Jun-Song
Jia
a,
Ying
Liang
*b and
Ying-Ming
Pan
*c
aCollege of Chemistry and Environmental Engineering, Key Laboratory of Green Catalysis of Higher Education Institutes of Sichuan, Sichuan University of Science and Engineering, Zigong, 643000, Sichuan, P. R. China
bSchool of Life and Environmental Sciences, Guilin University of Electronic Technology, Guilin 541004, P. R. China. E-mail: liangyi0774@guet.edu.cn
cState Key Laboratory for Chemistry and Molecular Engineering of Medicinal Resources, School of Chemistry and Pharmaceutical Sciences of Guangxi Normal University, Guilin 541004, P. R. China. E-mail: panym@mailbox.gxnu.edu.cn
First published on 8th August 2025
Direct conversion of waste CO2 avoids CO2 capture and lowers the cost of CO2 utilisation; however, this route remains a challenging research topic. Developing catalysts that facilitate the enrichment and conversion of waste CO2 is therefore essential. Porous polymer catalysts offer unique advantages due to their high surface area and tunable functionality. These materials catalyse the conversion of both simulated waste CO2 and CO2 present in industrial waste gases, such as anaerobic fermentation gas, lime kiln waste gas, and coal-fired flue gas. This review summarises recent progress on the direct conversion of waste CO2 using porous polymer catalysts. It analyses the structural features of these catalysts, their CO2 adsorption properties, and the associated catalytic mechanisms. A quantitative comparison of catalytic performance—such as turnover frequency, stability, and CO2 adsorption capacity—is also provided. The findings may support the rational design and synthesis of catalysts for the direct utilisation of waste CO2, and provide parameters for the industrialisation of porous polymer catalysts.
CO2 is a low-cost, safe, non-toxic, and abundant C1 source.8,9 Its conversion by chemical methods helps reduce CO2 emissions and yields a range of high value-added products, including synthons, fine chemicals, pharmaceuticals, and hydrocarbon fuels.10–13 Many methods have been developed to convert CO2 into bulk chemicals, such as CO, CH3OH, and HCOOH.14,15 CO2 can also be introduced into organic structures through catalysis to synthesise carboxylic acids, esters, and carbonyl compounds.16–19 However, these reactions require pure CO2 and generally convert CO2 under high pressure (Scheme 1A). Currently, the capture of CO2 from industrial waste gas is usually achieved through liquid or solid adsorption, followed by heating and desorption, and then compression and storage.20 Desorption and compression require a large amount of energy, resulting in high costs. For example, the cost of capturing one tonne of CO2 by amine washing is $59.10 for a typical coal-fired power plant.21 Therefore, the steps for capture and storage between CO2 emissions and utilisation must be reduced. If the CO2 in exhaust gas generated by burning fossil fuels is directly utilised, the steps of capture and storage can be eliminated, thereby reducing the cost of CO2 utilisation (Scheme 1B).
The exhaust gas emitted from industrial production has a low CO2 concentration. For example, flue gas contains CO2 at 10–15 vol%.20 Anaerobic fermentation gas (AFG) typically consists of CH4, CO2, N2, and H2S, with a CO2 concentration in the range of 30–35 vol%.22,23 Waste gas from lime kilns mainly comprises N2, O2, CO2, CO, SO2, H2O, and dust, with a CO2 concentration of 20–42 vol%.24–26 In addition, activated CO2 can be converted into small molecule compounds such as CO and HCOOH.27–29 In catalytic systems that convert CO2 into high value-added products, these small molecule products are often overlooked. Therefore, the direct conversion of waste CO2 faces three challenges: (1) the catalyst must have the capacity for in situ enrichment of low-concentration CO2; (2) the catalyst must accurately identify CO2 in exhaust gases without interference from other gas molecules; and (3) selective conversion of CO2. Hence, a new catalytic system with enrichment and conversion effects for low-concentration CO2 must be developed.
Porous polymer catalysts are a new class of porous materials with the advantages of high specific surface area, diverse pore structures, low skeleton density, and facile functionalisation.30–32 They have become versatile hard ligands and carriers for converting many homogeneous catalysts into heterogeneous nano or single-site active-centre catalysts with unique catalytic activity and stability.33–35 Compared to metal–organic framework (MOF) and covalent organic framework (COF)-based catalysts, polymer catalysts present several advantages: (1) the structure of polymer catalysts is characterised by weak pore uniformity and topological ordering.36–38 The hierarchical pore structure promotes the physical adsorption of CO2. (2) Polymer catalysts are more thermally stable. They are also more stable in water, acid, and base, which broadens their applicability.37,39 (3) After the loading of metal onto polymer catalysts, the metal is distributed on the surface, facilitating its contact with reactants. The metals within MOFs are typically positioned at the core of the crystal structure, reducing the accessibility of the substrate.40,41 In addition, porous polymer catalysts have advantages over non-porous catalytic systems: (1) the porous 3D structure increases the catalyst's surface area and allows more active sites to be loaded per unit volume, increasing the probability of CO2 coming in contact with the catalyst.30,39 (2) The hierarchical pore structure, especially micropores, contributes to CO2 enrichment.42 (3) The swelling properties of porous structures can create a “quasi-homogeneous” catalytic system that accelerates mass transfer and improves catalytic efficiency.43
Porous polymer catalysts include porous organic polymers (POPs), porous ionic liquid polymers (PILs), porous organic ligand polymers (POLs), and hypercrosslinked polymers (HCPs).39 Incorporating active sites that can activate or enrich CO2 into the framework of porous polymers creates an ideal platform for heterogeneous catalytic CO2 conversion. Chemists have reviewed the use of porous catalysts for the conversion of pure CO2. In 2018, Huang, Liu, and co-workers reviewed the design and synthesis progress of POPs for catalytic CO2 conversion.21 The design strategies of these catalysts include nitrogen doping, metallisation, and ionic functionalisation of POPs. In 2022, Wang, Zhou, and co-workers summarised the synthesis of porous PILs and the catalytic details for achieving cycloaddition between pure CO2 and epoxides.42 However, only a few reports are available on the enrichment and conversion of CO2 in exhaust gas catalysed by porous polymers.
