Next Article in Journal
Microstructure and Properties of Nickel-Based Gradient Coatings Prepared Using Cold Spraying Combined with Laser Cladding Methods
Next Article in Special Issue
Rheology of Recycled PET
Previous Article in Journal
Wooden Solar Evaporator Design Based on the Water Transpiration Principle of Trees
Previous Article in Special Issue
Effect of Polydopamine Coating of Cellulose Nanocrystals on Performance of PCL/PLA Bio-Nanocomposites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Porous Polymer Materials for CO2 Capture and Electrocatalytic Reduction

1
Key Laboratory for Green Chemical Technology of Ministry of Education, Haihe Laboratory of Sustainable Chemical Transformations, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
2
State Key Laboratory of Physical Chemistry of Solid Surfaces, Collaborative Innovation Center of Chemistry for Energy Materials, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China
3
Department of Mechanical Engineering, Research Institute for Smart Energy, The Hong Kong Polytechnic University, Hong Kong, China
4
Graphene Composite Research Center, College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518060, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(4), 1630; https://doi.org/10.3390/ma16041630
Submission received: 5 January 2023 / Revised: 8 February 2023 / Accepted: 9 February 2023 / Published: 15 February 2023
(This article belongs to the Special Issue Advanced Polymeric Materials: Synthesis, Properties, and Applications)

Abstract

:
Efficient capture of CO2 and its conversion into other high value-added compounds by electrochemical methods is an effective way to reduce excess CO2 in the atmosphere. Porous polymeric materials hold great promise for selective adsorption and electrocatalytic reduction of CO2 due to their high specific surface area, tunable porosity, structural diversity, and chemical stability. Here, we review recent research advances in this field, including design of porous organic polymers (POPs), porous coordination polymers (PCPs), covalent organic frameworks (COFs), and functional nitrogen-containing polymers for capture and electrocatalytic reduction of CO2. In addition, key issues and prospects for the optimal design of porous polymers for future development are elucidated. This review is expected to shed new light on the development of advanced porous polymer electrocatalysts for efficient CO2 reduction.

