. Marken, Polymer
of intrinsic microporosity (PIM) films and membranes in electrochemical energy storage and
conversion: A mini-review, Electrochem. Commun. 118 (2020) 106798.
https://doi.org/10.1016/j.elecom.2020.106798
[27] A.R. Antonangelo, N. Hawkins, M. Carta, Polymers of intrinsic microporosity (PIMs) for
catalysis: A perspective, Curr. Opin. Chem. Eng. 35 (2022) 100766.
https://doi.org/10.1016/j.coche.2021.100766
[28] F. Topuz, M.H. Abdellah, P.M. Budd, M.A. Abdulhamid, Advances in polymers of intrinsic
microporosity (PIMs)-based materials for membrane, environmental, catalysis, sensing and energy
applications, Polym. Rev. (2023) 1-55. https://doi.org/10.1080/15583724.2023.2236677
[29] W.F. Yong, H. Zhang, Recent advances in polymer blend membranes for gas separation and
pervaporation, Prog. Mater. Sci. 116 (2021) 100713.
https://doi.org/10.1016/j.pmatsci.2020.100713
[30] L.M. Robeson, Correlation of separation factor versus permeability for polymeric membranes,
J. Membr. Sci. 62 (1991) 165-185. https://doi.org/10.1016/0376-7388(91)80060-J
[31] L.M. Robeson, The upper bound revisited, J. Membr. Sci. 320 (2008) 390-400.
https://doi.org/10.1016/j.memsci.2008.04.030
[32] R. Swaidan, B. Ghanem, I. Pinnau, Fine-tuned intrinsically ultramicroporous polymers
redefine the permeability/selectivity upper bounds of membranebased air and hydrogen
separations, ACS Macro Lett. 4 (2015) 947-951. https://doi.org/10.1021/acsmacrolett.5b00512
[33] B. Comesaña-Gándara, J. Chen, C.G. Bezzu, M. Carta, I. Rose, M.-C. Ferrari, E. Esposito, A.
Fuoco, J.C. Jansen, N.B. McKeown, Redefining the Robeson upper bounds for CO2/CH4 and
CO2/N2 separations using a series of ultrapermeable benzotriptycene-based polymers of intrinsic
microporosity, Energy Environ. Sci. 12 (2019) 2733-2740. https://doi.org/10.1039/c9ee01384a
[34] R.G. Yin, Z.Y. Wang, M. Bockstaller, K. Matyjaszewski, Tuning dispersity of linear polymers
and polymeric brushes grown from nanoparticles by atom transfer radical polymerization, Polym.
Chem. 12 (2021) 6071-6082. https://doi.org/10.1039/D1PY01178B
[35] M.R. Abbasi, A. Shamiri, M.A. Hussain, Dynamic modeling and molecular weight
distribution of ethylene copolymerization in an industrial gas-phase fluidized-bed reactor, Adv.
Powder Technol. 27 (2016) 1526-1538. 10.1016/j.apt.2016.05.014
[36] R. Whitfield, N.P. Truong, D. Messmer, K. Parkatzidis, M. Rolland, A. Anastasaki, Tailoring
polymer dispersity and shape of molecular weight distributions: Methods and applications, Chem.
Sci. 10 (2019) 8724-8734. https://doi.org/10.1039/C9SC03546J
[37] M.R. Abbasi, A. Shamiri, M.A. Hussain, Dynamic modeling and molecular weight
distribution of ethylene copolymerization in an industrial gas-phase fluidized-bed reactor, Adv.
Powder Technol. 27 (2016) 1526-1538. https://doi.org/10.1016/j.apt.2016.05.014
[38] M.M. Abdelghafour, A. Orban, A. Deak, L. Lamch, E. Frank, R. Nagy, A. Adam, P. Sipos,
E. Farkas, F. Bari, L. Janovak, The effect of molecular weight on the solubility properties of

125
biocompatible poly(ethylene succinate) polyester, Polymers 13 (2021) 2725.
https://doi.org/10.3390/polym13162725
[39] R.G. Yin, Z.Y. Wang, M. Bockstaller, K. Matyjaszewski, Tuning dispersity of linear polymers
and polymeric brushes grown from nanoparticles by atom transfer radical polymerization, Polym.
Chem. 12 (2021) 6071-6082. https://doi.org/10.1039/D1PY01178B
[40] N. Ghahramani, M. Rahmati, The effect of the molecular weight and polydispersity index on
the thermal conductivity of polyamide 6: A molecular dynamics study, Int. J. Heat Mass Transf.
154 (2020) 119487. https://doi.org/10.1016/j.ijheatmasstransfer.2020.119487
[41] B. Satilmis, M.N. Alnajrani, P.M. Budd, Hydroxyalkylaminoalkylamide PIMs: Selective
adsorption by ethanolamine- and diethanolamine-modified PIM-1, Macromolecules 48 (2015)
5663-5669. https://doi.org/10.1021/acs.macromol.5b01196
[42] A.B. Foster, M. Tamaddondar, J.M. Luque-Alled, W.J. Harrison, Z. Li, P. Gorgojo, P.M.
Budd, Understanding the topology of the polymer of intrinsic microporosity PIM-1: Cyclics,
tadpoles, and network structures and their impact on membrane performance, Macromolecules 53
(2020) 569-583. https://doi.org/10.1021/acs.macromol.9b02185
[43] P.M. Budd, B.S. Ghanem, S. Makhseed, N.B. McKeown, K.J. Msayib, C.E. Tattershall,
Polymers of intrinsic microporosity (PIMs): Robust, solution-processable, organic nanoporous
materials, Chem. Commun. (2004) 230-231. https://doi.org/10.1039/B311764B
[44] B. Ghanem, N. McKeown, P. Budd, D. Fritsch, Polymers of intrinsic microporosity derived
from bis(phenazyl) monomers, Macromolecules 41 (2008) 1640-1646.
https://doi.org/10.1021/ma071846r
[45] A. Fuoco, B. Satilmis, T. Uyar, M. Monteleone, E. Esposito, C. Muzzi, E. Tocci, M. Longo,
M.P. De Santo, M. Lanč, K. Friess, O. Vopička, P. Izák, J.C. Jansen, Comparison of pure and
mixed gas permeation of the highly fluorinated polymer of intrinsic microporosity PIM-2 under
dry and humid conditions: Experiment and modelling, J. Membr. Sci. 594 (2020) 117460.
https://doi.org/10.1016/j.memsci.2019.117460
[46] M. Cook, P.R.J. Gaffney, L.G. Peeva, A.G. Livingston, Roll-to-roll dip coating of three
different PIMs for organic solvent nanofiltration, J. Membr. Sci. 558 (2018) 52-63.
https://doi.org/10.1016/j.memsci.2018.04.046
[47] A. Mahajan, S.K. Bhattacharya, S. Rochat, A.D. Burrows, P.J. Fletcher, Y. Rong, A.B. Dalton,
N.B. McKeown, F. Marken, Polymer of intrinsic microporosity (PIM-7) coating affects triphasic
palladium electrocatalysis, ChemElectroChem 6 (2019) 4307-4317.
https://doi.org/10.1002/celc.201801359
[48] M. Tian, S. Rochat, H. Fawcett, A.D. Burrows, C.R. Bowen, T.J. Mays, Chemical
modification of the polymer of intrinsic microporosity PIM-1 for enhanced hydrogen storage,
Adsorption 26 (2020) 1083-1091. https://doi.org/10.1007/s10450-020-00239-y
[49] M.D. Guiver, M. Yahia, M.M. Dal-Cin, G.P. Robertson, S. Saeedi Garakani, N. Du, N.
Tavajohi, Gas transport in a polymer of intrinsic microporosity (PIM-1) substituted with pseudo-
ionic liquid tetrazole-type structures, Macromolecules 53 (2020) 8951-8959.
https://doi.org/10.1021/acs.macromol.0c01321
[50] J. Xu, S. Ma, Y. Li, X. Li, J. Ou, M. Ye, Thiol-functionalized PIM-1 for removal and sensing
for mercury (II), J. Environ. Chem. Eng. 8 (2020) 104545.
https://doi.org/10.1016/j.jece.2020.104545
[51] C.G. Bezzu, M. Carta, M.-C. Ferrari, J.C. Jansen, M. Monteleone, E. Esposito, A. Fuoco, K.
Hart, T.P. Liyana-Arachchi, C.M. Colina, N.B. McKeown, The synthesis, chain-packing
simulation and long-term gas permeability of highly selective spirobifluorene-based polymers of

126
intrinsic microporosity, J. Mater. Chem. A 6 (2018) 10507-10514.
https://doi.org/10.1039/c8ta02601g
[52] N.B. McKeown, The structure-property relationships of Polymers of Intrinsic Microporosity
(PIMs), Curr. Opin. Chem. Eng. 36 (2022) 100785. https://doi.org/10.1016/j.coche.2021.100785
[53] C.G. Bezzu, A. Fuoco, E. Esposito, M. Monteleone, M. Longo, J.C. Jansen, G.S. Nichol, N.B.
McKeown, Ultrapermeable polymers of intrinsic microporosity containing spirocyclic units with
fused triptycenes, Adv. Funct. Mater. 31 (2021) 2104474.
https://doi.org/10.1002/adfm.202104474
[54] T.J. Corrado, Z. Huang, D. Huang, N. Wamble, T. Luo, R. Guo, Pentiptycene-based ladder
polymers with configurational free volume for enhanced gas separation performance and physical
aging resistance, Proc. Natl. Acad. Sci. 118 (2021) e2022204118.
https://doi.org/10.1073/pnas.2022204118
[55] E. Al-Hetlani, M.O. Amin, A.R. Antonangelo, H.L. Zhou, M. Carta, Triptycene and
triphenylbenzene-based polymers of intrinsic microporosity (PIMs) for the removal of
pharmaceutical residues from wastewater, Microporous Mesoporous Mater. 330 (2022) 111602.
https://doi.org/10.1016/j.micromeso.2021.111602
[56] O. Vopička, M. Lanč, K. Friess, Phenomenology of vapour sorption in polymers of intrinsic
microporosity PIM-1 and PIM-EA-TB: Envelopment of sorption isotherms, Curr. Opin. Chem.
Eng. 35 (2022) 100786. https://doi.org/10.1016/j.coche.2021.100786
[57] T.H. Chen, T. Su, L. Xia, L.Y. Gong, B.N. Dong, L.F. Liu, Corolliform morphology thin film
composite polyamide membrane constructed via Troger's base PIM for enhancing nanofiltration
separation performances, Desalination 548 (2023) 116234.
https://doi.org/10.1016/j.desal.2022.116234
[58] S. Liu, Z. Jin, Y.C. Teo, Y. Xia, Efficient synthesis of rigid ladder polymers via palladium
catalyzed annulation, J. Am. Chem. Soc. 136 (2014) 17434-17437.
https://doi.org/10.1021/ja5110415
[59] H.W.H. Lai, F.M. Benedetti, J.M. Ahn, A.M. Robinson, Y.G. Wang, I. Pinnau, Z.P. Smith,
Y. Xia, Hydrocarbon ladder polymer with ultrahigh permselectivity for membrane gas separations,
Science 375 (2022) 1390-1392. https://doi.org/10.1126/science.abl7163
[60] H.W.H. Lai, F.M. Benedetti, Z. Jin, Y.C. Teo, A.X. Wu, M.G.D. Angelis, Z.P. Smith, Y. Xia,
Tuning the molecular weights, chain packing, and gas-transport properties of CANAL ladder
polymers by short alkyl substitutions, Macromolecules 52 (2019) 6294-6302.
https://doi.org/10.1021/acs.macromol.9b01155
[61] K. Hazazi, Y.G. Wang, N.M.S. Bettahalli, X.H. Ma, Y. Xia, I. Pinnau, Catalytic arene-
norbornene annulation (CANAL) ladder polymer derived carbon membranes with unparalleled
hydrogen/carbon dioxide size-sieving capability, J. Membr. Sci. 654 (2022) 120548.
https://doi.org/10.1016/j.memsci.2022.120548
[62] L. Fu, Z. Ren, W. Si, Q. Ma, W. Huang, K. Liao, Z. Huang, Y. Wang, J. Li, P. Xu, Research
progress on CO2 capture and utilization technology, J. CO2 Util. 66 (2022) 102260.
https://doi.org/10.1016/j.jcou.2022.102260
[63] K. Wei, H. Guan, Q. Luo, J. He, S. Sun, Recent advances in CO2 capture and reduction,
Nanoscale 14 (2022) 11869-11891. https://doi.org/10.1039/d2nr02894h
[64] G. Chen, T. Wang, G. Zhang, G. Liu, W. Jin, Membrane materials targeting carbon capture
and utilization, Adv. Membr. 2 (2022) 100025. https://doi.org/10.1016/j.advmem.2022.100025

127
[65] H. Guo, J. Wei, Y. Ma, J. Deng, S. Yi, B. Wang, L. Deng, X. Jiang, Z. Dai, Facilitated
transport membranes for CO2/CH4 separation - State of the art, Adv. Membr. 2 (2022) 100040.
https://doi.org/10.1016/j.advmem.2022.100040
[66] B. Zhu, S. He, Y. Wu, S. Li, L. Shao, One-step synthesis of structurally stable CO2-philic
membranes with ultra-high PEO loading for enhanced carbon capture, Engineering 26 (2023) 220-
228. https://doi.org/10.1016/j.eng.2022.03.016
[67] H.S. Lau, W.F. Yong, Recent progress and prospects of polymeric hollow fiber membranes
for gas application, water vapor separation and particulate matter removal, J. Mater. Chem. A 9
(2021) 26454-26497. https://doi.org/10.1039/d1ta07093b
[68] Z. Wang, Q. Shen, J. Liang, Y. Zhang, J. Jin, Adamantane-grafted polymer of intrinsic
microporosity with finely tuned interchain spacing for improved CO2 separation performance, Sep.
Purif. Technol. 233 (2020) 116008. https://doi.org/10.1016/j.seppur.2019.116008
[69] N.B. McKeown, P.M. Budd, Exploitation of intrinsic microporosity in polymer-based
materials, Macromolecules 43 (2010) 5163-5176. https://doi.org/10.1021/ma1006396
[70] P.M. Budd, N.B. McKeown, D. Fritsch, Free volume and intrinsic microporosity in polymers,
J. Mater. Chem. 15 (2005) 1977. https://doi.org/10.1039/B417402J
[71] P.M. Budd, K.J. Msayib, C.E. Tattershall, B.S. Ghanem, K.J. Reynolds, N.B. McKeown, D.
Fritsch, Gas separation membranes from polymers of intrinsic microporosity, J. Membr. Sci. 251
(2005) 263-269. https://doi.org/10.1016/j.memsci.2005.01.009
[72] H. Sato, S. Nakajo, Y. Oishi, Y. Shibasaki, Synthesis of linear polymer of intrinsic
microporosity from 5,5 ′ ,6,6 ′ -tetrahydroxy-3,3,3 ′ ,3 ′ -tetramethylspirobisindane and
decafluorobiphenyl, React. Funct. Polym. 125 (2018) 70-76.
https://doi.org/10.1016/j.reactfunctpolym.2018.02.006
[73] P. Stanovsky, M. Karaszova, Z. Petrusova, M. Monteleone, J.C. Jansen, B. Comesaña-
Gándara, N.B. McKeown, P. Izak, Upgrading of raw biogas using membranes based on the
ultrapermeable polymer of intrinsic microporosity PIM-TMN-Trip, J. Membr. Sci. 618 (2021)
118694. https://doi.org/10.1016/j.memsci.2020.118694
[74] S.A. Felemban, C.G. Bezzu, B. Comesaña-Gándara, J.C. Jansen, A. Fuoco, E. Esposito, M.
Carta, N.B. McKeown, Synthesis and gas permeation properties of tetraoxidethianthrene-based
polymers of intrinsic microporosity, J. Mater. Chem. A 9 (2021) 2840-2849.
https://doi.org/10.1039/d0ta10134f
[75] H. Sun, W. Gao, Y. Zhang, X. Cao, S. Bao, P. Li, Z. Kang, Q.J. Niu, Bis(phenyl)fluorene-
based polymer of intrinsic microporosity/functionalized multi-walled carbon nanotubes mixed
matrix membranes for enhanced CO2 separation performance, React. Funct. Polym. 147 (2020)
104465. https://doi.org/10.1016/j.reactfunctpolym.2019.104465
[76] M. Longo, M.P. De Santo, E. Esposito, A. Fuoco, M. Monteleone, L. Giorno, B. Comesaña-
Gándara, J. Chen, C.G. Bezzu, M. Carta, Correlating gas permeability and Young’s modulus
during the physical aging of polymers of intrinsic microporosity using atomic force microscopy,
Ind. Eng. Chem. Res. 59 (2019) 5381-5391.
[77] C.H. Ma, Q.H. Li, Q. Wang, M. Gao, J.X. Wang, X.Z. Cao, High performance membranes
containing rigid contortion units prepared by interfacial polymerization for CO2 separation, J.
Membr. Sci. 652 (2022) 120459. https://doi.org/10.1016/j.memsci.2022.120459
[78] S. Kang, G. Huo, Z. Zhang, T. Guo, Z. Dai, N. Li, Polymers of intrinisic porosity with bulky
side groups on the spirobisindane moieties for gas separation, ACS Appl. Mater. Interfaces 5 (2023)
8660-8669. https://doi.org/10.1021/acsapm.3c01762