In this review, the most recent advances in the design and synthesis of porous polymer catalysts for the conversion of low-concentration CO2 are summarised. The discussion highlights the synthesis strategies of porous polymers with enrichment effects for low CO2 concentration and the accurate identification of CO2 molecules in exhaust gas (simulated exhaust gas). Problems in this field are analysed and future development directions are outlined. This review is divided into two sections: the first addresses simulation of waste CO2, and the second addresses real waste CO2. Within each section, further categorisation distinguishes waste CO2 conversion with metal catalysis from that without metal catalysis.
Porphyrins are a class of conjugated macrocyclic structures with 18π electron systems, which enable strong interactions with CO2.47 Furthermore, the cavity structure of porphyrin-based polymers can coordinate with metals to facilitate their loading. Therefore, many porphyrin-based porous materials have been used to convert pure CO2. In 2016, Xiao, Meng, and co-workers utilised porphyrin POP loaded with Co3+ to synthesise a heterogeneous catalyst, Co/POM-TPP, which enabled the reaction of low-concentration CO2 (15 vol%) with epichlorohydrin 1 by cycloaddition, yielding cyclic carbonates 2 with a 45.4% conversion rate and 88.7% selectivity (Scheme 2a).48 The catalytic system required the addition of tetrabutylammonium bromide (TBAB) as a cocatalyst. The study found that the activity of the heterogeneous Co/POM-TPP catalyst exceeded that of the corresponding homogeneous Co/TPP catalyst. This difference can be attributed to the strong adsorption capacity of Co/POP-TPP for 15 vol% CO2 (adsorption capacity of 8 cm3 g−1), whereas the nonporous Co/TPP showed negligible CO2 adsorption. Activation of the adsorbed CO2 can be attributed to two factors: the acid–base interaction between nitrogen atoms in porphyrins and CO2, and the strong interaction between the porphyrin π system and CO2. This catalyst can convert low concentrations of CO2 at 1 atm and room temperature, which is relevant for utilising CO2 generated during industrial processes. Although the catalyst successfully enabled the conversion of simulated waste CO2, the product yield remained relatively low.
In the cycloaddition reaction between epoxides and CO2, it is customary to add metal ions or halide anions, both of which serve to activate the substrate.21 Therefore, the design and synthesis of porous polymer catalysts containing both metal ions and halogen anions will contribute to the conversion of waste CO2. In 2021, Xiao's group continued to investigate and synthesise heterogeneous catalysts with multiple synergistic active sites to address the challenge of adding large quantities of soluble cocatalyst TBAB, as shown in Scheme 2a.46 They prepared a porous organic catalyst, POP–PBnCl–TPPMg-x, containing both Lewis acidic and Lewis basic active sites, by copolymerising different proportions of Mg–porphyrin monomers and phosphonium salt monomers (Scheme 2b). Among these, the catalyst POP–PBnCl–TPPMg-12 catalysed the cycloaddition of 15 vol% CO2 with epoxides to form cyclic carbonates with a 95.1% conversion rate and 73.8% selectivity.
In general, increasing the surface area of polymer catalysts is considered beneficial for enhancing reactant accessibility to active sites.21 However, the results from the Co/POM-TPP and POP-PBnCl-TPPMg-12 studies suggest that the Brunauer–Emmett–Teller (BET) surface area and pore volume are not the determining factors in waste CO2 conversion. Halogen anions integrated onto polymers provide improved cocatalytic effects, likely owing to better dispersion of halogens on the polymer. Additionally, POP–PBnCl–TPPMg-12 does not require cocatalysts and offers a broad range of parameters for converting low-concentration CO2. The use of MeCN as a solvent in the catalytic system significantly increased the cyclic carbonate yield, although the mechanism underlying this enhancement was not explained by the authors.