Graphical Abstract

1. Introduction

The usage of fossil energy has enabled human society to prosper, but it has also led to serious environmental problems [1]. Since the burning of coal, petroleum, and natural gas, the CO2 fixed by plants from nature through millions of years of photosynthesis has been released by humans in just a few hundred years [2]. Moreover, the greenhouse effect caused by excessive CO2 emissions contributes to the climate crisis, sea level rise, and ocean acidity increase [3,4,5]. Up to now, “zero carbon” or “carbon neutral” climate goals have been set up in more than 130 countries or regions [6]. In order to achieve this goal, research work focuses on finding clean energy alternatives to fossil fuels to reduce CO2 emissions, and developing technologies such as CO2 capture, storage, utilization, and chemical conversion to reduce the concentration of CO2 in the air [7,8,9].
The International Energy Agency predicted that CO2 capture and storage (CCS) can reduce CO2 emissions by about 14% annually, which is an important green “bridging” strategy in greenhouse gas reduction [10]. CO2 capture primarily includes pre-combustion, oxyfuel combustion, and post-combustion. CO2 sequestration is the last link of the CCS chain, which focuses on marine and geological sequestration and mineral carbonation [11]. In addition, the research focus of CO2 capture lies on post-combustion technology, which can be applied to the capture of low-pressure CO2 in flue gases [12]. In industry, CO2 in mixed exhaust gases can be effectively captured by amine solutions such as monoethanolamine (MEA) and diethanolamine (DEA). However, the high degree of toxicity, corrosion risk, as well as intense energy consumption hinder their wide applications [13]. It is necessary to develop environmental-friendly CO2 capture materials when considering people’s health and gradually serious environmental problems [14,15]. Morphology, functionality, structure stability, and pore structure are key parameters that determine the efficiency of an adsorbent. In particular, the size of the pore regulates the physical/chemical interaction between the pore and the gas. Porous materials with a pore size closer to the kinetic diameter of CO2 have a higher CO2 affinity. Typically, microporous materials show excellent adsorption performance at lower CO2 pressures [16], while mesoporous materials perform better for CO2 adsorption at higher pressures [17]. The equilibrium adsorption capacity of adsorbent varies with pressure or temperature. In industry, CO2 regeneration is mainly achieved by reducing pressure or increasing temperature [18,19].
From another perspective, CO2 is a low-cost, atoxic, and renewable carbon resource. So, the achievement of CO2 capture and conversion not only mitigates the greenhouse effect but also reduces human industry’s dependence on fossil fuels. However, CO2 has a low standard Gibbs free energy ΔG0 = −394.38 kJ mol−1. Moreover, CO2 shows kinetic inertia and thermodynamic stability, which make it difficult to be activated [20]. Therefore, the transformation of CO2 generally requires reactions with high-energy molecules (e.g., H2, epoxides, unsaturated complexes, organometallic complexes) or relies on external energy (e.g., thermal energy, solar energy, or electricity). In the above strategy, electrochemical technology enables an efficient conversion between electrical and chemical energy. CO2 electrocatalytic reduction is valued due to its mild reaction conditions and renewable clean energy source [21]. Developing novel electrocatalysts, atomic engineering at the catalysts’ surface interface and increasing the specific surface area all contribute to the boosting the catalytic activity of the electrocatalysts [22,23,24]. In addition, a large number of metallic (e.g., metals, alloys, metal oxides, metal carbides, metal sulfides, etc.), non-metallic (e.g., carbon nanotubes, graphene, polymers), and combinations of them served as efficient electrocatalysts for different electrocatalytic reactions [25,26,27]. In previous studies, porous materials have been shown to have more accessible catalytic sites, larger specific surface areas, rich structural designs, and suitable constraints on active species for superior catalytic performance compared to non-porous materials. Therefore, porous materials have the potential to be applied to CO2 capture and electrocatalysis (Scheme 1). Moreover, the functional modules of CO2 capture and electrocatalytic reduction can be integrated in a single system, where the porous material adsorbs CO2 and subsequently completes the catalytic reaction process at the catalytic active site to directly produce a chemical with high added value (e.g., methanol, acetic acid). Recently, interesting research advances have been made in CO2 capture and electrocatalytic reduction, involving porous materials such as zeolites [28], porous carbon materials [29], and porous polymers [30].
Among them, porous polymers are particularly attractive due to their high specific surface area, regulatable porosity, and precisely designed active sites. Some previous reviews have covered the studies of porous polymers in the field of CO2 capture and conversion [31,32], but a thorough overview of the structure–activity relationship, especially the effect of porosity on the capture capacity and catalytic activity of CO2 is still lacking. In this review paper, the porosity regulation in porous polymers and the influence of pore sizes on CO2 capture and electrocatalytic reduction are comprehensively summarized. Special attention will be paid to the interaction between pore structure with CO2 and reaction intermediates. Porous polymers as CO2 adsorbents and catalysts for CO2 electrocatalytic reduction will be discussed in this paper. Moreover, the existing limitations, the challenges in porous polymer design, and the future research direction have also been prospected.