128
[79] J. Chen, M. Longo, A. Fuoco, E. Esposito, M. Monteleone, B. Comesana Gandara, J. Carolus
Jansen, N.B. McKeown, Dibenzomethanopentacene-based polymers of intrinsic microporosity for
use in gas-separation membranes, Angew. Chem. Int. Ed. 62 (2023) e202215250.
https://doi.org/10.1002/anie.202215250
[80] A. Devarajan, E.D. Asuquo, M.Z. Ahmad, A.B. Foster, P.M. Budd, Influence of polymer
topology on gas separation membrane performance of the polymer of intrinsic microporosity PIM-
Py, ACS Appl. Polym. Mater. 3 (2021) 3485-3495. https://doi.org/10.1021/acsapm.1c00415
[81] S. Zhou, J. Guan, Z. Li, Q. Zhang, J. Zheng, S. Li, S. Zhang, Synthesis of fluorinated
poly(phenyl-alkane)s of intrinsic microporosity by regioselective aldehyde (A2) + aromatics (B2)
Friedel–Crafts polycondensation, Macromolecules 54 (2021) 6543-6551.
https://doi.org/10.1021/acs.macromol.1c00468
[82] Y. Weng, Q. Li, J. Li, Z. Gao, L. Zou, X. Ma, Facile synthesis of Bi-functionalized intrinsic
microporous polymer with fully carbon backbone for gas separation application, Sep. Purif.
Technol. 279 (2021) 119681. https://doi.org/10.1016/j.seppur.2021.119681
[83] I. Kammakakam, K.E. O’Harra, J.E. Bara, E.M. Jackson, Spirobisindane-containing
imidazolium polyimide ionene: Structural design and gas separation performance of “Ionic PIMs”,
Macromolecules 55 (2022) 4790-4802. https://doi.org/10.1021/acs.macromol.1c02317
[84] M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, J.C. Jansen, P. Bernardo, F. Bazzarelli,
N.B. McKeown, An efficient polymer molecular siece for membrane gas separation, Science 339
(2013) 303-307. https://doi.org/10.1126/science.1228032
[85] M.A. Spielman, The structure of Troger's base, J. Am. Chem. Soc. 57 (1935) 583-585.
https://doi.org/10.1021/ja01306a060
[86] C. Zhang, J. Yan, Z. Tian, X. Liu, B. Cao, P. Li, Molecular Design of Tröger’s Base-Based
Polymers Containing Spirobichroman Structure for Gas Separation, Ind. Eng. Chem. Res. 56 (2017)
12783-12788. https://doi.org/10.1021/acs.iecr.7b03434
[87] R. Malpass-Evans, I. Rose, A. Fuoco, P. Bernardo, G. Clarizia, N.B. McKeown, J.C. Jansen,
M. Carta, Effect of bridgehead methyl substituents on the gas permeability of Troger's-base
derived polymers of intrinsic microporosity, Membranes 10 (2020) 62.
https://doi.org/10.3390/membranes10040062
[88] A.R. Antonangelo, N. Hawkins, E. Tocci, C. Muzzi, A. Fuoco, M. Carta, Troger's base
network polymers of intrinsic microporosity (TB-PIMs) with tunable pore size for heterogeneous
catalysis, J. Am. Chem. Soc. 144 (2022) 15581-15594. https://doi.org/10.1021/jacs.2c04739
[89] R. Williams, L.A. Burt, E. Esposito, J.C. Jansen, E. Tocci, C. Rizzuto, M. Lanč, M. Carta,
N.B. McKeown, A highly rigid and gas selective methanopentacene-based polymer of intrinsic
microporosity derived from Tröger's base polymerization, J. Mater. Chem. A 6 (2018) 5661-5667.
https://doi.org/10.1039/c8ta00509e
[90] 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, J. Membr. Sci. 457 (2014) 95–102.
https://doi.org/10.1016/j.memsci.2014.01.055
[91] X. Ma, I. Pinnau, A novel intrinsically microporous ladder polymer and copolymers derived
from 1,1′,2,2′-tetrahydroxy-tetraphenylethylene for membrane-based gas separation, Polym.
Chem. 7 (2016) 1244-1248. https://doi.org/10.1039/C5PY01796C
[92] Y. Wang, V. Kumar, F. Elahi, B. Ghanem, M. Balcik, J. Shen, Y. Han, I. Pinnau, Amidoxime-
functionalized tetraphenylethylene ladder polymer for efficient membrane-based gas separations,
Eur. Polym. J. 209 (2024) 112896. https://doi.org/10.1016/j.eurpolymj.2024.112896

129
[93] J.B. Tang, W.L. Sun, H.D. Tang, M. Radosz, Y.Q. Shen, Enhanced CO2 absorption of
poly(ionic liquid)s, Macromolecules 38 (2005) 2037-2039. https://doi.org/10.1021/ma047574z
[94] K.M. Rodriguez, S. Lin, A.X. Wu, G. Han, J.J. Teesdale, C.M. Doherty, Z.P. Smith,
Leveraging free volume manipulation to improve the membrane separation performance of amine-
functionalized PIM-1, Angew. Chem. Int. Ed. 60 (2021) 6593-6599.
https://doi.org/10.1002/anie.202012441
[95] S.-L. Li, Z. Zhu, J. Li, Y. Hu, X. Ma, Synthesis and gas separation properties of OH-
functionalized Tröger's base-based PIMs derived from 1,1′-binaphthalene-2,2′-OH, Polymer
193 (2020) 122369. https://doi.org/10.1016/j.polymer.2020.122369
[96] Y. Peng, Z. Xia, Y. Song, Y. Zhang, Y. Li, P. Zhang, Highly carboxylic acid functionalized
PIM‐1 by hydrothermal method: Mechanistic study of gas separation properties, Nano Select 3
(2022) 1395-1406. https://doi.org/10.1002/nano.202200082
[97] N. Du, G.P. Robertson, J. Song, I. Pinnau, M.D. Guiver, High-performance carboxylated
polymers of intrinsic microporosity (PIMs) with tunable gas transport properties, Macromolecules
42 (2009) 6038-6043. https://doi.org/10.1021/ma9009017
[98] B. Satilmis, P.M. Budd, Base-catalysed hydrolysis of PIM-1: Amide versus carboxylate
formation, RSC Adv. 4 (2014) 52189-52198. https://doi.org/10.1039/C4RA09907A
[99] W.F. Yong, T.-S. Chung, Mechanically strong and flexible hydrolyzed polymers of intrinsic
microporosity (PIM-1) membranes, J. Polym. Sci. Part B: Polym. Phys. 55 (2017) 344-354.
https://doi.org/10.1002/polb.24279
[100] K.M. Rodriguez, A.X. Wu, Q. Qian, G. Han, S. Lin, F.M. Benedetti, H. Lee, W.S. Chi, C.M.
Doherty, Z.P. Smith, Facile and time-tfficient carboxylic acid functionalization of PIM-1: Effect
on molecular packing and gas separation performance, Macromolecules 53 (2020) 6220-6234.
https://doi.org/10.1021/acs.macromol.0c00933
[101] W. Han, C. Zhang, M. Zhao, F. Yang, Y. Yang, Y. Weng, Post-modification of PIM-1 and
simultaneously in situ synthesis of porous polymer networks into PIM-1 matrix to enhance CO2
separation performance, J. Membr. Sci. 636 (2021) 119544.
https://doi.org/10.1016/j.memsci.2021.119544
[102] W.H. Wu, P. Thomas, P. Hume, J. Jin, Effective conversion of amide to carboxylic acid on
polymers of intrinsic microporosity (PIM-1) with nitrous acid, Membranes 8 (2018) 20.
https://doi.org/10.3390/membranes8020020
[103] H. Ling, O.T. Qazvini, S.G. Telfer, J. Jin, Effective enhancement of selectivities and
capacities for CO2 over CH4 and N2 of polymers of intrinsic microporosity via postsynthesis
metalation, J. Polym. Sci. 58 (2020) 2619-2624. https://doi.org/10.1002/pol.20200104
[104] X. Chen, Z. Zhang, L. Wu, X. Liu, S. Xu, J.E. Efome, X. Zhang, N. Li, Polymers of intrinsic
microporosity having bulky substitutes and cross-linking for gas separation membranes, ACS Appl.
Polym. Mater. 2 (2020) 987-995. https://doi.org/10.1021/acsapm.9b01193
[105] K. Halder, P. Georgopanos, S. Shishatskiy, V. Filiz, V. Abetz, Investigation of gas transport
and other physical properties in relation to the bromination degree of polymers of intrinsic
microporosity, J. Polym. Sci. Part A: Polym. Chem. 56 (2018) 2752-2761.
https://doi.org/10.1002/pola.29262
[106] K. Halder, S. Neumann, G. Bengtson, M.M. Khan, V. Filiz, V. Abetz, Polymers of intrinsic
microporosity postmodified by vinyl groups for membrane applications, Macromolecules 51
(2018) 7309-7319. https://doi.org/10.1021/acs.macromol.8b01252

130
[107] K. Mizrahi Rodriguez, F.M. Benedetti, N. Roy, A.X. Wu, Z.P. Smith, Sorption-enhanced
mixed-gas transport in amine functionalized polymers of intrinsic microporosity (PIMs), J. Mater.
Chem. A 9 (2021) 23631-23642. https://doi.org/10.1039/D1TA06530K
[108] W.F. Yong, F.Y. Li, Y.C. Xiao, P. Li, K.P. Pramoda, Y.W. Tong, T.S. Chung, Molecular
engineering of PIM-1/Matrimid blend membranes for gas separation, J. Membr. Sci. 407-408
(2012) 47-57. https://doi.org/10.1016/j.memsci.2012.03.038
[109] W.F. Yong, F.Y. Li, T.-S. Chung, Y.W. Tong, Highly permeable chemically modified PIM-
1/Matrimid membranes for green hydrogen purification, J. Mater. Chem. A 1 (2013) 13914-13925.
https://doi.org/10.1039/C3TA13308G
[110] E. Esposito, I. Mazzei, M. Monteleone, A. Fuoco, M. Carta, N.B. McKeown, R. Malpass-
Evans, J.C. Jansen, Highly permeable Matrimid®/PIM-EA(H2)-TB blend membrane for gas
separation, Polymers 11 (2018) 46. https://doi.org/10.3390/polym11010046
[111] L. Wang, X. Guo, F. Zhang, N. Li, Blending and in situ thermally crosslinking of dual rigid
polymers for anti-plasticized gas separation membranes, J. Membr. Sci. 638 (2021) 119668.
https://doi.org/10.1016/j.memsci.2021.119668
[112] L. Hou, Z. Wang, Z. Chen, W. Chen, C. Yang, PIM-1 as an organic filler to enhance CO2
separation performance of poly (arylene fluorene ether ketone), Sep. Purif. Technol. 242 (2020)
116766. https://doi.org/10.1016/j.seppur.2020.116766
[113] M. Longo, M. Monteleone, E. Esposito, A. Fuoco, E. Tocci, M.-C. Ferrari, B. Comesaña-
Gándara, R. Malpass-Evans, N.B. McKeown, J.C. Jansen, Thin film composite membranes based
on the polymer of intrinsic microporosity PIM-EA(Me2)-TB blended with Matrimid®5218,
Membranes 12 (2022) 881. https://doi.org/10.3390/membranes12090881
[114] A.K. Sekizkardes, V.A. Kusuma, J.S. McNally, David W. Gidley, K. Resnik, S.R. Venna,
D. Hopkinson, Microporous polymeric composite membranes with advanced film properties: Pore
intercalation yields excellent CO2 separation performance, J. Mater. Chem. A 6 (2018) 22472-
22477. https://doi.org/10.1039/c8ta07424k
[115] J. Sánchez-Laínez, B. Zornoza, M. Carta, R. Malpass-Evans, N.B. McKeown, C. Téllez, J.
Coronas, Hydrogen separation at high temperature with dense and asymmetric membranes based
on PIM-EA(H2)-TB/PBI blends, Ind. Eng. Chem. Res. 57 (2018) 16909-16916.
https://doi.org/10.1021/acs.iecr.8b04209
[116] G.S. Larsen, P. Lin, K.E. Hart, C.M. Colina, Molecular simulations of PIM-1-like polymers
of intrinsic microporosity, Macromolecules 44 (2011) 6944-6951.
https://doi.org/10.1021/ma200345v
[117] S. Zhao, J. Liao, D. Li, X. Wang, N. Li, Blending of compatible polymer of intrinsic
microporosity (PIM-1) with Tröger's Base polymer for gas separation membranes, J. Membr. Sci.
566 (2018) 77-86. https://doi.org/10.1016/j.memsci.2018.08.010
[118] X. Chen, Y. Fan, L. Wu, L. Zhang, D. Guan, C. Ma, N. Li, Ultra-selective molecular-sieving
gas separation membranes enabled by multi-covalent-crosslinking of microporous polymer blends,
Nat. Commun. 12 (2021) 6140. https://doi.org/10.1038/s41467-021-26379-5
[119] F. Almansour, A.B. Foster, A.W. Ameen, S. Mohsenpour, P.M. Budd, P. Gorgojo, High gas
permeance in CO2-selective thin film composite membranes from bis(phenyl)fluorene-containing
blends with PIM-1, J. Membr. Sci. 699 (2024) 122652.
https://doi.org/10.1016/j.memsci.2024.122652
[120] P. Wang, Z. Liu, X. Li, J. Peng, W. Hu, B. Liu, Toward enhanced conductivity of high-
temperature proton exchange membranes: Development of novel PIM-1 reinforced PBI alloy
membranes, Chem. Commun. 55 (2019) 6491-6494. https://doi.org/10.1039/C9CC02102G