The disordered hierarchical pore structure is a prominent feature of porous polymer catalysts. Pore structures of varying volumes in polymers may serve different functions in the conversion of waste CO2. In 2016, Yang, Li, and co-workers prepared a hierarchical meso/microporous polymer catalyst containing a 2,2′-bipyridine Zn(II) coordination structure, Bp-Zn@MA (Scheme 3).49 When TBAB was used as a cocatalyst, Bp-Zn@MA catalysed the cycloaddition reaction of simulated flue gas (20 vol% CO2) and epoxide to produce cyclic carbonates 2. The N2 adsorption test showed that the micropore volume of Bp-Zn@MA was 0.002 cm3 g−1, representing 0.198% of the total pore volume. The authors verified, using Henry's law and ideal adsorption solution theory, that Bp-Zn@MA exhibits higher adsorption selectivity for CO2 than N2 in mixed gases. Notably, the micropores and abundant N sites in Bp-Zn@MA show selective adsorption for CO2 in CO2/N2 mixed gases, whereas the role of mesopores is to facilitate the rapid diffusion of reactants and products. Therefore, micropores exert a more pronounced enrichment effect on waste CO2. However, the “division of labour” mechanism between micropores and mesopores in catalysis requires further investigation.
In parallel, the authors proposed a mechanism for this reaction. The catalytic cycle begins with the coordination of epoxyethane compounds with zinc to produce compound 3. Bromide anions then perform a nucleophilic attack on epoxyethane to generate Zn-coordinated ring-opening intermediate 4, which combines with CO2 to afford zinc carbonate species 5. Finally, the intramolecular cyclisation of zinc carbonate species generates the final product 2 and releases bromide anions and regenerated catalyst Bp-Zn@MA, thus completing the catalytic cycle. In this process, zinc activates epoxides, whereas the activation of CO2 is mainly due to the acid–base interaction between the alkaline N site and CO2. Unfortunately, the reaction conditions are harsh (high temperature and pressure) and not suitable for large-scale catalyst application. In addition, the catalytic efficiency decreases when the reaction is carried out in a mixed gas at atmospheric pressure.
Nitrogen heterocyclic carbene (NHC) ligands contain multiple nitrogen atoms that can interact with the carbon in CO2 to activate it. NHC polymer catalysts also exhibit stronger alkalinity than N-doped polymers, facilitating acid–base interactions with CO2.21 PILs combine the advantages of porous materials, polymers, and ionic liquids, and have shown promise in CO2 fixation by cycloaddition.42 PILs synthesised from NHC ligand polymers demonstrate excellent catalytic performance in the conversion of low-concentration CO2.50–54 In 2024, the Liu group prepared a poly(ionic liquid)@porous carbon nanocomposite (PIL-Brx@Zn-CTM) through the pyrolysis of fluid precursors and in situ polymerisation (Scheme 4).55 PIL-Brx@Zn-CTM contains Lewis acidic (Zn2+) and Lewis basic (N) sites and a nucleophilic anion Br−. PIL-Br1.0@Zn-CTM can absorb low concentrations of CO2 and catalyse its reaction with epoxides to produce cyclic carbonates 2. The catalytic system does not require cocatalysts or solvents, making it a green approach. In this heterogeneous catalyst, zinc has strong coordination ability, whereas bromide anions have high nucleophilicity. Therefore, the presence of zinc and bromide anions activates epoxides. In addition to the alkaline NHC and N-doped sites, the strong interaction between the large conjugated structure of carbon nanocomposites and CO2 may also benefit CO2 adsorption and conversion. However, this catalytic reaction requires a high temperature, which limits its industrial application.
![]() | ||
Scheme 4 Poly(ionic liquid)@porous carbon nanocomposites for the capture and in situ conversion of low-concentration CO2. |
The process of PIL-Br1.0@Zn-CTM adsorbing and converting low-concentration CO2 is illustrated in Scheme 4. The alkaline N sites on the catalyst surface adsorb and activate low-concentration CO2. Simultaneously, Zn2+ coordinates with the epoxides 1 and induces C–O bond polarisation. The nucleophilic Br− attacks the β-C of epoxyethane, which has minimal steric hindrance, to generate the ring-opening intermediate, which undergoes nucleophilic addition with the CO2 adsorbed by the catalyst to form alkyl carbonate compounds. Finally, intramolecular cyclisation occurs to generate the final product 2 and release PIL-Br1.0@Zn-CTM.
Due to the disorder of the polymer structure, its BET surface area, pore volume, and pore size are difficult to control during preparation. The development of methodologies capable of precisely modulating the surface area and pore volume of polymers remains a significant challenge in this field. In 2017, Wang, Zhou, and co-workers described the selective capture and conversion of CO2 from simulated flue gas (15 vol% CO2) under mild conditions using imidazolinium-based porous hypercrosslinked ionic polymers (HIPs) with cocatalyst ZnBr2 (Scheme 5).56 Notably, the authors adjusted the porosity of HIPs by modifying the initial gel composition. The catalytic effect of this heterogeneous catalyst is substantially better than that of homogeneous ionic liquid monomers and post-modified analogues. The halogen anions in HIPs show high nucleophilicity and high leaving ability, which promote the selective adsorption of CO2. Furthermore, the authors propose that HCP exhibits advantages for CO2 capture and utilisation due to its high surface area and narrowly dispersed micropores. The CO2 conversion mechanism is the same as for Scheme 3. In addition to halide anions, the alkalinity of NHCs contributes to CO2 activation. Zn2+ assists in epoxide activation.