2. Porous Polymers for CO2 Capture and Electrocatalytic Reduction

Porous polymer materials used for CO2 capture and electrocatalytic reduction are mainly divided into four categories, namely, porous organic polymers (POPs), porous coordination polymers (PCPs), covalent organic frameworks (COFs), and nitrogen-containing polymers (N-polymers). Such porous polymeric materials can be prepared through a variety of synthetic methods, including solvothermal, ionothermal, microwave, mechanochemical, and interfacial synthesis methods. Among them, the solvothermal method is the most common preparation method [33]. The advantage of the solvent thermal method is that it can improve the synthesis efficiency by changing the solvent, but the biggest problem is that the solvent used is mostly organic solvent, which will cause environmental pollution when used in large quantities, and the reaction conditions are harsh, so it is not suitable for large-scale production. The ionothermal method is mainly used for the synthesis of covalent triazine frameworks (CTFs), and the reaction conditions required are more severe than those of the solvent thermal method, so the ionothermal method is not suitable for industrial applications [34]. The microwave method can effectively reduce the reaction time and maintain the porosity by microwave, which is a promising synthesis method [35]. The mechanochemical method does not require solvents and therefore avoids environmental pollution and is a good green synthesis method because of the simplicity of the process [36].
The CO2 Reduction Reaction (CO2RR) generally takes three steps: CO2 adsorption, electrocatalytic reduction, and desorption. To be specific, gaseous CO2 molecules are first dissolved in the solvent and form a CO2-saturated electrolyte. Then, the dissolved CO2 in the solvent reacts on the electrode surface. Finally, the reaction products are desorbed from the electrode surface to complete the whole reaction process. Because of the large number of products and complex intermediates in multi-step reactions of CO2RR, definite reaction mechanisms remain unclear. Currently reported CO2RRs occurring in aqueous solutions are summarized in Table 1. From the reaction equations in the table, it can be seen that CO2 can undergo different degrees of electrochemical reduction reactions by transferring different numbers of electrons, and the products include HCOOH, CH3OH, CO, CH4, CH3COOH, C2H4, CH3CH2OH, C2H6, etc. Additionally, the reduction of CO2 is accompanied by HER on the cathode.