131
[121] L. Hao, P. Li, T.-S. Chung, PIM-1 as an organic filler to enhance the gas separation
performance of Ultem polyetherimide, J. Membr. Sci. 453 (2014) 614-623.
https://doi.org/10.1016/j.memsci.2013.11.045
[122] W.F. Yong, Z.K. Lee, T.S. Chung, M. Weber, C. Staudt, C. Maletzko, Blends of a polymer
of intrinsic microporosity and partially sulfonated polyphenylenesulfone for gas separation,
ChemSusChem 9 (2016) 1953-1962. https://doi.org/10.1002/cssc.201600354
[123] 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, J. Membr. Sci. 462 (2014) 119-
130. https://doi.org/10.1016/j.memsci.2014.03.046
[124] P. Salehian, W.F. Yong, T.-S. Chung, Development of high performance carboxylated PIM-
1/P84 blend membranes for pervaporation dehydration of isopropanol and CO2/CH4 separation, J.
Membr. Sci. 518 (2016) 110-119. https://doi.org/10.1016/j.memsci.2016.06.027
[125] I. Hossain, A. Husna, D. Kim, T.-H. Kim, (PIM-co-Ellagic Acid)-based copolymer
membranes for high performance CO2 separation, Membr. J. 30 (2020) 420-432.
https://doi.org/10.14579/membrane_journal.2020.30.6.420
[126] I.I. Ponomarev, K.M. Skupov, K.A. Lyssenko, I.V. Blagodatskikh, A.V. Muranov, Y.A.
Volkova, D.Y. Razorenov, I.I. Ponomarev, L.E. Starannikova, D.A. Bezgin, A.Y. Alentiev, Y.P.
Yampolskii, New PIM-1 copolymers containing 2,3,6,7-anthracenetetrayl moiety and their use as
gas separation membranes, Mendeleev Commun. 30 (2020) 734-737.
https://doi.org/10.1016/j.mencom.2020.11.015
[127] G. Bengtson, S. Neumann, V. Filiz, Membranes of polymers of intrinsic microporosity
(PIM-1) modified by poly(ethylene glycol), Membranes 7 (2017) 28.
https://doi.org/10.3390/membranes7020028
[128] X. Mei Wu, Q. Gen Zhang, P. Ju Lin, Y. Qu, A. Mei Zhu, Q. Lin Liu, Towards enhanced
CO2 selectivity of the PIM-1 membrane by blending with polyethylene glycol, J. Membr. Sci. 493
(2015) 147-155. https://doi.org/10.1016/j.memsci.2015.05.077
[129] R. Lin, B.V. Hernandez, L. Ge, Z. Zhu, Metal organic framework based mixed matrix
membranes: An overview on filler/polymer interfaces, J. Mater. Chem. A 6 (2018) 293-312.
https://doi.org/10.1039/C7TA07294E
[130] Y. Wang, X. Wang, J. Guan, L. Yang, Y. Ren, N. Nasir, H. Wu, Z. Chen, Z. Jiang, 110th
Anniversary: Mixed matrix membranes with fillers of intrinsic nanopores for gas separation, Ind.
Eng. Chem. Res. 58 (2019) 7706-7724. https://doi.org/10.1021/acs.iecr.9b01568
[131] S.H. Goh, H.S. Lau, W.F. Yong, Metal-organic frameworks (MOFs)-based mixed matrix
membranes (MMMs) for gas separation: A review on advanced materials in harsh environmental
applications, Small 18 (2022) e2107536. https://doi.org/10.1002/smll.202107536
[132] N. Tien-Binh, D. Rodrigue, S. Kaliaguine, In-situ cross interface linking of PIM-1 polymer
and UiO-66-NH2 for outstanding gas separation and physical aging control, J. Membr. Sci. 548
(2018) 429-438. https://doi.org/10.1016/j.memsci.2017.11.054
[133] B. Zhu, Y. Yang, L. Guo, K. Wang, Y. Lu, X. He, S. Zhang, L. Shao, Ultrapermeable gel
membranes enabling superior carbon capture, Angew. Chem. Int. Ed. 63 (2024) e202315607.
https://doi.org/10.1002/anie.202315607
[134] S. Chen, D. Zhao, Y. Feng, H. Liu, S. Li, Y. Qiu, J. Ren, The preparation and characterization
of gel-mixed matrix membranes (g-MMMs) with high CO2 permeability and stability performance,
J. Membr. Sci. 652 (2022) 120471. https://doi.org/10.1016/j.memsci.2022.120471

132
[135] K. Zhang, X. Luo, S. Li, X. Tian, Q. Wang, C. Liu, Y. Tang, X. Feng, R. Zhang, S. Yin, S.
Wang, ZIF-8 gel/PIM-1 mixed matrix membranes for enhanced H2/CH4 separations, Chem. Eng.
J. 484 (2024) 149489. https://doi.org/10.1016/j.cej.2024.149489
[136] L. Yang, Z. Tian, X. Zhang, X. Wu, Y. Wu, Y. Wang, D. Peng, S. Wang, H. Wu, Z. Jiang,
Enhanced CO2 selectivities by incorporating CO2-philic PEG-POSS into polymers of intrinsic
microporosity membrane, J. Membr. Sci. 543 (2017) 69-78.
https://doi.org/10.1016/j.memsci.2017.08.050
[137] J.M. Luque-Alled, A.W. Ameen, M. Alberto, M. Tamaddondar, A.B. Foster, P.M. Budd, A.
Vijayaraghavan, P. Gorgojo, Gas separation performance of MMMs containing (PIM-1)-
functionalized GO derivatives, J. Membr. Sci. 623 (2021) 118902.
https://doi.org/10.1016/j.memsci.2020.118902
[138] P. Su, S. Chen, L. Chen, W. Li, Constructing polymer/metal-organic framework nanohybrids
to design compatible polymer-filler-polymer membranes for CO2 separation, J. Membr. Sci. 691
(2024) 122246. https://doi.org/10.1016/j.memsci.2023.122246
[139] N. Sakaguchi, M. Tanaka, M. Yamato, H. Kawakami, Superhigh CO2-permeable mixed
matrix membranes composed of a polymer of intrinsic microporosity (PIM-1) and surface-
modified silica nanoparticles, ACS Appl. Polym. Mater. 1 (2019) 2516-2524.
https://doi.org/10.1021/acsapm.9b00624
[140] Y. Kinoshita, K. Wakimoto, A.H. Gibbons, A.P. Isfahani, H. Kusuda, E. Sivaniah, B. Ghalei,
Enhanced PIM-1 membrane gas separation selectivity through efficient dispersion of
functionalized POSS fillers, J. Membr. Sci. 539 (2017) 178-186.
https://doi.org/10.1016/j.memsci.2017.05.072
[141] F. Almansour, M. Alberto, A.B. Foster, S. Mohsenpour, P.M. Budd, P. Gorgojo, Thin film
nanocomposite membranes of superglassy PIM-1 and amine-functionalised 2D fillers for gas
separation, J. Mater. Chem. A 10 (2022) 23341-23351. https://doi.org/10.1039/d2ta06339e
[142] Y. Wang, Y. Ren, H. Wu, X. Wu, H. Yang, L. Yang, X. Wang, Y. Wu, Y. Liu, Z. Jiang,
Amino-functionalized ZIF-7 embedded polymers of intrinsic microporosity membrane with
enhanced selectivity for biogas upgrading, J. Membr. Sci. 602 (2020) 117970.
https://doi.org/10.1016/j.memsci.2020.117970
[143] B.K. Voon, H.S. Lau, C.Z. Liang, W.F. Yong, Functionalized two-dimensional g-C3N4
nanosheets in PIM-1 mixed matrix membranes for gas separation, Sep. Purif. Technol. 296 (2022)
121354. https://doi.org/10.1016/j.seppur.2022.121354
[144] G. Yu, X. Zou, L. Sun, B. Liu, Z. Wang, P. Zhang, G. Zhu, Constructing connected paths
between UiO-66 and PIM-1 to improve membrane CO2 separation with crystal-like gas selectivity,
Adv. Mater. 31 (2019) e1806853. https://doi.org/10.1002/adma.201806853
[145] Z. Wang, H. Ren, S. Zhang, F. Zhang, J. Jin, Polymers of intrinsic microporosity/metal–
organic framework hybrid membranes with improved interfacial interaction for high-performance
CO2 separation, J. Mater. Chem. A 5 (2017) 10968-10977. https://doi.org/10.1039/c7ta01773a
[146] B. Zhu, S. He, Y. Yang, S. Li, C.H. Lau, S. Liu, L. Shao, Boosting membrane carbon capture
via multifaceted polyphenol-mediated soldering, Nat. Commun. 14 (2023) 1697.
https://doi.org/10.1038/s41467-023-37479-9
[147] J. Han, L. Bai, H. Jiang, S. Zeng, B. Yang, Y. Bai, X. Zhang, Task-specific Ionic Liquids
tuning ZIF-67/PIM-1 mixed matrix membranes for efficient CO2 separation, Ind. Eng. Chem. Res.
60 (2021) 593-603. https://doi.org/10.1021/acs.iecr.0c04830

133
[148] J. Han, H. Jiang, S. Zeng, Y. Bai, X. Zhang, L. Bai, CO2 separation performance for PIM
based mixed matrix membranes embedded by superbase ionic liquids, J. Mol. Liq. 359 (2022)
119375. https://doi.org/10.1016/j.molliq.2022.119375
[149] X. Wu, Y. Ren, G. Sui, G. Wang, G. Xu, L. Yang, Y. Wu, G. He, N. Nasir, H. Wu, Z. Jiang,
Accelerating CO2 capture of highly permeable polymer through incorporating highly selective
hollow zeolite imidazolate framework, AIChE J. 66 (2019) e16800.
https://doi.org/10.1002/aic.16800
[150] C. Ye, X. Wu, H. Wu, L. Yang, Y. Ren, Y. Wu, Y. Liu, Z. Guo, R. Zhao, Z. Jiang,
Incorporating nano-sized ZIF-67 to enhance selectivity of polymers of intrinsic microporosity
membranes for biogas upgrading, Chem. Eng. Sci. 216 (2020) 115497.
https://doi.org/10.1016/j.ces.2020.115497
[151] A.F. Bushell, M.P. Attfield, C.R. Mason, P.M. Budd, Y. Yampolskii, L. Starannikova, A.
Rebrov, F. Bazzarelli, P. Bernardo, J. Carolus Jansen, M. Lanč, K. Friess, V. Shantarovich, V.
Gustov, V. Isaeva, Gas permeation parameters of mixed matrix membranes based on the polymer
of intrinsic microporosity PIM-1 and the zeolitic imidazolate framework ZIF-8, J. Membr. Sci.
427 (2013) 48-62. https://doi.org/10.1016/j.memsci.2012.09.035
[152] B. Ghalei, K. Sakurai, Y. Kinoshita, K. Wakimoto, Ali P. Isfahani, Q. Song, K. Doitomi, S.
Furukawa, H. Hirao, H. Kusuda, S. Kitagawa, E. Sivaniah, Enhanced selectivity in mixed matrix
membranes for CO2 capture through efficient dispersion of amine-functionalized MOF
nanoparticles, Nat. Energy 2 (2017) 17086. https://doi.org/10.1038/nenergy.2017.86
[153] M. Khdhayyer, A.F. Bushell, P.M. Budd, M.P. Attfield, D. Jiang, A.D. Burrows, E. Esposito,
P. Bernardo, M. Monteleone, A. Fuoco, G. Clarizia, F. Bazzarelli, A. Gordano, J.C. Jansen, Mixed
matrix membranes based on MIL-101 metal–organic frameworks in polymer of intrinsic
microporosity PIM-1, Sep. Purif. Technol. 212 (2019) 545-554.
https://doi.org/10.1016/j.seppur.2018.11.055
[154] B. Qiu, M. Yu, J.M. Luque-Alled, S. Ding, A.B. Foster, P.M. Budd, X. Fan, P. Gorgojo,
High gas permeability in aged superglassy membranes with nanosized UiO-66-NH2/cPIM-1
network fillers, Angew. Chem. Int. Ed. 63 (2024) e202316356.
https://doi.org/10.1002/anie.202316356
[155] S. He, B. Zhu, X. Jiang, G. Han, S. Li, C.H. Lau, Y. Wu, Y. Zhang, L. Shao, Symbiosis-
inspired de novo synthesis of ultrahigh MOF growth mixed matrix membranes for sustainable
carbon capture, Proc. Natl. Acad. Sci. 119 (2022) e2114964119.
https://doi.org/10.1073/pnas.2114964119
[156] J. Ahn, W.-J. Chung, I. Pinnau, J. Song, N. Du, G.P. Robertson, M.D. Guiver, Gas transport
behavior of mixed-matrix membranes composed of silica nanoparticles in a polymer of intrinsic
microporosity (PIM-1), J. Membr. Sci. 346 (2010) 280-287.
https://doi.org/10.1016/j.memsci.2009.09.047
[157] Q. Song, S. Cao, R.H. Pritchard, H. Qiblawey, E.M. Terentjev, A.K. Cheetham, E. Sivaniah,
Nanofiller-tuned microporous polymer molecular sieves for energy and environmental processes,
J. Mater. Chem. A 4 (2016) 270-279. https://doi.org/10.1039/C5TA09060A
[158] M.G. De Angelis, R. Gaddoni, G.C. Sarti, Gas solubility, diffusivity, permeability, and
selectivity in mixed matrix membranes based on PIM-1 and fumed silica, Ind. Eng. Chem. Res.
52 (2013) 10506-10520. https://doi.org/10.1021/ie303571h
[159] C.R. Mason, M.G. Buonomenna, G. Golemme, P.M. Budd, F. Galiano, A. Figoli, K. Friess,
V. Hynek, New organophilic mixed matrix membranes derived from a polymer of intrinsic

134
microporosity and silicalite-1, Polymer 54 (2013) 2222-2230.
https://doi.org/10.1016/j.polymer.2013.02.032
[160] S.-W. Kuo, F.-C. Chang, POSS related polymer nanocomposites, Prog. Polym. Sci. 36 (2011)
1649-1696. https://doi.org/10.1016/j.progpolymsci.2011.05.002
[161] W.F. Yong, K.H.A. Kwek, K.-S. Liao, T.-S. Chung, Suppression of aging and plasticization
in highly permeable polymers, Polymer 77 (2015) 377-386.
https://doi.org/10.1016/j.polymer.2015.09.075
[162] N. Konnertz, Y. Ding, W.J. Harrison, P.M. Budd, A. Schönhals, M. Böhning, Molecular
mobility and gas transport properties of nanocomposites based on PIM-1 and polyhedral
oligomeric phenethyl-silsesquioxanes (POSS), J. Membr. Sci. 529 (2017) 274-285.
https://doi.org/10.1016/j.memsci.2017.02.007
[163] A. Gonciaruk, K. Althumayri, W.J. Harrison, P.M. Budd, F.R. Siperstein, PIM-1/graphene
composite: A combined experimental and molecular simulation study, Microporous Mesoporous
Mater. 209 (2015) 126-134. https://doi.org/10.1016/j.micromeso.2014.07.007
[164] M. Chen, F. Soyekwo, Q. Zhang, C. Hu, A. Zhu, Q. Liu, Graphene oxide nanosheets to
improve permeability and selectivity of PIM-1 membrane for carbon dioxide separation, J. Ind.
Eng. Chem. 63 (2018) 296-302. https://doi.org/10.1016/j.jiec.2018.02.030
[165] M. Alberto, R. Bhavsar, J.M. Luque-Alled, A. Vijayaraghavan, P.M. Budd, P. Gorgojo,
Impeded physical aging in PIM-1 membranes containing graphene-like fillers, J. Membr. Sci. 563
(2018) 513-520. https://doi.org/10.1016/j.memsci.2018.06.026
[166] Z. Tian, S. Wang, Y. Wang, X. Ma, K. Cao, D. Peng, X. Wu, H. Wu, Z. Jiang, Enhanced
gas separation performance of mixed matrix membranes from graphitic carbon nitride nanosheets
and polymers of intrinsic microporosity, J. Membr. Sci. 514 (2016) 15-24.
https://doi.org/10.1016/j.memsci.2016.04.019
[167] A.W. Ameen, J. Ji, M. Tamaddondar, S. Moshenpour, A.B. Foster, X. Fan, P.M. Budd, D.
Mattia, P. Gorgojo, 2D boron nitride nanosheets in PIM-1 membranes for CO2/CH4 separation, J.
Membr. Sci. 636 (2021) 119527. https://doi.org/10.1016/j.memsci.2021.119527
[168] M.M. Khan, V. Filiz, G. Bengtson, S. Shishatskiy, M. Rahman, V. Abetz, Functionalized
carbon nanotubes mixed matrix membranes of polymers of intrinsic microporosity for gas
separation, Nanoscale Res. Lett. 7 (2012) 504. https://doi.org/10.1186/1556-276X-7-504
[169] Q. Li, Z. Zhu, Y. Wang, H. Wang, J. Li, X. Ma, Unprecedented gas separation performance
of ITTB/CNT nanocomposite membranes at low temperature by strong interfacial interaction
enhanced rigidity, J. Membr. Sci. 636 (2021) 119590.
https://doi.org/10.1016/j.memsci.2021.119590
[170] M.M. Khan, V. Filiz, G. Bengtson, S. Shishatskiy, M.M. Rahman, J. Lillepaerg, V. Abetz,
Enhanced gas permeability by fabricating mixed matrix membranes of functionalized multiwalled
carbon nanotubes and polymers of intrinsic microporosity (PIM), J. Membr. Sci. 436 (2013) 109-
120. https://doi.org/10.1016/j.memsci.2013.02.032
[171] K. Golzar, H. Modarress, S. Amjad-Iranagh, Effect of pristine and functionalized single- and
multi-walled carbon nanotubes on CO2 separation of mixed matrix membranes based on polymers
of intrinsic microporosity (PIM-1): A molecular dynamics simulation study, J. Mol. Model. 23
(2017) 266. https://doi.org/10.1007/s00894-017-3436-3
[172] Z. Tian, D. Li, W. Zheng, Q. Chang, Y. Sang, F. Lai, J. Wang, Y. Zhang, T. Liu, M.
Antonietti, Heteroatom-doped noble carbon-tailored mixed matrix membranes with
ultrapermeability for efficient CO2 separation, Mater. Horiz. 10 (2023) 3660-3667.
https://doi.org/10.1039/d3mh00463e