As depicted in Scheme 6, the mechanism of this transformation begins with the formation of hydrogen bonds between the epoxyethane compound 1 and imidazolium salts to afford 6, followed by Br− nucleophilic attack on epoxyethane to generate ring-opening products 7. CO2 insertion into O−⋯H then produces opening carbonates 8, which undergo nucleophilic attack to form ring-closing products 2 and release the catalysts. It can be inferred from the mechanism that NHC monomers also activate epoxides. However, the mechanisms of CO2 adsorption and activation were not clarified by the authors.
Progress has been made in enriching and converting simulated waste CO2 using porous polymer catalysts. Some catalysts can achieve the conversion of low-concentration CO2 at room temperature, atmospheric pressure, and without co-catalysts, confirming the potential of polymer catalysts for industrial waste CO2 conversion. However, further research is required into the underlying mechanisms of CO2 enrichment by porous polymer catalysts. In addition, the above waste CO2 reactions are carried out in closed reactors. Given the fluidity of industrial waste gases, studying the conversion of CO2 in flowing gases will provide more relevant reference data.
As shown in Scheme 8, a continuous flow device introduces 30 vol% CO2 into a heterogeneous catalytic system via bubbling. Under analogous conditions, the continuous flow apparatus achieves a product yield of 96%, whereas the fixed reactor yields 72%. The authors suggest that the observed difference in yield results from more effective contact between the waste CO2, the catalyst, and the substrate in the continuous flow device. This finding highlights the potential application of the catalyst for CO2 enrichment and conversion in industrial waste gas. However, the authors did not explore the mechanism of selective adsorption and enrichment of CO2 in lime kiln exhaust gas by the hierarchical pore structure of the catalyst.
Increasing the number of NHC and metal sites in the monomer enhances the probability of CO2 contacting active sites, thereby improving the conversion efficiency. In 2023, Liang, Pan and co-workers synthesised the nanocatalyst Ag@POP-NL-3 by pre-coordination and polymerisation of a tridentate NHC monomer with AgNO3 (Scheme 9).60 Coordinating before polymerising ensures uniform distribution of Ag on POL-NL-3 and improves catalytic efficiency. The hierarchical pore structure and abundant nitrogen sites of the catalyst guarantee the adsorption and conversion of CO2. Using Ag@POP-NL-3 as the heterogeneous catalyst and 0.6 equivalent of DBU as an additive, CO2 (30 vol%) in lime kiln waste gas reacts with propargyl alcohol 15 to form cyclic carbonates 16. Raman spectroscopy detected the stretching vibration (1183 cm−1) and antisymmetric stretching (2211 cm−1) of the N–CO2 structure. This result confirms the adsorption and activation of CO2 by N species within the polymer.
Following the addition of Ag@POP-NL-3 to propargylic alcohol, the 1H nuclear magnetic resonance absorption peak signal of propargylic alcohol decreased, indicating that Ag@POP-NL-3 exerts an adsorption effect on the substrate. The mechanism of this reaction is shown in Scheme 9. First, the hierarchical pore structure and abundant nitrogen sites of the catalyst enrich the low-concentration CO2 to a locally high concentration. DBU and propargylic alcohol 15 then interact through hydrogen bonding, and Ag coordinates with CC to form intermediate 17. The oxygen atom of propargylic alcohol attacks CO2, inserting CO2 and providing intermediate 18 while releasing DBU-H+19. Intermediate 18 undergoes intramolecular cyclisation with carbonate and C
C to form intermediate 20, followed by reaction with 19 to produce cyclic carbonate 16. Through controlled experiments, the authors demonstrated that DBU activates propargylic alcohol. It is speculated that, as an organic base, DBU may also activate CO2. Therefore, exploring the catalytic effect of DBU on this reaction in the absence of polymer catalysts would improve the accuracy of the proposed mechanism.
The addition of nitrogen-containing components and co-polymerisation with NHC monomers can effectively increase the alkaline nitrogen sites in the polymer. In 2024, Pan, Liang and co-workers added copolymer allyl aniline to the bidentate NHC ligand to synthesise polymer POP-6 with abundant nitrogen sites, and then loaded it with Ag2O to form POP-6@Ag2O (Scheme 10).61 This polymer catalyst enables the cyclisation between CO2 in coal-fired flue gas (8 vol%) and propargylamides 21 to produce oxazolidine-2,4-diones 22. The amine groups in the polymer selectively absorb CO2 from coal-fired flue gas, and imidazole and silver nanoparticles activate CO2 and propargylamides.
The mechanism of POP-6@Ag2O adsorbing and converting waste CO2 is shown in Scheme 10. The CO2 in coal-fired flue gas is first adsorbed in the hierarchical pores of the polymer and then onto the amino group. The CO2 adsorbed by the amine group is transferred to the carbene nitrogen for activation. The nitrogen in propargylamine attacks activated CO2, and Ag activates alkynes. Finally, intermediate 23 undergoes intramolecular cyclisation to generate the target product 22. The enrichment and activation of waste CO2 with a concentration below 10 vol% in the presence of other gases is a challenging process. The hierarchical porous structure of polymer catalysts facilitates the enrichment of CO2 and other gases. Therefore, the selectivity of polymers towards waste CO2 is predominantly attributed to the interaction between the alkaline sites of the polymer and CO2. Notably, the stronger alkalinity of aliphatic amines compared to aromatic amines means that integrating alkyl amines into polymer catalysts could improve their ability to enrich and activate waste CO2.