2.1. Porous Organic Polymers (POPs)

Porous organic polymers (POPs) are 2D or 3D network structures expanded by organic groups or fragments generated through chemical polymerization [37,38]. With a large specific surface area, variable porosity, excellent thermal stability, as well as controllable functionality, POPs are widely applied in gas adsorption, heterogeneous catalysis, energy storage, and other fields [36,39,40,41,42]. A large specific surface area and abundant porosity lead to high CO2 absorption capacity, bringing out the importance of pore regulation in the design of CO2 capture materials (Table 2).
In recent years, many studies have focused on POPs serving as hosts or heteroatom catalysts to capture and convert CO2 [31,43,44,45]. Han et al. [46] obtained 1,3-dialkyne linked porous polymer LKK-CMP-1 through 1,3,6,8-Tetraethynylpyrene homopolymerization for CO2 capture. The N2 adsorption/desorption isotherm for LKK-CMP-1 is a typical I-type, indicating a microporous structure (Figure 1a). The pore size of LKK-CMP-1 is mainly distributed around 0.59 nm (Figure 1b), close to the dynamic diameter of CO2 (0.33 nm). The CO2 isosteric heat (Qst) for LKK-CMP-1 was 35.0 kJ mol−1 (Figure 1c), which proves CO2 molecules have a strong interaction with LKK-CMP-1 pore walls. The CO2 adsorption capacity can reach 9.78 wt% at 273 K (1 bar). Song et al. [47] tuned the size of micropores through side-chain engineering. Crosslinkable dianhydride monomers PEPHADA and PEQDA were utilized to construct microporous networks (Figure 1d). The crosslinkable alkyne group on the PDQDA side branch is closer to the molecular skeleton, which adjusts the pore size of porous polymers to 0.57 nm (Figure 1e). Therefore, the pore size of crosslinked polyimide can be precisely regulated by adjusting the position of the crosslinkable group in polyimide. The 6FA-PE-CL with well-developed micropores shows a higher surface area and stronger CO2 uptake ability. At 273 K and 1 bar, the CO2 uptake ability of 6FA-PE-CL is 8.9 wt% (Figure 1f). 6FA-PE-CL exhibits a significant Qst value downward trend compared to 6FA-PEPH-CL with the increase of CO2 adsorption, which indicates that there is a stronger interaction between ultramicropores (0.56 nm) and CO2 molecules.
Table 2. The textural features and CO2 adsorption characteristics of reported POPs.
Table 2. The textural features and CO2 adsorption characteristics of reported POPs.
POPsSBET
(m2 g−1)
Vtotal
(cm3 g−1)
CO2 Uptake
(mmol g−1)
273 K
CO2 Qst
(kJ mol−1 )
CO2/N2
Selectivity
273 K
Ref
isox-CTF-50016830.704.922929[43]
acac-CTF-50015561.203.302846[43]
cCTFs-4007440.362.8649-[45]
cCTFs-4508610.592.2546-[45]
cCTFs-50012471.043.0243-[45]
LKK-CMP-1467-2.2435.044.2[46]
6FA-PEPH-CL6530.371.65-38.0[47]
6FA-PE-CL6980.462.02-58.0[47]
TTF-112340.592.86--[48]
POP is an ideal porous structure, which can be applied for the electrocatalytic reduction of CO2 through functional design. Dai et al. [48] reported a porous triazine linking framework TTF-1 with a pyridine linker as the reactive site (Figure 2a). The existence of a well-developed porous structure enables a large contact area between electrodes and electrolytes, and facilitates the charge and mass transfer (Figure 2b). Amorphous porous polymer POP often shows a broad pore size distribution, which is attributed to the inter-particle void. In particular, a large number of micropores promote the adsorption of CO2. The permanent CO2 uptake of TTF-1 is up to 64.4 cm3 g−1 (298 K, 1 bar). TTF-1 also shows excellent electrocatalytic activity for CO2 reduction, with a Faraday efficiency of about 82%.
The size of the pore can affect the charge density of the polymer and the confinement of the pore to the active site. Tang et al. [49] synthesized a series of porous polymers, POP-Py (n) (n means the quantity of benzene rings), through the reaction of benzyl bromide and pyridine. POP-Py (n) with pore sizes from micropore to mesopore were obtained by changing the amount of benzene rings (Figure 3a). The highly efficient electrocatalyst was obtained by encapsulating cobalt meso-tetra(4-carboxyphenyl)porphyrin (CoTCPP) (Figure 3b) into POP-Py (n) by the ion exchange method, with single-molecule dispersion. POP-Py (0)/CoTCPP showed the highest FEco of 83% at −0.6 V vs. RHE, which is twice as high as the activity of the pure CoTCPP molecular electrocatalytic conversion of CO2 to CO (Figure 3c). The high activity is caused by the monomolecular dispersion of CoTCPP in the pores of POP-Py (0), which promotes the electrochemical accessibility of active sites. Moreover, the positively charged substituted pyridine ring in POP-Py can stabilize the key reaction intermediate *COOH to accelerate the conversion of CO2. When a neutral benzyl ring is inserted between two pyridine rings, the positive charge on the pyridine ring will be delocalized through the conjugation effect. Hence, the interaction between POP-Py and key reaction intermediate *COOH becomes weaker. In addition, at a high overpotential, the current attenuation of POPy-Py(2) indicated the aggregation of CoTCPP (Figure 3d).
However, many POPs have poor electrical conductivity, which hinders their application in electrochemistry. Alkordi et al. [50] homogenously deposited porous organic polymer containing pyrimidine atop a graphene sheet to provide composite material PyPOP@G (Figure 4a). It shows significant CO2 electrochemical reduction activity (5 mA cm−2 at −1.6 V), which is much higher than the separated component. PyPOP has an intentional affinity for CO2 due to its rich microporous structure (Figure 4b), while graphene enhances the electrical conductivity of the composite. The synergistic effect of the two components promotes the electrocatalytic reduction of CO2. Later, Alkordi et al. [51] prepared a composite material named POP@MWCNT by synthesizing POP on multi-walled carbon nanotubes in a one-pot method. It was verified by the N2 physical adsorption/desorption test that the high porosity of POP@MWCNT is mainly contributed to by POP, not MWCNTs (Figure 4c). The Qst of CO2 into the POP and the POP@G are almost the same, which confirms that the adsorption of CO2 by the composites mainly occurs in the pores of POP (Figure 4d). Alkordi et al. proposed post-synthetic modification of the electrode material by initial impregnation in CuBr acetonitrile solution. Impregnation of Cu ions enhances electrode catalytic activity. The current density for the POP-Cu@MWCNTs was enhanced by 75% compared to POP@MWCNTs at −1.5 V (Figure 4e).