135
[173] K. Wang, D. Chen, J. Tang, Z. Hong, Z. Zhu, Z. Yuan, Z. Lin, Y. Liu, R. Semiat, X. He,
PIM-1-based membranes mediated with CO2-philic MXene nanosheets for superior CO2/N2
separation, Chem. Eng. J. 483 (2024) 149305. https://doi.org/10.1016/j.cej.2024.149305
[174] M. Yahia, Q.N. Phan Le, N. Ismail, M. Essalhi, O. Sundman, A. Rahimpour, M.M. Dal-Cin,
N. Tavajohi, Effect of incorporating different ZIF-8 crystal sizes in the polymer of intrinsic
microporosity, PIM-1, for CO2/CH4 separation, Microporous Mesoporous Mater. 312 (2021)
110761. https://doi.org/10.1016/j.micromeso.2020.110761
[175] X. Wu, W. Liu, H. Wu, X. Zong, L. Yang, Y. Wu, Y. Ren, C. Shi, S. Wang, Z. Jiang,
Nanoporous ZIF-67 embedded polymers of intrinsic microporosity membranes with enhanced gas
separation performance, J. Membr. Sci. 548 (2018) 309-318.
https://doi.org/10.1016/j.memsci.2017.11.038
[176] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga, K.P. Lillerud, A
new zirconium inorganic building brick forming metal organic frameworks with exceptional
stability, J. Am. Chem. Soc. 130 (2008) 13850-13851. https://doi.org/10.1021/ja8057953
[177] M.R. Khdhayyer, E. Esposito, A. Fuoco, M. Monteleone, L. Giorno, J.C. Jansen, M.P.
Attfield, P.M. Budd, Mixed matrix membranes based on UiO-66 MOFs in the polymer of intrinsic
microporosity PIM-1, Sep. Purif. Technol. 173 (2017) 304-313.
https://doi.org/10.1016/j.seppur.2016.09.036
[178] C. Geng, Y. Sun, Z. Zhang, Z. Qiao, C. Zhong, Mitigated aging in a defective metal–organic
framework pillared polymer of an intrinsic porosity hybrid membrane for efficient gas separation,
ACS Sustain. Chem. Eng. 10 (2022) 3643-3650. https://doi.org/10.1021/acssuschemeng.1c08485
[179] Z. Zhou, X. Cao, D. Lv, F. Cheng, Hydrophobic metal–organic framework UiO-66-
(CF3)2/PIM-1 mixed-matrix membranes for stable CO2/N2 separation under high humidity, Sep.
Purif. Technol. 339 (2024) 126666. https://doi.org/10.1016/j.seppur.2024.126666
[180] S. Mandal, S. Natarajan, P. Mani, A. Pankajakshan, Post‐synthetic modification of metal–
organic frameworks toward applications, Adv. Funct. Mater. 31 (2020) 2006291.
https://doi.org/10.1002/adfm.202006291
[181] S.J. Smith, B.P. Ladewig, A.J. Hill, C.H. Lau, M.R. Hill, Post-synthetic Ti exchanged UiO-
66 metal-organic frameworks that deliver exceptional gas permeability in mixed matrix
membranes, Sci. Rep. 5 (2015) 7823. https://doi.org/10.1038/srep07823
[182] J. Gao, Y. Sun, F. Kang, F. Guo, G. He, H. Wang, Z. Yang, C. Ma, X. Jiang, W. Xiao,
Amidoxime modified UiO-66@PIM-1 mixed-matrix membranes to enhance CO2 separation and
anti-aging performance, Membranes 13 (2023) 781. https://doi.org/10.3390/membranes13090781
[183] P.F. Muldoon, S.R. Venna, D.W. Gidley, J.S. Baker, L. Zhu, Z. Tong, F. Xiang, D.P.
Hopkinson, S. Yi, A.K. Sekizkardes, N.L. Rosi, Mixed matrix membranes from a microporous
polymer blend and nanosized metal–organic frameworks with exceptional CO2/N2 separation
performance, ACS Mater. Lett. 2 (2020) 821-828.
https://doi.org/10.1021/acsmaterialslett.0c00156
[184] I.D. Carja, S.R. Tavares, O. Shekhah, A. Ozcan, R. Semino, V.S. Kale, M. Eddaoudi, G.
Maurin, Insights into the enhancement of MOF/polymer adhesion in mixed-matrix membranes via
polymer functionalization, ACS Appl. Mater. Interfaces 13 (2021) 29041-29047.
https://doi.org/10.1021/acsami.1c03859
[185] G. Férey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surblé, I. Margiolaki, A
chromium terephtalate-based solid with unusually large pore volumes and surface area, Science
309 (2005) 2040-2042. https://doi.org/10.1126/science.1116275

136
[186] Y. Wang, J. Yang, Z. Li, Z. Zhang, J. Li, Q. Yang, C. Zhong, Computational study of oxygen
adsorption in metal-organic frameworks with exposed cation sites: Effect of framework metal ions,
RSC Adv. 5 (2015) 33432-33437. https://doi.org/10.1039/c5ra04791a
[187] N. Tien-Binh, H. Vinh-Thang, X.Y. Chen, D. Rodrigue, S. Kaliaguine, Crosslinked MOF-
polymer to enhance gas separation of mixed matrix membranes, J. Membr. Sci. 520 (2016) 941-
950. https://doi.org/10.1016/j.memsci.2016.08.045
[188] Q. Shen, S. Cong, J. Zhu, Y. Zhang, R. He, S. Yi, Y. Zhang, Novel pyrazole-based MOF
synergistic polymer of intrinsic microporosity membranes for high-efficient CO2 capture, J.
Membr. Sci. 664 (2022) 121107. https://doi.org/10.1016/j.memsci.2022.121107
[189] H. Yin, A. Alkaş, Y. Zhang, Y. Zhang, S.G. Telfer, Mixed matrix membranes (MMMs)
using an emerging metal-organic framework (MUF-15) for CO2 separation, J. Membr. Sci. 609
(2020) 118245. https://doi.org/10.1016/j.memsci.2020.118245
[190] A. Fuoco, M.R. Khdhayyer, M.P. Attfield, E. Esposito, J.C. Jansen, P.M. Budd, Synthesis
and transport properties of novel MOF/PIM-1/MOF sandwich membranes for gas separation,
Membranes 7 (2017) 7. https://doi.org/10.3390/membranes7010007
[191] Y. Cheng, S.R. Tavares, C.M. Doherty, Y. Ying, E. Sarnello, G. Maurin, M.R. Hill, T. Li,
D. Zhao, Enhanced polymer crystallinity in mixed-matrix membranes induced by metal–organic
framework nanosheets for efficient CO2 capture, ACS Appl. Mater. Interfaces 10 (2018) 43095-
43103. https://doi.org/10.1021/acsami.8b16386
[192] Y. Pu, Z. Yang, V. Wee, Z. Wu, Z. Jiang, D. Zhao, Amino-functionalized NUS-8 nanosheets
as fillers in PIM-1 mixed matrix membranes for CO2 separations, J. Membr. Sci. 641 (2022)
119912. https://doi.org/10.1016/j.memsci.2021.119912
[193] D. Fan, A. Ozcan, N.A. Ramsahye, G. Maurin, R. Semino, Putting forward NUS-8-
CO2H/PIM-1 as a mixed matrix membrane for CO2 capture, ACS Appl. Mater. Interfaces 14 (2022)
16820-16829. https://doi.org/10.1021/acsami.2c00090
[194] D. Wang, Y. Ying, Y. Zheng, Y. Pu, Z. Yang, D. Zhao, Induced polymer crystallinity in
mixed matrix membranes by metal-organic framework nanosheets for gas separation, J. Membr.
Sci. Lett. 2 (2022) 100017. https://doi.org/10.1016/j.memlet.2022.100017
[195] O.T. Qazvini, R. Babarao, Z.-L. Shi, Y.-B. Zhang, S.G. Telfer, A robust ethane-trapping
metal-organic framework with a high capacity for ethylene purification, J. Am. Chem. Soc. 141
(2019) 5014-5020. https://doi.org/10.1021/jacs.9b00913
[196] N. Prasetya, B.C. Donose, B.P. Ladewig, A new and highly robust light-responsive Azo-
UiO-66 for highly selective and low energy post-combustion CO2 capture and its application in a
mixed matrix membrane for CO2/N2 separation, J. Mater. Chem. A 6 (2018) 16390-16402.
https://doi.org/10.1039/C8TA03553A
[197] Z. Wang, A. Knebel, S. Grosjean, D. Wagner, S. Bräse, C. Wöll, J. Caro, L. Heinke, Tunable
molecular separation by nanoporous membranes, Nat. Commun. 7 (2016) 13872.
https://doi.org/10.1038/ncomms13872
[198] A. Knebel, L. Sundermann, A. Mohmeyer, I. Strauß, S. Friebe, P. Behrens, J. Caro,
Azobenzene guest molecules as light-switchable CO2 valves in an ultrathin UiO-67 membrane,
Chem. Mater. 29 (2017) 3111-3117. https://doi.org/10.1021/acs.chemmater.7b00147
[199] N. Prasetya, B.P. Ladewig, Dynamic photo-switching in light-responsive JUC-62 for CO2
capture, Sci. Rep. 7 (2017) 13355. https://doi.org/10.1038/s41598-017-13536-4
[200] M. Wang, S. Zhou, S. Cao, Z. Wang, S. Liu, S. Wei, Y. Chen, X. Lu, Stimulus-responsive
adsorbent materials for CO2 capture and separation, J. Mater. Chem. A 8 (2020) 10519-10533.
https://doi.org/10.1039/D0TA01863E

137
[201] N. Prasetya, A.A. Teck, B.P. Ladewig, Matrimid-JUC-62 and Matrimid-PCN-250 mixed
matrix membranes displaying light-responsive gas separation and beneficial ageing characteristics
for CO2/N2 separation, Sci. Rep. 8 (2018) 2944. https://doi.org/10.1038/s41598-018-21263-7
[202] N. Prasetya, B.P. Ladewig, New Azo-DMOF-1 MOF as a photoresponsive low-energy CO2
adsorbent and its exceptional CO2/N2 separation performance in mixed matrix membranes, ACS
Appl. Mater. Interfaces 10 (2018) 34291-34301. https://doi.org/10.1021/acsami.8b12261
[203] N. Prasetya, B.P. Ladewig, An insight into the effect of azobenzene functionalities studied
in UiO-66 frameworks for low energy CO2 capture and CO2/N2 membrane separation, J. Mater.
Chem. A 7 (2019) 15164-15172. https://doi.org/10.1039/c9ta02096a
[204] F.P. Kinik, A. Uzun, S. Keskin, Ionic liquid/metal–organic framework composites: From
synthesis to applications, ChemSusChem 10 (2017) 2842-2863.
https://doi.org/10.1002/cssc.201700716
[205] Y. Ban, Z. Li, Y. Li, Y. Peng, H. Jin, W. Jiao, A. Guo, P. Wang, Q. Yang, C. Zhong, W.
Yang, Confinement of ionic liquids in nanocages: Tailoring the molecular sieving properties of
ZIF-8 for membrane-based CO2 capture, Angew. Chem. Int. Ed. 54 (2015) 15483-15487.
https://doi.org/10.1002/anie.201505508
[206] H. Li, L. Tuo, K. Yang, H.-K. Jeong, Y. Dai, G. He, W. Zhao, Simultaneous enhancement
of mechanical properties and CO2 selectivity of ZIF-8 mixed matrix membranes: Interfacial
toughening effect of ionic liquid, J. Membr. Sci. 511 (2016) 130-142.
https://doi.org/10.1016/j.memsci.2016.03.050
[207] J. Lu, X. Zhang, L. Xu, G. Zhang, J. Zheng, Z. Tong, C. Shen, Q. Meng, Preparation of
amino-functional UiO-66/PIMs mixed matrix membranes with [bmim][Tf2N] as regulator for
enhanced gas separation, Membranes 11 (2021) 35. https://doi.org/10.3390/membranes11010035
[208] W. Chen, Z. Zhang, C. Yang, J. Liu, H. Shen, K. Yang, Z. Wang, PIM-based mixed-matrix
membranes containing MOF-801/ionic liquid nanocomposites for enhanced CO2 separation
performance, J. Membr. Sci. 636 (2021) 119581. https://doi.org/10.1016/j.memsci.2021.119581
[209] C. Geng, Y. Sun, Z. Zhang, Z. Qiao, C. Zhong, Mixed matrix metal–organic framework
membranes for efficient CO2/N2 separation under humid conditions, AIChE J. 69 (2023) e18025.
https://doi.org/10.1002/aic.18025
[210] Z. Hu, E.M. Mahdi, Y. Peng, Y. Qian, B. Zhang, N. Yan, D. Yuan, J.-C. Tan, D. Zhao,
Kinetically controlled synthesis of two-dimensional Zr/Hf metal–organic framework nanosheets
via a modulated hydrothermal approach, J. Mater. Chem. A 5 (2017) 8954-8963.
https://doi.org/10.1039/C7TA00413C
[211] Y. Sun, C. Geng, Z. Zhang, Z. Qiao, C. Zhong, Two-dimensional basic cobalt carbonate
supported ZIF-67 composites towards mixed matrix membranes for efficient CO2/N2 separation,
J. Membr. Sci. 661 (2022) 120928. https://doi.org/10.1016/j.memsci.2022.120928
[212] Z. Xu, Z. Fan, C. Shen, Q. Meng, G. Zhang, C. Gao, Porous composite membrane based on
organic substrate for molecular sieving: Current status, opportunities and challenges, Adv. Membr.
2 (2022) 100027. https://doi.org/10.1016/j.advmem.2022.100027
[213] Z. Qu, C. Lai, G. Zhao, A. Knebel, H. Fan, H. Meng, Pore engineering in covalent organic
framework membrane for gas separation, Adv. Membr. 2 (2022) 100037.
https://doi.org/10.1016/j.advmem.2022.100037
[214] X. Wu, Z. Tian, S. Wang, D. Peng, L. Yang, Y. Wu, Q. Xin, H. Wu, Z. Jiang, Mixed matrix
membranes comprising polymers of intrinsic microporosity and covalent organic framework for
gas separation, J. Membr. Sci. 528 (2017) 273-283. https://doi.org/10.1016/j.memsci.2017.01.042

138
[215] G. Dai, Q. Zhang, S. Xiong, L. Deng, Z. Gao, A. Chen, X. Li, C. Pan, J. Tang, G. Yu,
Building interfacial compatible PIM-1-based mixed-matrix membranes with β-ketoenamine-
linked COF fillers for effective CO2/N2 separation, J. Membr. Sci. 676 (2023) 121561.
https://doi.org/10.1016/j.memsci.2023.121561
[216] G. Yu, Y. Li, Z. Wang, T.X. Liu, G. Zhu, X. Zou, Mixed matrix membranes derived from
nanoscale porous organic frameworks for permeable and selective CO2 separation, J. Membr. Sci.
591 (2019) 117343. https://doi.org/10.1016/j.memsci.2019.117343
[217] F. Emamverdi, J. Huang, N.M. Razavi, M.J. Bojdys, A.B. Foster, P.M. Budd, M. Bohning,
A. Schonhals, Molecular mobility and gas transport properties of mixed matrix membranes based
on PIM-1 and a phosphinine containing covalent organic framework, Macromolecules 57 (2024)
1829-1845. https://doi.org/10.1021/acs.macromol.3c02419
[218] P. Kuhn, M. Antonietti, A. Thomas, Porous, covalent triazine-based frameworks prepared
by ionothermal synthesis, Angew. Chem. Int. Ed. 47 (2008) 3450-3453.
https://doi.org/10.1002/anie.200705710
[219] Y. Zhao, K.X. Yao, B. Teng, T. Zhang, Y. Han, A perfluorinated covalent triazine-based
framework for highly selective and water–tolerant CO2 capture, Energy Environ. Sci. 6 (2013)
3684-3692. https://doi.org/10.1039/C3EE42548G
[220] H. Jiang, J. Zhang, T. Huang, J. Xue, Y. Ren, Z. Guo, H. Wang, L. Yang, Y. Yin, Z. Jiang,
M.D. Guiver, Mixed-matrix membranes with covalent triazine framework fillers in polymers of
intrinsic microporosity for CO2 separations, Ind. Eng. Chem. Res. 59 (2019) 5296-5306.
https://doi.org/10.1021/acs.iecr.9b04632
[221] C. Wang, F. Guo, H. Li, J. Xu, J. Hu, H. Liu, Porous organic polymer as fillers for fabrication
of defect-free PIM-1 based mixed matrix membranes with facilitating CO2-transfer chain, J.
Membr. Sci. 564 (2018) 115-122. https://doi.org/10.1016/j.memsci.2018.07.018
[222] T. Ben, H. Ren, S. Ma, D. Cao, J. Lan, X. Jing, W. Wang, J. Xu, F. Deng, J.M. Simmons, S.
Qiu, G. Zhu, Targeted synthesis of a porous aromatic framework with high stability and
exceptionally high surface area, Angew. Chem. Int. Ed. 48 (2009) 9457-9460.
https://doi.org/10.1002/anie.200904637
[223] Y. Li, T. Ben, B. Zhang, Y. Fu, S. Qiu, Ultrahigh gas storage both at low and high pressures
in KOH-activated carbonized porous aromatic frameworks, Sci. Rep. 3 (2013) 2420.
https://doi.org/10.1038/srep02420
[224] C.H. Lau, P.T. Nguyen, M.R. Hill, A.W. Thornton, K. Konstas, C.M. Doherty, R.J. Mulder,
L. Bourgeois, A.C.Y. Liu, D.J. Sprouster, J.P. Sullivan, T.J. Bastow, A.J. Hill, D.L. Gin, R.D.
Noble, Ending aging in super glassy polymer membranes, Angew. Chem. Int. Ed. 53 (2014) 5322-
5326. https://doi.org/10.1002/anie.201402234
[225] C.H. Lau, K. Konstas, C.M. Doherty, S. Kanehashi, B. Ozcelik, S.E. Kentish, A.J. Hill, M.R.
Hill, Tailoring physical aging in super glassy polymers with functionalized porous aromatic
frameworks for CO2 capture, Chem. Mater. 27 (2015) 4756-4762.
https://doi.org/10.1021/acs.chemmater.5b01537
[226] R. Hou, B.S. Ghanem, S.J.D. Smith, C.M. Doherty, C. Setter, H. Wang, I. Pinnau, M.R. Hill,
Highly permeable and selective mixed-matrix membranes for hydrogen separation containing
PAF-1, J. Mater. Chem. A 8 (2020) 14713-14720. https://doi.org/10.1039/D0TA05071G
[227] R. Hou, S.J.D. Smith, K. Konstas, C.M. Doherty, C.D. Easton, J. Park, H. Yoon, H. Wang,
B.D. Freeman, M.R. Hill, Synergistically improved PIM-1 membrane gas separation performance
by PAF-1 incorporation and UV irradiation, J. Mater. Chem. A 10 (2022) 10107-10119.
https://doi.org/10.1039/d2ta00138a