More importantly, the authors demonstrated through controlled experiments that neither SO2 nor NO2 affects the cyclisation reaction. This reaction is unaffected by SO2, NOX and volatile organic compounds in actual coal-fired flue gas. Therefore, it can be deduced that the presence of other gases in the mixed gas does not affect the catalytic performance of the polymer catalyst. However, the efficiency of the polymer catalyst may decrease as the reaction progresses if the metal loaded on it is sensitive to water and oxygen.
Compared with the two-component reaction involving waste CO2, the three-component reaction is more challenging. In 2025, Pan, Liang and co-workers copolymerised a six-membered NHC monomer with DVB to generate POP-1, which was then loaded with Pd(PPh3)4 to form the catalyst POP-1-Pd (Scheme 11).62 At 298.15 K, the CO2 adsorption capacity of POP-1-Pd is 11.62 m3 g−1. Under the action of NaOtBu, POP-1-Pd catalyses a three-component reaction of CO2 from lime kiln exhaust gas with propargylamine 9 and aryl iodide 24 to produce oxazolidinone 25. When NHC:
DVB = 1
:
4, the catalyst has the best effect on CO2 adsorption and conversion. This finding indicates that the specific surface area of the catalyst has a significant effect on the utilisation of CO2. Moreover, this catalytic system can be applied to the gram-level synthesis of oxazolidinone (100 vol% CO2, 1.72 g).
CO2 is enriched near the catalytic sites through the adsorption of imidazolium salts. In parallel, Pd0 undergoes oxidative addition with aryl iodide 24 and coordinates with the CC of propargylamine 9 to form intermediate 26. CO2 is then integrated into intermediate 27via the imidazole structure. The N− in propargylamine attacks CO2 to produce intermediate 28, which undergoes intramolecular cyclisation to generate intermediate 29. Finally, 29 undergoes reductive elimination to yield the desired product 25 and regenerate the catalyst. The hierarchical pores of POP-1-Pd provide a microenvironment with increased concentration after adsorbing CO2 and substrates. According to the mechanism proposed by the authors, NHC adsorbs and activates waste CO2, while palladium activates iodobenzene and propargylamine. There is no direct interaction between palladium and CO2. Furthermore, the addition of stoichiometric strong base NaOtBu renders the reaction conditions harsh.
Interestingly, the anions in the imidazole polymer can be exchanged with other anions, thereby enhancing the ability of the polymer to enrich and convert waste CO2. In 2025, Pan, Liang and co-workers developed an Ag@POP-HCO3 catalysed reaction of propargylamine 30 with CO2 in coal-fired flue gas (8 vol%), producing oxazolidinones 31 (Scheme 12).63 Ag@NHC–HCO3 can be obtained by modifying the Ag@NHC ligand with HCO3−, which enhances the catalyst's ability to adsorb CO2. The specific surface area and micropore volume of the catalyst can be controlled by changing the NHC:
HCO3−
:
DVB ratio. Moreover, the authors designed simulated flue gas with different SO2 and NO2 contents. Subsequent controlled experimentation confirmed that SO2 and NO2 have no significant effect on the activity of the catalyst.
The mechanism of this conversion is shown in Scheme 12. CO2 is enriched by the porous structure of polymers and free HCO3− to give 32. The N-nucleophile of propargylamine then attacks the activated CO2, and Ag interacts with CC through π electrons to form the carbamate intermediate 33, followed by intramolecular cyclisation to generate vinyl-silver 34. Finally, 34 is converted into the final product 31, and the catalyst Ag@NHC–HCO3 is regenerated. HCO3− may promote the conversion of waste CO2 in two ways: (1) as the HCO3− content increases, the specific surface area and microporosity of the catalyst increase, which favours CO2 adsorption; (2) HCO3− contributes to the formation of the NHC–CO2 structure between CO2 and NHC under alkaline conditions.64,65 According to this mechanism, CO2 may be converted into other small-molecule compounds, such as H2CO3. Further analysis of these compounds could clarify the issue of selective CO2 conversion.
As illustrated in Scheme 13, the authors proposed the catalytic cycle. Initially, CO2 in AFG gas is selectively adsorbed onto IPOP-3 through the combined effect of hydroxide ions, nitrogen sites, and pore channels, creating a local atmosphere of high CO2 concentration. PhSiH3 is activated via nucleophilic interaction with the OH− groups in IPOP-3, forming a high-valence silicon centre 37 that reduces CO2 to formate anions and intermediate 38. The formate anion undergoes nucleophilic substitution with intermediate 38 to generate intermediate 39 and regenerate the IPOP-3 catalyst. Finally, amine 35 undergoes nucleophilic substitution with 39 to yield the formamide product 36 and byproduct phenylsilanol. It is evident that waste CO2 is also activated by NHC, whereas the reactant amine is not activated by any species. Nevertheless, the use of stoichiometric phenylsilane and the low atomic utilisation rate of CO2 reduce the atomic economy of this method.
Real waste CO2 can be directly converted into various high-value-added products using porous polymer catalysis, providing a solution to the environmental problems caused by industrial CO2 emissions. Although this process is currently at the pilot stage, it provides important information for potential industrial applications. Additionally, the Liang group59 designed a prototype continuous-flow device for converting waste CO2, demonstrating the possibility of directly introducing emitted exhaust gas into a porous polymer catalytic system to achieve selective adsorption and conversion of CO2. Designing a comprehensive flow catalytic system and exploring further parameters—including exhaust gas introduction, product separation, catalyst regeneration and circulation, and exhaust gas treatment—will help porous polymer catalysts achieve continuous conversion of waste CO2 on a large scale. However, there remains a paucity of discussion regarding selectivity in real CO2 conversion processes.
Catalysts | Preparation method | TOFa | Stability/°C | CO2 adsorption capacityb | Pore sizes/nm |
---|---|---|---|---|---|
a TOF = [n(product)]/[n(ionic or metal sites) × (reaction time)]. b Unless indicated otherwise, the term refers to the adsorption capacity for pure CO2 and 25 °C. | |||||
Co/POP-TPP | Radical polymerization | 4.4 | 350 | 8 cm3 g−1 (15% CO2) | 0.4–2, 12–150 |
POP–PBnCl–TPPMg | Copolymerization | 19.8 | 200 | 50.6 cm3 g−1 | 0.7–9.3 |
Bp-Zn@MA | Polycondensation | 21–1580 | 300 | 0.85 mmol g−1 | 1.5, 4–11 |
PIL-Br1.0@Zn-CTM | Pyrolysis and polymerization | 39–72 | 300 | 1.2 mmol g−1 (0 °C) | 3.9 (average) |
HIP-Br-2 | Friedel–Crafts alkylation | 0.09–0.2 | 300 | 1.9 mmol g−1 | 0.55–1.61, ∼3 |
HCP | Friedel–Crafts alkylation | 136–240 | 750 | 2.02 mmol g−1 (0 °C) | 6.4 (average) |
Cu@NHC-1 | Copolymerization | 4–9.6 | 400 | 13.39 cm3 g−1 | <5 |
Ag@POP-NL-3 | Copolymerization | 1.6–1.8 | 280 | ∼10.8 cm3 g−1 | 50–100 |
POP-6@Ag2O | Copolymerization | — | 350 | 10.5 cm3 g−1 | 1.85–20.48 |
POP-1-Pd | Copolymerization | 5.5–6.8 | 300 | 11.62 cm3 g−1 | 5.87 (average) |
Ag@NHC–HCO3 | Copolymerization | 81–106 | 400 | ∼11.2 cm3 g−1 | ∼3 |
IPOP-3 | Copolymerization | 0.3–0.4 | 250 | 12.3 cm3 g−1 | ∼4 |
Waste CO2 | Additives | T/°C | t/h | Conv. or yield % | Selectivity/% | Cycles | Ref. |
---|---|---|---|---|---|---|---|
15% CO2 + 85% N2, 1 atm | TBAB | 29 | 48 | 45.4 | 88.7 | 18 | 48 |
15% CO2 + 85% N2, 1 atm | — | 40 | 96 | 95.1 | 73.8 | 5 | 46 |
20% CO2 + 80% N2, 2 Mpa | TBAB | 100 | 6 | 11–99 | >99 | 5 | 49 |
15% CO2 + 85% N2, 1 Mpa | — | 100 or 120 | 13 or 20 | 51–99 | ≥99 | 7 | 55 |
15% CO2 + 85% N2, 1 bar | ZnBr2 | 55 | 120 | 85–96 | >99 | 5 | 56 |
15% CO2 + 85% N2, 3 Mpa | — | 120 | 20–32 | 85–96 | 99 | 5 | 57 |
Lime kiln gas (30% CO2), 1 atm | — | 50 | 7.5 or 16 | 56–94 | — | 5 | 59 |
Lime kiln gas (30% CO2), 1 atm | DBU | rt | 12 | 85–95 | — | 7 | 60 |
Coal-fired flue gas (8% CO2), 1 atm | — | rt | 7 | 78–92 | — | 6 | 61 |
Lime kiln gas (30% CO2), 1 atm | NaOtBu | rt | 12 | 54–67 | — | 6 | 62 |
Coal-fired flue gas (8% CO2), 1 atm | — | rt | 36 | 62–81 | — | 3 | 63 |
Anaerobic fermentation gas (46% CO2), 1 atm | PhSiH3 | rt | 24 | 62–90 | — | 5 | 66 |
Waste CO2 | Waste gas | Gas compositions | Experimental conditions | Ref. | ||||
---|---|---|---|---|---|---|---|---|
Catalyst | Additive | Solvent | T/°C | t/h | ||||
Simulated exhaust gas | — | 15% CO2 + 85% N2 | Co/POP-TPP | TBAB | — | 29 | 48 | 48 |
POP–PBnCl–TPPMg | — | MeCN | 40 | 96 | 46 | |||
PIL-Br1.0@Zn-CTM | — | — | 100 or 120 | 13 or 20 | 55 | |||
HIP-Br-2 | ZnBr2 | DMF | 55 | 120 | 56 | |||
HCP | — | — | 120 | 20–32 | 57 | |||
20% CO2 + 80% N2 | Bp-Zn@MA | TBAB | — | 100 | 6 | 49 | ||
Real exhaust gas | Lime kiln gas | 67.9% N2, 30% CO2, 2% O2, 0.1% CO, 15 ppm SO2 | Cu@NHC-1 | — | MeCN | 50 | 7.5 or 16 | 59 |
Ag@POP-NL-3 | DBU | MeCN | rt | 12 | 60 | |||
67% N2, 30% CO2, 2% O2, 0.1% CO, 100 ppm SO2 | POP-1-Pd | NaOtBu | DMSO | rt | 12 | 62 | ||
Coal-fired flue gas | 8% CO2 | POP-6@Ag2O | — | MeCN | rt | 7 | 61 | |
Ag@NHC–HCO3 | — | DMSO | rt | 36 | 63 | |||