139
[228] L. Tan, B. Tan, Hypercrosslinked porous polymer materials: Design, synthesis, and
applications, Chem. Soc. Rev. 46 (2017) 3322-3356. https://doi.org/10.1039/C6CS00851H
[229] R.T. Woodward, L.A. Stevens, R. Dawson, M. Vijayaraghavan, T. Hasell, I.P. Silverwood,
A.V. Ewing, T. Ratvijitvech, J.D. Exley, S.Y. Chong, F. Blanc, D.J. Adams, S.G. Kazarian, C.E.
Snape, T.C. Drage, A.I. Cooper, Swellable, water- and acid-tolerant polymer sponges for
chemoselective carbon dioxide capture, J. Am. Chem. Soc. 136 (2014) 9028-9035.
https://doi.org/10.1021/ja5031968
[230] R.S. Bhavsar, T. Mitra, D.J. Adams, A.I. Cooper, P.M. Budd, Ultrahigh-permeance PIM-1
based thin film nanocomposite membranes on PAN supports for CO2 separation, J. Membr. Sci.
564 (2018) 878-886. https://doi.org/10.1016/j.memsci.2018.07.089
[231] R. Hou, S.J.D. Smith, C.D. Wood, R.J. Mulder, C.H. Lau, H. Wang, M.R. Hill, Solvation
effects on the permeation and aging performance of PIM-1-based MMMs for gas separation, ACS
Appl. Mater. Interfaces 11 (2019) 6502-6511. https://doi.org/10.1021/acsami.8b19207
[232] M. Tamaddondar, A.B. Foster, J.M. Luque-Alled, K.J. Msayib, M. Carta, S. Sorribas, P.
Gorgojo, N.B. McKeown, P.M. Budd, Intrinsically microporous polymer nanosheets for high-
performance gas separation membranes, Macromol. Rapid Commun. 41 (2020) e1900572.
https://doi.org/10.1002/marc.201900572
[233] M. Tamaddondar, A.B. Foster, M. Carta, P. Gorgojo, N.B. McKeown, P.M. Budd,
Mitigation of physical aging with mixed matrix membranes based on cross-linked PIM-1 fillers
and PIM-1, ACS Appl. Mater. Interfaces 12 (2020) 46756-46766.
https://doi.org/10.1021/acsami.0c13838
[234] J. Liu, Y. Xiao, K.-S. Liao, T.-S. Chung, Highly permeable and aging resistant 3D
architecture from polymers of intrinsic microporosity incorporated with beta-cyclodextrin, J.
Membr. Sci. 523 (2017) 92-102. https://doi.org/10.1016/j.memsci.2016.10.001
[235] J. Wu, J. Liu, T.S. Chung, Structural tuning of polymers of intrinsic microporosity via the
copolymerization with macrocyclic 4 ‐ tert ‐ butylcalix[4]arene for enhanced gas separation
performance, Adv. Sustainable Syst. 2 (2018) 1800084. https://doi.org/10.1002/adsu.201800044
[236] 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. https://doi.org/10.1021/ma202667y
[237] Q. Song, S. Cao, R.H. Pritchard, B. Ghalei, S.A. Al-Muhtaseb, E.M. Terentjev, A.K.
Cheetham, E. Sivaniah, Controlled thermal oxidative crosslinking of polymers of intrinsic
microporosity towards tunable molecular sieve membranes, Nat. Commun. 5 (2014) 1-12.
https://doi.org/10.1038/ncomms5813
[238] S. He, X. Jiang, S. Li, F. Ran, J. Long, L. Shao, Intermediate thermal manipulation of
polymers of intrinsic microporous (PIMs) membranes for gas separations, AIChE J. 66 (2020)
e16543. https://doi.org/10.1002/aic.16543
[239] T. Zhang, L. Deng, P. Li, Decarboxylation cross-linking of triptycene-based Tröger’s Base
polymers for gas separation, Ind. Eng. Chem. Res. 59 (2020) 18640-18648.
https://doi.org/10.1021/acs.iecr.0c03740
[240] T.O. McDonald, R. Akhtar, C.H. Lau, T. Ratvijitvech, G. Cheng, R. Clowes, D.J. Adams,
T. Hasell, A.I. Cooper, Using intermolecular interactions to crosslink PIM-1 and modify its gas
sorption properties, J. Mater. Chem. A 3 (2015) 4855-4864. https://doi.org/10.1039/C4TA06070A
[241] L. Chen, P. Su, J. Liu, S. Chen, J. Huang, X. Huang, H. Zhang, B. Ye, W. Yang, W. Li, Post
‐ synthesis amination of polymer of intrinsic microporosity membranes for CO2 separation,
AIChE J. 69 (2023) e18050. https://doi.org/10.1002/aic.18050
140
[242] H. Sun, S. Zhao, Y. Niu, K. Wang, Z. Xu, B. Wei, P. Li, Y. Hou, Facile surface amination
strategy of PIM-1 based membranes for efficient CO2 capture, Sep. Purif. Technol. 331 (2024)
125643. https://doi.org/10.1016/j.seppur.2023.125643
[243] R.A. Hayes, Polyimide gas separation membranes, US patent 4717393, 1988.
[244] F.Y. Li, Y. Xiao, Y.K. Ong, T.-S. Chung, UV-rearranged PIM-1 polymeric membranes for
advanced hydrogen purification and production, Adv. Energy Mater. 2 (2012) 1456-1466.
https://doi.org/10.1002/aenm.201200296
[245] Q. Song, S. Cao, P. Zavala-Rivera, L. Ping Lu, W. Li, Y. Ji, S.A. Al-Muhtaseb, A.K.
Cheetham, E. Sivaniah, Photo-oxidative enhancement of polymeric molecular sieve membranes,
Nat. Commun. 4 (2013) 1918. https://doi.org/10.1038/ncomms2942
[246] P. Ray, D. Gidley, J.V. Badding, A.D. Lueking, UV and chemical modifications of polymer
of intrinsic microporosity 1 to develop vibrational spectroscopic probes of surface chemistry and
porosity, Microporous Mesoporous Mater. 277 (2019) 29-35.
https://doi.org/10.1016/j.micromeso.2018.09.013
[247] F.Y. Li, T.-S. Chung, Physical aging, high temperature and water vapor permeation studies
of UV-rearranged PIM-1 membranes for advanced hydrogen purification and production, Int. J.
Hydrog. Energy 38 (2013) 9786-9793. https://doi.org/10.1016/j.ijhydene.2013.05.056
[248] C.A. Scholes, S. Kanehashi, Polymer of intrinsic microporosity (PIM-1) membranes treated
with supercritical CO2, Membranes 9 (2019) 41. https://doi.org/10.3390/membranes9030041
[249] L.E. Starannikova, A.Y. Alentiev, R.Y. Nikiforov, I.I. Ponomarev, I.V. Blagodatskikh, A.Y.
Nikolaev, V.P. Shantarovich, Y.P. Yampolskii, Effects of different treatments of films of PIM-1
on its gas permeation parameters and free volume, Polymer 212 (2021) 123271.
https://doi.org/10.1016/j.polymer.2020.123271
[250] K. Nagai, A. Higuchi, T. Nakagawa, Gas permeability and stability of poly(1-trimethylsilyl-
1-propyne-co-1-phenyl-1-propyne) membranes, J. Polym. Sci. Part B: Polym. Phys. 33 (1995)
289-298. http://dx.doi.org/10.1002/polb.1995.090330214
[251] A. Brunetti, M. Cersosimo, G. Dong, K.T. Woo, J. Lee, J.S. Kim, Y.M. Lee, E. Drioli, G.
Barbieri, In situ restoring of aged thermally rearranged gas separation membranes, J. Membr. Sci.
520 (2016) 671-678. https://doi.org/10.1016/j.memsci.2016.07.030
[252] P. Bernardo, F. Bazzarelli, F. Tasselli, G. Clarizia, C.R. Mason, L. Maynard-Atem, P.M.
Budd, M. Lanč, K. Pilnáček, O. Vopička, K. Friess, D. Fritsch, Y.P. Yampolskii, V. Shantarovich,
J.C. Jansen, Effect of physical aging on the gas transport and sorption in PIM-1 membranes,
Polymer 113 (2017) 283-294. https://doi.org/10.1016/j.polymer.2016.10.040
[253] X. Chen, L. Wu, H. Yang, Y. Qin, X. Ma, N. Li, Tailoring the microporosity of polymers of
intrinsic microporosity for advanced gas separation by atomic layer deposition, Angew. Chem. Int.
Ed. 60 (2021) 17875-17880. https://doi.org/10.1002/anie.202016901
[254] X. Niu, G. Dong, D. Li, Y. Zhang, Y. Zhang, Atomic layer deposition modified PIM-1
membranes for improved CO2 separation: A comparative study on the microstructure-performance
relationships, J. Membr. Sci. 664 (2022) 121103. https://doi.org/10.1016/j.memsci.2022.121103
[255] L. Hao, K.-S. Liao, T.-S. Chung, Photo-oxidative PIM-1 based mixed matrix membranes
with superior gas separation performance, J. Mater. Chem. A 3 (2015) 17273-17281.
https://doi.org/10.1039/c5ta03776j
[256] X. Huang, L. Chen, S. Chen, P. Su, W. Li, Reconstruction of mixed matrix membranes by
in situ vapor aminolysis for CO2/N2 and CH4/N2 separations, J. Membr. Sci. 685 (2023) 121984.
https://doi.org/10.1016/j.memsci.2023.121984

141
[257] Y. Huang, D.R. Paul, Physical aging of thin glassy polymer films monitored by gas
permeability, Polymer 45 (2004) 8377-8393. https://doi.org/10.1016/j.polymer.2004.10.019
[258] M.M. Merrick, R. Sujanani, B.D. Freeman, Glassy polymers: Historical findings, membrane
applications, and unresolved questions regarding physical aging, Polymer 211 (2020) 123176.
https://doi.org/10.1016/j.polymer.2020.123176
[259] R. Swaidan, B. Ghanem, E. Litwiller, I. Pinnau, Physical aging, plasticization and their
effects on gas permeation in “rigid” polymers of intrinsic microporosity, Macromolecules 48 (2015)
6553-6561. https://doi.org/10.1021/acs.macromol.5b01581
[260] S. Mukherjee, D. Sensharma, K.J. Chen, M.J. Zaworotko, Crystal engineering of porous
coordination networks to enable separation of C2 hydrocarbons, Chem. Commun. 56 (2020)
10419-10441. https://doi.org/10.1039/d0cc04645k
[261] T. Numpilai, C.K. Cheng, J. Limtrakul, T. Witoon, Recent advances in light olefins
production from catalytic hydrogenation of carbon dioxide, Process Saf. Environ. Prot. 151 (2021)
401-427. https://doi.org/10.1016/j.psep.2021.05.025
[262] X. Wang, Y. Wang, B. Robinson, Q. Wang, J. Hu, Ethane oxidative dehydrogenation by
CO2 over stable CsRu/CeO2 catalyst, J. Catal. 413 (2022) 138-149.
https://doi.org/10.1016/j.jcat.2022.06.021
[263] G. Chen, X. Chen, Y. Pan, Y. Ji, G. Liu, W. Jin, M-gallate MOF/6FDA-polyimide mixed-
matrix membranes for C2H4/C2H6 separation, J. Membr. Sci. 620 (2021) 118852.
https://doi.org/10.1016/j.memsci.2020.118852
[264] I. Amghizar, L.A. Vandewalle, K.M. Van Geem, G.B. Marin, New trends in olefin
production, Engineering 3 (2017) 171-178. https://doi.org/10.1016/j.Eng.2017.02.006
[265] Y. Wang, S.B. Peh, D. Zhao, Alternatives to cryogenic distillation: Advanced porous
materials in adsorptive light olefin/paraffin separations, Small 15 (2019) 1900058.
https://doi.org/10.1002/smll.201900058
[266] D.S. Sholl, P.R. Lively, Seven chemical separations to change the world, Nature 532 (2016)
435–437. https://doi.org/10.1038/532435a
[267] C.Y. Chuah, H. Lee, T.-H. Bae, Recent advances of nanoporous adsorbents for light
hydrocarbon (C1 – C3) separation, Chem. Eng. J. 430 (2022) 132654.
https://doi.org/10.1016/j.cej.2021.132654
[268] C.Y. Chuah, T.-H. Bae, Recent advances in Mixed-matrix membranes for light hydrocarbon
(C1–C3) separation, Membranes 12 (2022) 201. https://doi.org/10.3390/membranes12020201
[269] N. Gholamipour, M. Sadeghi, M. Shafiei, Effect of silica nanoparticles on the performance
of polysulfone membranes for olefin‐paraffin separation, Chem. Eng. Technol. 42 (2019) 2292-
2301. https://doi.org/10.1002/ceat.201800147
[270] M. Najafi, M. Sadeghi, A.A. Shamsabadi, M. Dinari, M. Soroush, Polysulfone membranes
incorporated with reduced graphene oxide nanoparticles for enhanced olefin/paraffin separation,
ChemistrySelect 5 (2020) 3675-3681. https://doi.org/10.1002/slct.202000240
[271] H. An, K.Y. Cho, S. Back, X.H. Do, J.-D. Jeon, H.K. Lee, K.-Y. Baek, J.S. Lee, The
significance of the interfacial interaction in mixed matrix membranes for enhanced
propylene/propane separation performance and plasticization resistance, Sep. Purif. Technol. 261
(2021) 118279. https://doi.org/10.1016/j.seppur.2020.118279
[272] H.R. Amedi, M. Aghajani, Modified zeolitic-midazolate framework 8/poly(ether-block-
amide) mixed-matrix membrane for propylene and propane separation, J. Appl. Polym. Sci. 135
(2018) 46273. https://doi.org/10.1002/app.46273