Anaerobic fermentation gas | 49% CH4, 46% CO2, 1.5% O2, 66 ppm H2S | IPOP-3 | PhSiH3 | MeCN | rt | 24 | 66 |
Porous polymer catalysts | Functional groups | Metal incorporation | Mechanistic features | Ref. |
---|---|---|---|---|
Porous polymer catalysts loaded with metals | ||||
![]() |
Porphyrin | CoCl2·6H2O | • CO2 activation mode: the acid–base interaction between nitrogen atoms in porphyrins and CO2, and the strong interaction between the porphyrin π system and CO2 | 48 |
• Both cobalt and TBAB activate epichlorohydrins | ||||
![]() |
Porphyrin and phosphonium salt | MgBr2·Et2O | • CO2 activation mode: the acid–base interaction between nitrogen atoms in porphyrins and CO2, and the strong interaction between the porphyrin π system and CO2 | 46 |
• Both Mg and Cl− activate epichlorohydrins | ||||
![]() |
1,3,5-Triazine, bipyridine, and secondary amine group | ZnBr2 | • CO2 activation mode: the acid–base interaction between alkaline N sites and CO2 | 49 |
• Micropores and abundant nitrogen sites selectivity adsorption for CO2 | ||||
• Mesopores is to facilitate the rapid diffusion of reactants | ||||
• Both Zn and Br− activate epoxides | ||||
![]() |
Carbon nanocomposite, NHC and Br− | Zn(OAc)2·2H2O | • CO2 activation mode: the acid–base interactions between CO2 and NHCs and N-basic sites, as well as a strong interaction between carbon nanocomposites and CO2 | 55 |
• Both Zn and Br− activate epoxides | ||||
![]() |
NHC | Cu(OAc)2 | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 59 |
• Cu activates epoxides | ||||
![]() |
NHC and Br− | AgNO3 | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 60 |
• Both DBU and Ag activate propargylic alcohols | ||||
![]() |
NHC and aniline | Ag2O | • CO2 activation mode: the acid–base interactions of CO2 with NHCs and anilines, respectively | 61 |
• Ag activates propargylamines | ||||
![]() |
NHC and Cl− | Pd(PPh3)4 | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 62 |
• Pd activates both iodobenzenes and propargylamines | ||||
![]() |
NHC and HCO3− | AgNO3 | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 63 |
• Ag activates propargylamines | ||||
• HCO3− promotes NHC–CO2 formation | ||||
![]() |
||||
Porous polymer catalyst without metal loading | ||||
![]() |
NHC and Br− | ZnBr as the additive | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 56 |
• Br− activates epoxides | ||||
![]() |
NHC and Br− | — | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 57 |
• Br− activates epoxides | ||||
![]() |
NHC and OH− | — | • CO2 activation mode: the acid–base interaction between NHCs and CO2 | 66 |
Catalysts | Conv. or yield/% | Selectivity/% | TOF | Ref. |
---|---|---|---|---|
![]() |
||||
Co/POM-TPP | 45.4 | 88.7 | 4.4 | 48 |
POP–PBNCI–TPPMg | 95.1 | 73.8 | 19.8 | 46 |
Bp-Zn@MA | 11–99 | >99 | 21–1580 | 49 |
PIL-Br1.0@Zn-CTM | 51–99 | ≥99 | 39–72 | 55 |
HIP-Br-2 | 85–96 | >99 | 0.09–0.2 | 56 |
HCP | 85–96 | 99 | 136–240 | 57 |
![]() |
||||
![]() |
||||
Cu@NHC-1 | 56–94 | — | 4–9.6 | 59 |
POP-6@Ag2O | 78–92 | — | — | 61 |
POP-1-Pd | 54–67 | — | 5.5–6.8 | 62 |
Ag@POP-HCO3 | 62–81 | — | 81–106 | 63 |
![]() |
||||
![]() |
||||
Ag@POP-NL-3 | 85–95 | — | 1.6–1.8 | 60 |
![]() |
||||
![]() |
||||
IPOP-3 | 62–90 | — | 0.3–0.4 | 66 |
(1) The mechanism of enrichment and activation of waste CO2 by porous polymer catalysts is as follows: (i) acid–base interactions between active sites and CO2 can alter the electron distribution of CO2; (ii) interactions between specific polymer structures and CO2, such as strong interactions observed between conjugated macrocyclic structures and CO2; (iii) the hierarchical pore structure of the polymer (primarily micropores) enables CO2 enrichment, increasing its concentration near catalytic active sites. However, the mechanism of CO2 enrichment by the pore structure of polymers remains unclear. For mixed gases, the hierarchical pore structure of polymers may enrich each gas molecule. Therefore, further research is required to elucidate the enrichment mechanism of waste CO2 by the pore structure.
(2) The BET surface area and total pore volume can be modulated by adjusting the ratio of crosslinking agent to ligand. However, no direct relationship is observed between the BET surface area or total pore volume of the polymer and the conversion of low-concentration CO2 catalysed by the polymer. Accurately preparing polymer catalysts with specific surface area and pore volume tailored to the size of CO2 would be more effective for enriching and converting waste CO2. In addition, the principle underlying the regulation of BET surface area and total pore volume through the addition of anions requires further investigation.
(3) Some reports using gram-scale reactions indicate that porous polymer catalysts may catalyse the large-scale conversion of waste CO2. The development of continuous flow devices demonstrates the feasibility of introducing industrial waste gas directly into the catalytic system for enrichment and conversion of waste CO2. To our knowledge, no published studies have demonstrated the use of polymeric catalysts for industrial-scale CO2 separation and conversion, nor have the economic viability and technical feasibility of such processes been systematically evaluated. Therefore, further research is required to transition catalysts from laboratory to industrial production. Further development of milder and more efficient porous polymer catalysts is required. Sufficient attention should also be given to reactor design and optimisation.
(4) Traditionally, the discovery of catalysts for CO2 absorption and conversion involves a sequential process of polymer synthesis, catalyst optimisation and extensive data collection. Nevertheless, this empirical trial-and-error approach is not only labour-intensive, but also inherently inefficient. In the context of the rapid advancement of artificial intelligence, data-driven machine learning (ML) approaches can now be used to discover high-performance catalysts through computational prediction, thus eliminating the need for extensive experimental procedures.67 Therefore, implementing ML-based predictive modelling for catalyst optimisation can significantly reduce the time taken to develop porous polymer catalysts, while ensuring enhanced catalytic efficacy. On the other hand, ML shows great promise in characterising the surface architectures of heterogeneous catalysts using computational modelling techniques such as density functional theory, molecular dynamics and Monte Carlo simulations.68 As previously established, microstructural parameters of porous polymer catalysts, including pore size distribution, total pore volume and BET-specific surface area, affect the performance of CO2 capture and enrichment. Consequently, compiling a comprehensive database of these surface characteristics from catalysts reported in the literature is essential for the ML-driven optimisation of porous polymer parameters.
(5) The adsorption outcomes of porous polymer catalysts on waste CO2 have been detected using FT-IR and Raman spectroscopy.69 Nevertheless, the dynamic adsorption and conversion process of CO2 on polymer surfaces remains to be elucidated. In order to facilitate more intuitive detection of the conversion process of low concentration CO2, there is a critical research frontier in the development of advanced in situ spectroscopic methodologies capable of real-time monitoring of both selective adsorption and activation phenomena.
(6) In recent years, the development of bifunctional materials (BFMs) for capturing and converting CO2 has emerged as a significant research area. Typically, BFMs comprise two phases: a catalyst phase and a high-temperature adsorbent phase. Meanwhile, porous polymer catalysts have excellent thermal stability (300–750 °C, Table 1), making them compatible with high-temperature adsorption.70 Therefore, the integration of porous polymer catalysts with high-temperature adsorption phases, such as CaO, to prepare novel BFMs for CO2 adsorption and conversion is also worth studying. The good adsorption capabilities of porous polymers for waste CO2 may facilitate the capture and conversion of CO2 by BFMs, obviating the necessity for elevated temperature conditions.
(7) Polymer-based membranes exhibit significant advantages in manufacturability and functional versatility, enabling their widespread adoption in liquid/gas separation processes.71 Notably, porous organic polymers offer superior design flexibility due to their tunable wettability, chemical stability, and facile modification capabilities.72 These attributes position porous polymer catalysts as promising alternatives to conventional membrane materials—including polydimethylsiloxane, polyvinylidene fluoride, poly(vinyl alcohol), and polyethylene glycol laurate—with demonstrated potential to enhance performance metrics in critical applications ranging from gas separation and osmotic processes to advanced filtration and solvent recovery systems. On the other hand, millimetre-sized composite beads are prepared using a double cross-linking strategy, combining powdered, porous polymer catalysts with additives such as polyacrylic acid, CaCl2 and sodium alginate.73 The assembly of fixed-bed catalytic systems utilizing these composite beads as structured packing media has evolved into a burgeoning research focus, particularly for enabling continuous-flow transformation of reactants with enhanced mass transfer efficiency. For gas-phase reaction systems targeting waste CO2 valorization, continuous gas feeding into the flow reactor configuration enables simultaneous concentration enhancement and catalytic conversion through optimized gas–liquid–solid interfacial contact.
We hope that this review provides a new perspective on the enrichment and conversion of waste CO2 using porous polymer catalysts and offers direction for the design and synthesis of heterogeneous catalysts that can facilitate more efficient conversion of waste CO2.
This journal is © The Royal Society of Chemistry 2025 |