142
[273] S. Shrestha, P.K. Dutta, Modification of a continuous zeolite membrane grown within porous
polyethersulfone with Ag(I) cations for enhanced propylene/propane gas separation, Microporous
Mesoporous Mater. 279 (2019) 178-185. https://doi.org/10.1016/j.micromeso.2018.12.032
[274] T.H. Lee, B.K. Lee, C. Youn, J.H. Kang, Y.J. Kim, K.I. Kim, Y.R. Ha, Y. Han, H.B. Park,
Interface engineering in MOF/crosslinked polyimide mixed matrix membranes for enhanced
propylene/propane separation performance and plasticization resistance, J. Membr. Sci. 667 (2023)
121182. https://doi.org/10.1016/j.memsci.2022.121182
[275] S. Rico-Martínez, C. Álvarez, A. Hernández, J.A. Miguel, Á.E. Lozano, Mixed matrix
membranes loaded with a porous organic polymer having bipyridine moieties, Membranes 12
(2022) 547. https://doi.org/10.3390/membranes12060547
[276] J.H. Shin, H.J. Yu, J. Park, A.S. Lee, S.S. Hwang, S.-J. Kim, S. Park, K.Y. Cho, W. Won,
J.S. Lee, Fluorine-containing polyimide/polysilsesquioxane carbon molecular sieve membranes
and techno-economic evaluation thereof for C3H6/C3H8 separation, J. Membr. Sci. 598 (2020)
117660. https://doi.org/10.1016/j.memsci.2019.117660
[277] S. Kunjattu H, U.K. Kharul, PPO-ZIF MMMs possessing metal-polymer interactions for
propane/propylene separation, J. Membr. Sci. 668 (2023) 121208.
https://doi.org/10.1016/j.memsci.2022.121208
[278] W. Ying, X. Peng, Graphene oxide nanoslit-confined AgBF4/ionic liquid for efficiently
separating olefin from paraffin, Nanotechnology 31 (2020) 085703. https://doi.org/10.1088/1361-
6528/ab53af
[279] M. Doosti, R. Abedini, Polyethyleneglycol-modified cellulose acetate membrane for
efficient olefin/paraffin separation, Energy Fuels 36 (2022) 10082-10095.
https://doi.org/10.1021/acs.energyfuels.2c01768
[280] H.-X. Sun, B.-B. Yuan, P. Li, T. Wang, Y.-Y. Xu, Preparation of nanoporous graphene and
the application of its nanocomposite membrane in propylene/propane separation, Funct. Mater.
Lett. 8 (2015) 1550019. https://doi.org/10.1142/s1793604715500198
[281] B. Yuan, H. Sun, T. Wang, Y. Xu, P. Li, Y. Kong, Q.J. Niu, Propylene/propane permeation
properties of ethyl cellulose (EC) mixed matrix membranes fabricated by incorporation of
nanoporous graphene nanosheets, Sci. Rep. 6 (2016) 28509. https://doi.org/10.1038/srep28509
[282] G. Deng, J. Luo, X. Liu, T. Hu, Y. Wang, X. Zong, S. Xue, Fabrication of analogous mixed
matrix membranes via partially in-situ generation of rigid porous moieties without interfacial
defects, J. Membr. Sci. 644 (2022) 120164. https://doi.org/10.1016/j.memsci.2021.120164
[283] S.Y. Kim, Y. Cho, S.W. Kang, Correlation between functional group and formation of
nanoparticles in PEBAX/Ag salt/Al salt complexes for olefin separation, Polymers 12 (2020) 667.
https://doi.org/10.3390/polym12030667
[284] S.Y. Kim, Y. Cho, S.W. Kang, Preparation and characterization of PEBAX-
5513/AgBF4/BMIMBF4 membranes for olefin/paraffin separation, Polymers 12 (2020) 1550.
https://doi.org/10.3390/polym12071550
[285] M. Monteleone, G. Barbieri, P. Bernardo, A. Borgogno, A. Brunetti, G. Clarizia, E. Esposito,
J.C. Jansen, F. Tasselli, E. Tocci, I.N. Odeh, J.R. Johnson, K.K. Kopeć, L. Giorno, Strategies to
stabilize silver salt in composite Pebax2533©/Ag(NH3)2OH and Pebax©2533 [Ag(15-crown-5-
ether)] membranes for enhanced ethylene/ethane separation, Sep. Sci. Technol. 58 (2022) 1190-
1201. https://doi.org/10.1080/01496395.2022.2130077
[286] M.A. Abdulhamid, G. Genduso, Y. Wang, X. Ma, I. Pinnau, Plasticization-resistant
carboxyl-functionalized 6FDA-polyimide of intrinsic microporosity (PIM–PI) for membrane-

143
based gas separation, Ind. Eng. Chem. Res. 59 (2019) 5247-5256.
https://doi.org/10.1021/acs.iecr.9b04994
[287] X. Ma, R. Swaidan, Y. Belmabkhout, Y. Zhu, E. Litwiller, M. Jouiad, I. Pinnau, Y. Han,
Synthesis and gas transport properties of hydroxyl-functionalized polyimides with intrinsic
microporosity, Macromolecules 45 (2012) 3841-3849. https://doi.org/10.1021/ma300549m
[288] O. Salinas, X. Ma, E. Litwiller, I. Pinnau, High-performance carbon molecular sieve
membranes for ethylene/ethane separation derived from an intrinsically microporous polyimide,
J. Membr. Sci. 500 (2016) 115-123. https://doi.org/10.1016/j.memsci.2015.11.013
[289] O. Salinas, X. Ma, Y. Wang, Y. Han, I. Pinnau, Carbon molecular sieve membrane from a
microporous spirobisindane-based polyimide precursor with enhanced ethylene/ethane mixed-gas
selectivity, RSC Adv. 7 (2017) 3265-3272. https://doi.org/10.1039/c6ra24699k
[290] K.M. Steel, W.J. Koros, Investigation of porosity of carbon materials and related effects on
gas separation properties, Carbon 41 (2003) 253-266. https://doi.org/10.1016/S0008-
6223(02)00309-3
[291] P.S. Tin, T.S. Chung, A.J. Hill, Advanced fabrication of carbon molecular sieve membranes
by nonsolvent pretreatment of precursor polymers, Ind. Eng. Chem. Res. 43 (2004) 6476-6483.
https://doi.org/10.1021/ie049606c
[292] M. Rungta, L. Xu, W.J. Koros, Carbon molecular sieve dense film membranes derived from
Matrimid® for ethylene/ethane separation, Carbon 50 (2012) 1488-1502.
https://doi.org/10.1016/j.carbon.2011.11.019
[293] A.B. Fuertes, I. Menendez, Separation of hydrocarbon gas mixtures using phenolic resin-
based carbon membranes, Sep. Purif. Technol. 28 (2002) 29-41. https://doi.org/10.1016/S1383-
5866(02)00006-0
[294] M. Guo, M. Kanezashi, Recent progress in a membrane-based technique for
propylene/propane separation, Membranes 11 (2021) 310.
https://doi.org/10.3390/membranes11050310
[295] S.N. Ringelberg, A. Meetsma, B. Hessen, J.H. Teuben, Thiophene C−H activation as a
chain-transfer mechanism in ethylene polymerization: catalytic formation of thienyl-capped
polyethylene, J. Am. Chem. Soc. 121 (1999) 6082-6083. https://doi.org/10.1021/ja984288s
[296] O. Salinas, X. Ma, E. Litwiller, I. Pinnau, Ethylene/ethane permeation, diffusion and gas
sorption properties of carbon molecular sieve membranes derived from the prototype ladder
polymer of intrinsic microporosity (PIM-1), J. Membr. Sci. 504 (2016) 133-140.
https://doi.org/10.1016/j.memsci.2015.12.052
[297] K.-S. Liao, J.-Y. Lai, T.-S. Chung, Metal ion modified PIM-1 and its application for
propylene/propane separation, J. Membr. Sci. 515 (2016) 36-44.
https://doi.org/10.1016/j.memsci.2016.05.032
[298] K.-S. Liao, S. Japip, J.-Y. Lai, T.-S. Chung, Boron-embedded hydrolyzed PIM-1 carbon
membranes for synergistic ethylene/ethane purification, J. Membr. Sci. 534 (2017) 92-99.
https://doi.org/10.1016/j.memsci.2017.04.017
[299] K. Hazazi, Y. Wang, B. Ghanem, X. Hu, T. Puspasari, C. Chen, Y. Han, I. Pinnau, Precise
molecular sieving of ethylene from ethane using triptycene-derived submicroporous carbon
membranes, Nature Materials 22 (2023) 1218-1226. https://doi.org/10.1038/s41563-023-01629-7
[300] S. Jiang, L. Guo, L. Chen, C. Song, B. Liu, Q. Yang, Z. Zhang, Y. Yang, Q. Ren, Z. Bao, A
strongly hydrophobic ethane-selective metal-organic framework for efficient ethane/ethylene
separation, Chem. Eng. J. 442 (2022) 136152. https://doi.org/10.1016/j.cej.2022.136152

144
[301] Y. Chen, H. Wu, D. Lv, R. Shi, Y. Chen, Q. Xia, Z. Li, Highly adsorptive separation of
ethane/ethylene by an ethane-selective MOF MIL-142A, Ind. Eng. Chem. Res. 57 (2018) 4063-
4069. https://doi.org/10.1021/acs.iecr.7b05260
[302] Y. Chen, Z. Qiao, H. Wu, D. Lv, R. Shi, Q. Xia, J. Zhou, Z. Li, An ethane-trapping MOF
PCN-250 for highly selective adsorption of ethane over ethylene, Chem. Eng. Sci. 175 (2018) 110-
117. https://doi.org/10.1016/j.ces.2017.09.032
[303] Y. Sun, Q. Yu, C. Geng, G.-R. Zhang, Z. Zhang, Z. Qiao, C. Zhong, Ethane-selective
permeation in mixed matrix membranes containing fluorinated carboxylic acid functionalized
metal-organic frameworks, Chem. Eng. J. 479 (2024) 147656.
https://doi.org/10.1016/j.cej.2023.147656
[304] A. Noonikara-Poyil, H. Cui, B. Wang, Y. Shi, B. Chen, H.V.R. Dias, Remarkably selective
propylene-propane separation using a copper scorpionate, Small 19 (2023) e2206984.
https://doi.org/10.1002/smll.202206984
[305] Y. Ren, B. Chong, W. Xu, Z. Zhang, L. Liu, Y. Wu, Y. Liu, H. Jiang, X. Liang, H. Wu, H.
Zhang, B. Ye, C. Zhong, G. He, Z. Jiang, Coordination-driven structure reconstruction in polymer
of intrinsic microporosity membranes for efficient propylene/propane separation, Innovation 3
(2022) 100334. https://doi.org/10.1016/j.xinn.2022.100334
[306] J. Liu, Y. Xiao, T.-S. Chung, Flexible thermally treated 3D PIM-CD molecular sieve
membranes exceeding the upper bound line for propylene/propane separation, J. Mater. Chem. A
5 (2017) 4583-4595. https://doi.org/10.1039/c6ta09751k
[307] R.J. Swaidan, X. Ma, I. Pinnau, Spirobisindane-based polyimide as efficient precursor of
thermally-rearranged and carbon molecular sieve membranes for enhanced propylene/propane
separation, J. Membr. Sci. 520 (2016) 983-989. https://doi.org/10.1016/j.memsci.2016.08.057
[308] A. Yerzhankyzy, B.S. Ghanem, Y. Wang, N. Alaslai, I. Pinnau, Gas separation performance
and mechanical properties of thermally-rearranged polybenzoxazoles derived from an intrinsically
microporous dihydroxyl-functionalized triptycene diamine-based polyimide, J. Membr. Sci. 595
(2020) 117512. https://doi.org/10.1016/j.memsci.2019.117512
[309] Q. Shen, S. Cong, R. He, Z. Wang, Y. Jin, H. Li, X. Cao, J. Wang, B. Van der Bruggen, Y.
Zhang, SIFSIX-3-Zn/PIM-1 mixed matrix membranes with enhanced permeability for
propylene/propane separation, J. Membr. Sci. 588 (2019) 117201.
https://doi.org/10.1016/j.memsci.2019.117201
[310] D. Peng, X. Feng, G. Yang, X. Niu, Z. Liu, Y. Zhang, In-situ growth of silver complex on
ZIF-8 towards mixed matrix membranes for propylene/propane separation, J. Membr. Sci. 668
(2023) 121267. https://doi.org/10.1016/j.memsci.2022.121267
[311] S. Cong, X. Feng, L. Guo, D. Peng, J. Wang, J. Chen, Y. Zhang, X. Shen, G. Yang, Rational
design of mixed matrix membranes modulated by trisilver complex for efficient propylene/propane
separation, Adv. Sci. 10 (2023) e2206858. https://doi.org/10.1002/advs.202206858
[312] X. Ma, R.J. Swaidan, Y. Wang, C.-e. Hsiung, Y. Han, I. Pinnau, Highly compatible
hydroxyl-functionalized microporous polyimide-ZIF-8 mixed matrix membranes for energy
efficient propylene/propane separation, ACS Appl. Nano Mater. 1 (2018) 3541-3547.
https://doi.org/10.1021/acsanm.8b00682
[313] Z. Wang, W. Wang, T. Zeng, D. Ma, P. Zhang, S. Zhao, L. Yang, X. Zou, G. Zhu, Covalent-
linking-enabled superior compatibility of ZIF-8 hybrid membrane for efficient propylene
separation, Adv. Mater. 34 (2022) e2104606. https://doi.org/10.1002/adma.202104606
[314] N.B. McKeown, Polymers of intrinsic microporosity (PIMs), Polymer 202 (2020) 122736.
https://doi.org/10.1016/j.polymer.2020.122736

145
[315] G.L. Dotto, G. McKay, Current scenario and challenges in adsorption for water treatment,
J. Environ. Chem. Eng. 8 (2020) 103988. https://doi.org/10.1016/j.jece.2020.103988
[316] B.P. Chaplin, Critical review of electrochemical advanced oxidation processes for water
treatment applications, Environ. Sci. Process. Impacts 16 (2014) 1182-1203.
https://doi.org/10.1039/c3em00679d
[317] H.O. Tugaoen, S. Garcia-Segura, K. Hristovski, P. Westerhoff, Challenges in photocatalytic
reduction of nitrate as a water treatment technology, Sci. Total Environ. 599-600 (2017) 1524-
1551. https://doi.org/10.1016/j.scitotenv.2017.04.238
[318] M.H. Mohamed Noor, S. Wong, N. Ngadi, I. Mohammed Inuwa, L.A. Opotu, Assessing the
effectiveness of magnetic nanoparticles coagulation/flocculation in water treatment: A systematic
literature review, Int. J. Environ. Sci. Technol. 19 (2021) 6935-6956.
https://doi.org/10.1007/s13762-021-03369-0
[319] T. Peters, Membrane technology for water treatment, Chem. Eng. Technol. 33 (2010) 1233-
1240. https://doi.org/10.1002/ceat.201000139
[320] H. Guo, H. Li, C. Jing, X. Wang, Soluble polymers with intrinsic porosity for efficient
removal of phenolic compounds from water, Microporous Mesoporous Mater. 319 (2021) 111068.
https://doi.org/10.1016/j.micromeso.2021.111068
[321] B. Satilmis, T. Uyar, Amine modified electrospun PIM-1 ultrafine fibers for an efficient
removal of methyl orange from an aqueous system, Appl. Surf. Sci. 453 (2018) 220-229.
https://doi.org/10.1016/j.apsusc.2018.05.069
[322] B. Satilmis, T. Isık, M.M. Demir, T. Uyar, Amidoxime functionalized polymers of intrinsic
microporosity (PIM-1) electrospun ultrafine fibers for rapid removal of uranyl ions from water,
Appl. Surf. Sci. 467-468 (2019) 648-657. https://doi.org/10.1016/j.apsusc.2018.10.210
[323] B. Satilmis, Amidoxime modified polymers of intrinsic microporosity (PIM-1); A versatile
adsorbent for efficient removal of charged dyes; Equilibrium, kinetic and thermodynamic studies,
J. Polym. Environ. 28 (2020) 995-1009. https://doi.org/10.1007/s10924-020-01664-4
[324] Q. Shi, K. Zhang, R. Lu, J. Jiang, Water desalination and biofuel dehydration through a thin
membrane of polymer of intrinsic microporosity: Atomistic simulation study, J. Membr. Sci. 545
(2018) 49-56. https://doi.org/10.1016/j.memsci.2017.09.057
[325] Y. Lan, P. Peng, Preparation of polymer of intrinsic microporosity composite membranes
and their applications for butanol recovery, J. Appl. Polym. Sci. 136 (2019) 46912.
https://doi.org/10.1002/app.46912
[326] J. Contreras-Martínez, S. Mohsenpour, A.W. Ameen, P.M. Budd, C. García-Payo, M.
Khayet, P. Gorgojo, High-flux thin film composite PIM-1 membranes for butanol recovery:
Experimental study and process simulations, ACS Appl. Mater. Interfaces 13 (2021) 42635-42649.
https://doi.org/10.1021/acsami.1c09112
[327] B. Satilmis, T. Uyar, Development of superhydrophobic electrospun fibrous membrane of
polymers of intrinsic microporosity (PIM-2), Eur. Polym. J. 112 (2019) 87-94.
https://doi.org/10.1016/j.eurpolymj.2018.12.029
[328] M.A. Abdulhamid, G. Szekely, Organic solvent nanofiltration membranes based on
polymers of intrinsic microporosity, Curr. Opin. Chem. Eng. 36 (2022) 100804.
https://doi.org/10.1016/j.coche.2022.100804
[329] Z. Wang, X. Luo, J. Zhang, F. Zhang, W. Fang, J. Jin, Polymer membranes for organic
solvent nanofiltration: Recent progress, challenges and perspectives, Adv. Membr. 3 (2023)
100063. https://doi.org/10.1016/j.advmem.2023.100063

146
[330] D. Fritsch, P. Merten, K. Heinrich, M. Lazar, M. Priske, High performance organic solvent
nanofiltration membranes: Development and thorough testing of thin film composite membranes
made of polymers of intrinsic microporosity (PIMs), J. Membr. Sci. 401-402 (2012) 222-231.
https://doi.org/10.1016/j.memsci.2012.02.008
[331] P. Gorgojo, S. Karan, H.C. Wong, M.F. Jimenez‐Solomon, J.T. Cabral, A.G. Livingston,
Ultrathin polymer films with intrinsic microporosity: Anomalous solvent permeation and high flux
membranes, Adv. Funct. Mater. 24 (2014) 4729-4737. https://doi.org/10.1002/adfm.201400400
[332] P. Agarwal, I. Tomlinson, R.E. Hefner, S. Ge, Y. Rao, T. Dikic, Thin film composite
membranes from polymers of intrinsic microporosity using layer-by-layer method, J. Membr. Sci.
572 (2019) 475-479. https://doi.org/10.1016/j.memsci.2018.11.028
[333] P. Agarwal, R.E. Hefner, S. Ge, I. Tomlinson, Y. Rao, T. Dikic, Nanofiltration membranes
from crosslinked Troger's base Polymers of Intrinsic Microporosity (PIMs), J. Membr. Sci. 595
(2020) 117501. https://doi.org/10.1016/j.memsci.2019.117501
[334] C. Jiang, L. Tian, Y. Hou, Q.J. Niu, Nanofiltration membranes with enhanced microporosity
and inner-pore interconnectivity for water treatment: Excellent balance between permeability and
selectivity, J. Membr. Sci. 586 (2019) 192-201. https://doi.org/10.1016/j.memsci.2019.05.075
[335] T.-H. Chen, T. Su, L. Xia, L.-Y. Gong, B.-N. Dong, L.-F. Liu, Corolliform morphology thin
film composite polyamide membrane constructed via Tröger's base PIM for enhancing
nanofiltration separation performances, Desalination 548 (2023) 116234.
https://doi.org/10.1016/j.desal.2022.116234
[336] A. Tang, C. Fang, W. Feng, J. Lu, J. Li, L. Zhu, Engineering novel thin-film composite
membranes with crater-like surface morphology using rigidly-contorted monomer for high flux
nanofiltration, Desalination 509 (2021) 115067. https://doi.org/10.1016/j.desal.2021.115067
[337] J. Dou, S. Han, S. Lin, Z. Qi, F. Huang, X. Feng, Z. Yao, J. Wang, L. Zhang, Tailoring the
selectivity of quasi-PIMs nanofiltration membrane via molecular flexibility of acyl chloride
monomers for desalination from dye effluents, J. Membr. Sci. 671 (2023) 121382.
https://doi.org/10.1016/j.memsci.2023.121382
[338] J. Gao, S. Japip, T.-S. Chung, Organic solvent resistant membranes made from a cross-linked
functionalized polymer with intrinsic microporosity (PIM) containing thioamide groups, Chem.
Eng. J. 353 (2018) 689-698. https://doi.org/10.1016/j.cej.2018.07.156
[339] S. Zhou, Y. Zhao, J. Zheng, S. Zhang, High-performance functionalized polymer of intrinsic
microporosity (PIM) composite membranes with thin and stable interconnected layer for organic
solvent nanofiltration, J. Membr. Sci. 591 (2019) 117347.
https://doi.org/10.1016/j.memsci.2019.117347
[340] H. Tian, J. Luo, X. Liu, X. Zong, S. Xue, Preparation of PIM-1 thin film composite
membranes with enhanced organic solvent resistance via thermal crosslinking, Sep. Purif. Technol.
323 (2023) 124431. https://doi.org/10.1016/j.seppur.2023.124431
[341] J. Li, M. Zhang, W. Feng, L. Zhu, L. Zhang, PIM-1 pore-filled thin film composite
membranes for tunable organic solvent nanofiltration, J. Membr. Sci. 601 (2020) 117951.
https://doi.org/10.1016/j.memsci.2020.117951
[342] Y. Jin, Q. Song, N. Xie, W. Zheng, J. Wang, J. Zhu, Y. Zhang, Amidoxime-functionalized
polymer of intrinsic microporosity (AOPIM-1)-based thin film composite membranes with
ultrahigh permeance for organic solvent nanofiltration, J. Membr. Sci. 632 (2021) 119375.
https://doi.org/10.1016/j.memsci.2021.119375

147
[343] E.K. McGuinness, F. Zhang, Y. Ma, R.P. Lively, M.D. Losego, Vapor phase infiltration of
metal oxides into nanoporous polymers for organic solvent separation membranes, Chem. Mater.
31 (2019) 5509-5518. https://doi.org/10.1021/acs.chemmater.9b01141
[344] C. Li, S. Li, L. Tian, J. Zhang, B. Su, M.Z. Hu, Covalent organic frameworks (COFs)-
incorporated thin film nanocomposite (TFN) membranes for high-flux organic solvent
nanofiltration (OSN), J. Membr. Sci. 572 (2019) 520-531.
https://doi.org/10.1016/j.memsci.2018.11.005
[345] H. Zhou, A. Akram, A.J.C. Semiao, R. Malpass-Evans, C.H. Lau, N.B. McKeown, W. Zhang,
Enhancement of performance and stability of thin-film nanocomposite membranes for organic
solvent nanofiltration using hypercrosslinked polymer additives, J. Membr. Sci. 644 (2022)
120172. https://doi.org/10.1016/j.memsci.2021.120172
[346] S. Li, R. Zhang, Q. Yao, B. Su, L. Han, C. Gao, High flux thin film composite (TFC)
membrane with non-planar rigid twisted structures for organic solvent nanofiltration (OSN), Sep.
Purif. Technol. 286 (2022) 120496. https://doi.org/10.1016/j.seppur.2022.120496
[347] M.A. Abdulhamid, S.-H. Park, Z. Zhou, D.A. Ladner, G. Szekely, Surface engineering of
intrinsically microporous poly(ether-ether-ketone) membranes: From flat to honeycomb structures,
J. Membr. Sci. 621 (2021) 118997. https://doi.org/10.1016/j.memsci.2020.118997
[348] J. Li, W. Feng, M. Zhang, X. Wang, C. Fang, J. Wang, L. Zhang, L. Zhu, Microporous
matrimid/PIM-1 thin film composite membranes with narrow pore size distribution used for
molecular separation in organic solvents, Macromol. Rapid Commun. 44 (2023) 2200826.
https://doi.org/10.1002/marc.202200826
[349] M.-L. Ouinten, A. Szymczyk, A. Ghoufi, Interactions between methanol/toluene binary
mixtures and an organic solvent nanofiltration PIM-1 membrane, J. Mol. Liq. 357 (2022) 119146.
https://doi.org/10.1016/j.molliq.2022.119146
[350] E. Yang, Y. Liang, N. Yanar, M. Kim, H. Park, H. Choi, Intermolecular cross-linked polymer
of intrinsic microporosity-1 (PIM-1)-based thin-film composite hollow fiber membrane for
organic solvent nanofiltration, J. Membr. Sci. 671 (2023) 121370.
https://doi.org/10.1016/j.memsci.2023.121370
[351] E. Yang, M. Kim, Y. Liang, J. Byun, H. Kim, J. Kim, H. Choi, Tailoring the pore size of
intermolecular cross-linked PIMs-based thin-film composite hollow fiber membranes using
different length cross-linkers for organic solvent nanofiltration, Chem. Eng. J. 474 (2023) 145339.
https://doi.org/10.1016/j.cej.2023.145339
[352] S.I. Polianciuc, A.E. Gurzau, B. Kiss, M.G. Stefan, F. Loghin, Antibiotics in the environment:
Causes and consequences, Med. Pharm. Rep. 93 (2020) 231-240. https://doi.org/10.15386/mpr-
1742
[353] M.N. Alnajrani, O.A. Alsager, Removal of antibiotics from water by polymer of intrinsic
microporosity: isotherms, kinetics, thermodynamics, and adsorption mechanism, Sci. Rep. 10
(2020) 794. https://doi.org/10.1038/s41598-020-57616-4
[354] G. Yan, Z. Yang, X. Zhang, H. Li, L. Wang, Z. Li, J. Chen, Y. Wu, Antibacterial
biodegradable nanofibrous membranes by hybrid needleless electrospinning for high-efficiency
particulate matter removal, Chem. Eng. J. 461 (2023) 142137.
https://doi.org/10.1016/j.cej.2023.142137
[355] S. Wang, N.L. Pomerantz, Z. Dai, W. Xie, E.E. Anderson, T. Miller, S.A. Khan, G.N.
Parsons, Polymer of intrinsic microporosity (PIM) based fibrous mat: Combining particle filtration
and rapid catalytic hydrolysis of chemical warfare agent simulants into a highly sorptive,

148
breathable, and mechanically robust fiber matrix, Mater. Today Adv. 8 (2020) 100085.
https://doi.org/10.1016/j.mtadv.2020.100085
[356] S.L. Malone Rubright, L.L. Pearce, J. Peterson, Environmental toxicology of hydrogen
sulfide, Nitric Oxide 71 (2017) 1-13. https://doi.org/10.1016/j.niox.2017.09.011
[357] S. Yi, B. Ghanem, Y. Liu, I. Pinnau, W.J. Koros, Ultraselective glassy polymer membranes
with unprecedented performance for energy-efficient sour gas separation, Sci. Adv. 5 (2019)
eaaw5459. https://doi.org/10.1126/sciadv.aaw5459
[358] X. Liu, L. Dai, Carbon-based metal-free catalysts, Nat. Rev. Mater. 1 (2016) 16064.
https://doi.org/10.1038/natrevmats.2016.64
[359] S.C. Perry, S.M. Gateman, R. Malpass-Evans, N. McKeown, M. Wegener, P. Nazarovs, J.
Mauzeroll, L. Wang, C. Ponce de León, Polymers with intrinsic microporosity (PIMs) for targeted
CO2 reduction to ethylene, Chemosphere 248 (2020) 125993.
https://doi.org/10.1016/j.chemosphere.2020.125993
[360] S. Rat, A. Chavez-Sanchez, M. Jerigová, D. Cruz, M. Antonietti, Acetic anhydride
polymerization as a pathway to functional porous organic polymers and their application in acid–
base catalysis, ACS Appl. Polym. Mater. 3 (2021) 2588-2597.
https://doi.org/10.1021/acsapm.1c00202
[361] E.M. Maya, E. Rangel-Rangel, U. Díaz, M. Iglesias, Efficient cycloaddition of CO2 to
epoxides using novel heterogeneous organocatalysts based on tetramethylguanidine-
functionalized porous polyphenylenes, J. CO2 Util. 25 (2018) 170-179.
https://doi.org/10.1016/j.jcou.2018.04.001
[362] Y. Pan, X. Zhai, J. Yin, T. Zhang, L. Ma, Y. Zhou, Y. Zhang, J. Meng, Hierarchical porous
and zinc-ion crosslinked PIM-1 nanocomposite as a CO2 cycloaddition catalyst with high
efficiency, ChemSusChem 12 (2019) 2231-2239. https://doi.org/10.1002/cssc.201803066
[363] A.R. Antonangelo, C. Grazia Bezzu, S.S. Mughal, T. Malewschik, N.B. McKeown, S.
Nakagaki, A porphyrin-based microporous network polymer that acts as an efficient catalyst for
cyclooctene and cyclohexane oxidation under mild conditions, Catal. Commun. 99 (2017) 100-
104. https://doi.org/10.1016/j.catcom.2017.05.024
[364] K. Halder, G. Bengtson, V. Filiz, V. Abetz, Catalytically active (Pd) nanoparticles supported
by electrospun PIM-1: Influence of the sorption capacity of the polymer tested in the reduction of
some aromatic nitro compounds, Appl. Catal. A: Gen. 555 (2018) 178-188.
https://doi.org/10.1016/j.apcata.2018.02.004
[365] Q. Tang, J. Gong, Q. Zhao, Efficient organic pollutant degradation under visible-light using
functional polymers of intrinsic microporosity, Catal. Sci. Technol. 9 (2019) 5383-5393.
https://doi.org/10.1039/C9CY01338E
[366] H. Xu, X. Li, H. Hao, X. Dong, W. Sheng, X. Lang, Designing fluorene-based conjugated
microporous polymers for blue light-driven photocatalytic selective oxidation of amines with
oxygen, Appl. Catal. B: Environ. 285 (2021) 119796.
https://doi.org/10.1016/j.apcatb.2020.119796
[367] S. Li, W. Zhang, S. Yang, F. Chen, C. Pan, J. Tang, K.A.I. Zhang, G. Yu, Phenothiazine-
based conjugated microporous polymers: Pore surface and bandgap engineering for visible light-
driven aerobic oxidative cyanation, Chem. Eng. J. 408 (2021) 127261.
https://doi.org/10.1016/j.cej.2020.127261
[368] S.D. Ahn, A. Kolodziej, R. Malpass-Evans, M. Carta, N.B. McKeown, S.D. Bull, A.
Buchard, F. Marken, Polymer of intrinsic microporosity induces host-guest substrate selectivity in

149
heterogeneous 4-benzoyloxy-TEMPO-catalysed alcohol oxidations, Electrocatalysis 7 (2016) 70-
78. https://doi.org/10.1007/s12678-015-0284-8
[369] Y. Rong, D. He, R. Malpass-Evans, M. Carta, N.B. McKeown, M.F. Gromboni, L.H.
Mascaro, G.W. Nelson, J.S. Foord, P. Holdway, S.E.C. Dale, S. Bending, F. Marken, High-
utilisation nanoplatinum catalyst (Pt@cPIM) obtained via vacuum carbonisation in a molecularly
rigid polymer of intrinsic microporosity, Electrocatalysis 8 (2017) 132-143.
https://doi.org/10.1007/s12678-016-0347-5
[370] S.X. Leong, M. Carta, R. Malpass-Evans, N.B. McKeown, E. Madrid, F. Marken, One-step
preparation of microporous Pd@cPIM composite catalyst film for triphasic electrocatalysis,
Electrochem. Commun. 86 (2018) 17-20. https://doi.org/10.1016/j.elecom.2017.11.007
[371] B. Patil, B. Satilmis, T. Uyar, Metal-free N-doped ultrafine carbon fibers from electrospun
Polymers of Intrinsic Microporosity (PIM-1) based fibers for oxygen reduction reaction, J. Power
Sources 451 (2020) 227799. https://doi.org/10.1016/j.jpowsour.2020.227799
[372] B. Patil, B. Satilmis, M.A. Khalily, T. Uyar, Atomic layer deposition of NiOOH/Ni(OH)2
on PIM-1-based N-doped carbon nanofibers for electrochemical water splitting in alkaline medium,
ChemSusChem 12 (2019) 1469-1477. https://doi.org/10.1002/cssc.201802500
[373] D. Jung, S. Su, Z.H. Syed, A. Atilgan, X. Wang, F. Sha, Y. Lei, N.C. Gianneschi, T.
Islamoglu, O.K. Farha, A catalytically accessible polyoxometalate in a porous fiber for
degradation of a mustard gas simulant, ACS Appl. Mater. Interfaces 14 (2022) 16687-16693.
https://doi.org/10.1021/acsami.2c01584
[374] G.H. Moon, H.J. Kim, I.S. Chae, S.C. Park, B.S. Kim, J. Jang, H. Kim, Y.S. Kang, An
artificial solid interphase with polymers of intrinsic microporosity for highly stable Li metal
anodes, Chem. Commun. 55 (2019) 6313-6316. https://doi.org/10.1039/C9CC01329F
[375] L. Qi, L. Shang, K. Wu, L. Qu, H. Pei, W. Li, L. Zhang, Z. Wu, H. Zhou, N.B. McKeown,
W. Zhang, Z. Yang, An interfacial layer based on polymers of intrinsic microporosity to suppress
dendrite growth on Li metal anodes, Chem. Eur. J. 25 (2019) 12052-12057.
https://doi.org/10.1002/chem.201902124
[376] L. Ma, C. Fu, L. Li, K.S. Mayilvahanan, T. Watkins, B.R. Perdue, K.R. Zavadil, B.A. Helms,
Nanoporous polymer films with a high cation transference number stabilize lithium metal anodes
in light-weight batteries for electrified transportation, Nano Lett. 19 (2019) 1387-1394.
https://doi.org/10.1021/acs.nanolett.8b05101
[377] X. Shen, X. Cheng, P. Shi, J. Huang, X. Zhang, C. Yan, T. Li, Q. Zhang, Lithium–matrix
composite anode protected by a solid electrolyte layer for stable lithium metal batteries, J. Energy
Chem. 37 (2019) 29-34. https://doi.org/10.1016/j.jechem.2018.11.016
[378] Y. Tian, C. Lin, Z. Wang, J. Jin, Polymer of intrinsic microporosity-based macroporous
membrane with high thermal stability as a Li-ion battery separator, RSC Adv. 9 (2019) 21539-
21543. https://doi.org/10.1039/C9RA02308A
[379] S.E. Doris, A.L. Ward, P.D. Frischmann, L. Li, B.A. Helms, Understanding and controlling
the chemical evolution and polysulfide-blocking ability of lithium–sulfur battery membranes cast
from polymers of intrinsic microporosity, J. Mater. Chem. A 4 (2016) 16946-16952.
https://doi.org/10.1039/C6TA06401A
[380] A.L. Ward, S.E. Doris, L. Li, M.A. Hughes, Jr., X. Qu, K.A. Persson, B.A. Helms, Materials
genomics screens for adaptive ion transport behavior by redox-switchable microporous polymer
membranes in lithium–sulfur batteries, ACS Central Sci. 3 (2017) 399-406.
https://doi.org/10.1021/acscentsci.7b00012

150
[381] X. Yu, S. Feng, M.J. Boyer, M. Lee, R.C. Ferrier, N.A. Lynd, G.S. Hwang, G. Wang, S.
Swinnea, A. Manthiram, Controlling the polysulfide diffusion in lithium-sulfur batteries with a
polymer membrane with intrinsic nanoporosity, Mater. Today Energy 7 (2018) 98-104.
https://doi.org/10.1016/j.mtener.2018.01.002
[382] Y. Zhao, X. Ma, P. Li, Y. Lv, J. Huang, H. Zhang, Y. Shen, Q. Deng, X. Liu, Y. Ding, Y.
Han, Bifunctional polymer-of-intrinsic-microporosity membrane for flexible Li/Na–H2O2
batteries with hybrid electrolytes, J. Mater. Chem. A 8 (2020) 3491-3498.
https://doi.org/10.1039/C9TA13210D
[383] J.W. Jeon, D.M. Kim, J. Lee, J.C. Lee, Y.S. Kim, K.T. Lee, B.G. Kim, PIM-1-based carbon-
sulfur composites for sodium-sulfur batteries that operate without the shuttle effect, J. Mater.
Chem. A 8 (2020) 3580–3585. https://doi.org/10.1039/C9TA10939K
[384] J.W. Jeon, J.H. Han, S.-K. Kim, D.-G. Kim, Y.S. Kim, D.H. Suh, Y.T. Hong, T.-H. Kim,
B.G. Kim, Intrinsically microporous polymer-based hierarchical nanostructuring of electrodes via
nonsolvent-induced phase separation for high-performance supercapacitors, J. Mater. Chem. A 6
(2018) 8909-8915. https://doi.org/10.1039/C8TA02451K
[385] N. Hernandez, J. Iniesta, V.M. Leguey, R. Armstrong, S.H. Taylor, E. Madrid, Y. Rong, R.
Castaing, R. Malpass-Evans, M. Carta, N.B. McKeown, F. Marken, Carbonization of polymers of
intrinsic microporosity to microporous heterocarbon: Capacitive pH measurements, Appl. Mater.
Today 9 (2017) 136-144. https://doi.org/10.1016/j.apmt.2017.06.003
[386] J.S. Bonso, G.D. Kalaw, J.P. Ferraris, High surface area carbon nanofibers derived from
electrospun PIM-1 for energy storage applications, J. Mater. Chem. A 2 (2014) 418-424.
https://doi.org/10.1039/C3TA13779A
[387] C.Y. Li, A.L. Ward, S.E. Doris, T.A. Pascal, D. Prendergast, B.A. Helms, Polysulfide-
blocking microporous polymer membrane tailored for hybrid Li-sulfur flow batteries, Nano Lett.
15 (2015) 5724–5729. https://doi.org/10.1021/acs.nanolett.5b02078
[388] D.-S. Yang, J.H. Han, J.W. Jeon, J.Y. Lee, D.-G. Kim, D.H. Seo, B.G. Kim, T.-H. Kim, Y.T.
Hong, Multimodal porous and nitrogen-functionalized electrode based on graphite felt modified
with carbonized porous polymer skin layer for all-vanadium redox flow battery, Mater. Today
Energy 11 (2019) 159-165. https://doi.org/10.1016/j.mtener.2018.11.003
[389] I.S. Chae, T. Luo, G.H. Moon, W. Ogieglo, Y.S. Kang, M. Wessling, Ultra-high
proton/vanadium selectivity for hydrophobic polymer membranes with intrinsic nanopores for
redox flow battery, Adv. Energy Mater. 6 (2016) 1600517.
https://doi.org/10.1002/aenm.201600517
[390] S.E. Doris, A.L. Ward, A. Baskin, P.D. Frischmann, N. Gavvalapalli, E. Chénard, C.S.
Sevov, D. Prendergast, J.S. Moore, B.A. Helms, Macromolecular design strategies for preventing
active-material crossover in non-aqueous all-organic redox-flow Batteries, Angew. Chem. Int. Ed.
56 (2017) 1595-1599. https://doi.org/10.1002/anie.201610582
[391] K.H. Hendriks, S.G. Robinson, M.N. Braten, C.S. Sevov, B.A. Helms, M.S. Sigman, S.D.
Minteer, M.S. Sanford, High-performance oligomeric catholytes for effective macromolecular
separation in nonaqueous redox flow batteries, ACS Central Sci. 4 (2018) 189-196.
https://doi.org/10.1021/acscentsci.7b00544
[392] M.B. Karimi, F. Mohammadi, K. Hooshyari, Recent approaches to improve Nafion
performance for fuel cell applications: A review, Int. J. Hydrog. Energy 44 (2019) 28919-28938.
https://doi.org/10.1016/j.ijhydene.2019.09.096

151
[393] H. Tang, G. Chao, J. Gao, Y. Shang, N. Li, K. Geng, Tailoring of microporosity of Tröger's
base (TB) high temperature proton exchange membrane by miscible polymer blending, J. Power
Sources 565 (2023) 232868. https://doi.org/10.1016/j.jpowsour.2023.232868
[394] C. Hu, Q. Zhang, C. Lin, Z. Lin, L. Li, F. Soyekwo, A. Zhu, Q. Liu, Multi-cation crosslinked
anion exchange membranes from microporous Tröger's base copolymers, J. Mater. Chem. A 6
(2018) 13302-13311. https://doi.org/10.1039/C8TA02153H
[395] S. Gong, L. Li, L. Ma, N.A. Qaisrani, J. Liu, G. He, F. Zhang, Blend anion exchange
membranes containing polymer of intrinsic microporosity for fuel cell application, J. Membr. Sci.
595 (2020) 117541. https://doi.org/10.1016/j.memsci.2019.117541
[396] S. Gong, L. Bai, L. Li, N.A. Qaisrani, L. Ma, G. He, F. Zhang, Block copolymer anion
exchange membrane containing polymer of intrinsic microporosity for fuel cell application, Int. J.
Hydrog. Energy 46 (2021) 2269-2281. https://doi.org/10.1016/j.ijhydene.2020.10.068
[397] N.A. Rakow, M.S. Wendland, J.E. Trend, R.J. Poirier, D.M. Paolucci, S.P. Maki, C.S. Lyons,
M.J. Swierczek, Visual indicator for trace organic volatiles, Langmuir 26 (2010) 3767-3770.
https://doi.org/10.1021/la903483q
[398] Y. Wang, N.B. McKeown, K.J. Msayib, G.A. Turnbull, I.D.W. Samuel, Laser chemosensor
with rapid responsivity and inherent memory based on a polymer of intrinsic microporosity,
Sensors 11 (2011) 2478-2487. https://doi.org/10.3390/s110302478
[399] K. Polak-Kraśna, M. Tian, S. Rochat, N. Gathercole, C. Yuan, Z. Hao, M. Pan, A.D. Burrows,
T.J. Mays, C.R. Bowen, Solvent sorption-induced actuation of composites based on a polymer of
intrinsic microporosity, ACS Appl. Polym. Mater. 3 (2021) 920-928.
https://doi.org/10.1021/acsapm.0c01215
[400] X. Wang, C. Yu, H. Guo, Y. Cheng, Y. Li, D. Zheng, S. Feng, Y. Lin, Robust fluorescent
detection of iodine vapor by a film sensor based on a polymer of intrinsic microporosity, Chem.
Eng. J. 438 (2022). https://doi.org/10.1016/j.cej.2022.135641
[401] B.R. Putra, M. Carta, R. Malpass-Evans, N.B. McKeown, F. Marken, Potassium cation
induced ionic diode blocking for a polymer of intrinsic microporosity | Nafion “heterojunction”
on a microhole substrate, Electrochim. Acta 258 (2017) 807-813.
https://doi.org/10.1016/j.electacta.2017.11.130
[402] B.R. Putra, B.D.B. Aaronson, E. Madrid, K. Mathwig, M. Carta, R. Malpass-Evans, N.B.
McKeown, F. Marken, Ionic diode characteristics at a polymer of intrinsic microporosity (PIM) |
Nafion “heterojunction” deposit on a microhole poly(ethylene-terephthalate) substrate,
Electroanalysis 29 (2017) 2217-2223. https://doi.org/10.1002/elan.201700247
[403] Z. Li, L. Wang, R. Malpass-Evans, M. Carta, N.B. McKeown, K. Mathwig, P.J. Fletcher, F.
Marken, Ionic diode and molecular pump phenomena associated with caffeic acid accumulated
into an intrinsically microporous polyamine (PIM-EA-TB), ChemElectroChem 8 (2021) 2044-
2051. https://doi.org/10.1002/celc.202100432
[404] N. Jahani, M. Amiri, M. Ghiasi, H. Imanzadeh, R. Boukherroub, S. Szunerits, F. Marken,
N.B. McKeown, Non-enzymatic electrochemical cholesterol sensor based on strong host-guest
interactions with a polymer of intrinsic microporosity (PIM) with DFT study, Anal. Bioanal. Chem.
413 (2021) 6523-6533. https://doi.org/10.1007/s00216-021-03616-w
[405] Y. Zhao, R. Malpass‐Evans, M. Carta, N.B. McKeown, P.J. Fletcher, G. Kociok‐Köhn,
D. Lednitzky, F. Marken, Size ‐ selective photoelectrochemical reactions in microporous
environments: Clark probe investigation of Pt@g‐C3N4 embedded into intrinsically microporous

152
polymer (PIM ‐ 1), ChemElectroChem 8 (2021) 3499-3505.
https://doi.org/10.1002/celc.202100732
[406] Y. Zhao, J. Dobson, C. Harabajiu, E. Madrid, T. Kanyanee, C. Lyall, S. Reeksting, M. Carta,
N.B. McKeown, L. Torrente-Murciano, K. Black, F. Marken, Indirect photo-electrochemical
detection of carbohydrates with Pt@g-C3N4 immobilised into a polymer of intrinsic microporosity
(PIM-1) and attached to a palladium hydrogen capture membrane, Bioelectrochemistry 134 (2020)
107499. https://doi.org/10.1016/j.bioelechem.2020.107499
[407] E. Madrid, D. He, J. Yang, C.F. Hogan, B. Stringer, K.J. Msayib, N.B. McKeown, P.R.
Raithby, F. Marken, Reagentless electrochemiluminescence from a nanoparticulate polymer of
intrinsic microporosity (PIM-1) immobilized onto tin-doped indium oxide, ChemElectroChem 3
(2016) 2160-2164. https://doi.org/10.1002/celc.201600419
[408] H. Afshary, M. Amiri, F. Marken, N.B. McKeown, M. Amiri, ECL sensor for selective
determination of citrate ions as a prostate cancer biomarker using polymer of intrinsic
microporosity-1 nanoparticles/nitrogen-doped carbon quantum dots, Anal. Bioanal. Chem. 415
(2023) 2727-2736. https://doi.org/10.1007/s00216-023-04672-0
[409] A. Haleem, M. Javaid, R.H. Khan, R. Suman, 3D printing applications in bone tissue
engineering, J Clin. Orthop. Trauma 11 (2020) S118-S124.
https://doi.org/10.1016/j.jcot.2019.12.002
[410] P.K. Bg, S. Mehrotra, S.M. Marques, L. Kumar, R. Verma, 3D printing in personalized
medicines: A focus on applications of the technology, Mater. Today Commun. 35 (2023) 105875.
https://doi.org/10.1016/j.mtcomm.2023.105875
[411] A. Ambrosi, M. Pumera, 3D-printing technologies for electrochemical applications, Chem.
Soc. Rev. 45 (2016) 2740-2755. https://doi.org/10.1039/c5cs00714c
[412] M.N. Nadagouda, M. Ginn, V. Rastogi, A review of 3D printing techniques for
environmental applications, Curr. Opin. Chem. Eng. 28 (2020) 173-178.
https://doi.org/10.1016/j.coche.2020.08.002
[413] Z. Liu, M. Zhang, B. Bhandari, Y. Wang, 3D printing: Printing precision and application in
food sector, Trends Food Sci. Technol. 69 (2017) 83-94. https://doi.org/10.1016/j.tifs.2017.08.018
[414] A. Vanderploeg, S.-E. Lee, M. Mamp, The application of 3D printing technology in the
fashion industry, Int. J. Fash. Des. Technol. Educ. 10 (2016) 170-179.
https://doi.org/10.1080/17543266.2016.1223355
[415] F. Zhang, C. Tuck, R. Hague, Y. He, E. Saleh, Y. Li, C. Sturgess, R. Wildman, Inkjet printing
of polyimide insulators for the 3D printing of dielectric materials for microelectronic applications,
J. Appl. Polym. Sci. 133 (2016) 43361. https://doi.org/10.1002/app.43361
[416] J. Kim, R. Kumar, A.J. Bandodkar, J. Wang, Advanced materials for printed wearable
electrochemical devices: A review, Adv. Electron. Mater. 3 (2017) 1600260.
https://doi.org/10.1002/aelm.201600260
[417] D.-Y. Koh, B.A. McCool, H.W. Deckman, R.P. Lively, Reverse osmosis molecular
differentiation of organic liquids using carbon molecular sieve membranes, Science 353 (2016)
804-807. https://doi.org/10.1126/science.aaf1343
[418] S. Jiang, L. Chen, M.E. Briggs, T. Hasell, A.I. Cooper, Functional porous composites by
blending with solution-processable molecular pores, Chem. Commun. 52 (2016) 6895-6898.
https://doi.org/10.1039/C6CC01034B
[419] Q. Song, S. Jiang, T. Hasell, M. Liu, S. Sun, A.K. Cheetham, E. Sivaniah, A.I. Cooper,
Porous organic cage thin films and molecular-sieving membranes, Adv. Mater. 28 (2016) 2629-
2637. https://doi.org/10.1002/adma.201505688

153
[420] S.W. Pattinson, A.J. Hart, Additive manufacturing of cellulosic materials with robust
mechanics and antimicrobial functionality, Adv. Mater. Technol. 2 (2017) 1600084.
https://doi.org/10.1002/admt.201600084
[421] F. Zhang, Y. Ma, J. Liao, V. Breedveld, R.P. Lively, Solution-based 3D printing of polymers
of intrinsic microporosity, Macromol. Rapid Commun. 39 (2018) e1800274.
https://doi.org/10.1002/marc.201800274

154

You might also like