2.2. Porous Coordination Polymers (PCPs)

Porous coordination polymers (PCPs), also known as metal-organic frameworks (MOFs), are crystalline materials. Such materials are composed of metal ions or metal particles with organic complexes through coordination linkages. PCPs have now become a significant category of nanoporous polymers with multiple uses in CO2 storage, CO2 separation, and CO2 electrocatalysis due to their organized porous structures with large surface areas and adjustable pore surface features [52,53,54,55,56]. The catalytic performance of PCPs catalysts in CO2 electroreduction is summarized in Table 3. Sun et al. [57] reported a two-dimensional bismuth metal-organic framework (Bi-MOF), a catalyst with a microporous structure, for the electrocatalytic reduction of CO2 to HCOOH. They prepared Bi-MOFs with tunable grain size and coordination environment by a simple solvothermal method, varying the reaction temperature and the type of organic linker. Figure 5d shows the crystal structure of a Bi-MOF (CAU-17) composed of a triple base linker (1,3,5-benzene tricarboxylic acid, H3BTC). Figure 5a shows a representative type-I N2 adsorption/desorption isotherm of Bi-MOF, and after measuring it was found that the average pore size was about 0.8 nm (inset of Figure 5a), indicating a remarkable microporous structure. In particular, the obtained Bi-MOF has a CO2 adsorption capacity of 33.0 mgCO2gcat−1, which is twice the CO2 adsorption capacity of Bi sheets under the same testing conditions (Figure 5b). The superior CO2 adsorption capacity of Bi-MOF is attributed to the large number of differently shaped channels in its structure, which are larger in size than the molecular diameter of CO2. The excellent CO2 uptake capacity promotes the adsorption of large amounts of CO2 on the surface of Bi-MOF catalysts, thus potentially contributing to higher catalytic conversion. This Bi-MOF shows large Faraday efficiency of HCOOH over a wide potential window, reaching 92.2% at −0.9 V (vs. RHE), and excellent stability over 30 h (Figure 5c).
The same microporous structure is a nickel phthalocyanine-based conductive MOF (NiPc-Ni(NH)4) constructed from 2,3,9,10,16,17,23,24-octaaminonickel(II) (NiPc-(NH2)8) as reported by Cao et al. [58]. In this research, Ni-phthalocyanines (NiPc), which was built block-based, with an activated CO2 molecular reduction site, were incorporated into a porous essentially conductive 2D MOF NiPc-Ni(NH)4 for the efficient CO2RR towards the CO2 conversion in an aqueous medium. NiPc-Ni(NH)4 showed a 1.78 nm one-dimensional rectangular channel while running along the c-direction with a distance of 0.34 nm between the stacked two-dimensional conjugated layers in the structural model (Figure 6a). Therefore, the N atoms of the amino groups in NiPc-(NH2)8 ligands were managed to coordinate with the Ni(II) ion, forming a fully π-conjugated two-dimensional backbone in NiPc-Ni(NH)4. The accessible pore size of ~1.7 nm was obtained by N2 adsorption/desorption isotherms (Figure 6b), and this large microporous structure facilitates the diffusion and mass transfer of reactants to improve the activity of electrocatalytic reactions. Similarly, the porous NiPc-Ni(NH)4 has a great CO2 adsorption capacity (Figure 6c), which indicates that NiPc-Ni(NH)4 with a nitrogen-rich structure has a strong affinity for CO2 to enhance its electrocatalytic activity in CO2RR. It can be seen that NiPc-Ni(NH)4 nanosheets reach a high CO Faraday efficiency of 96.4% at −0.7V (vs. RHE).
Among PCP materials, not only are microporous structures often seen, but also mesoporous structures are common for CO2RR catalysts. Gu et al. [59] prepared a series of imidazole (Im)-based zeolite imidazolium salt backbone Ni(Im)2 nanosheets of different thicknesses using the liquid phase top-down exfoliation method (Figure 7a). Zeolite imidazole frameworks (ZIFs) are a relatively stable class of MOF materials consisting of imidazole groups and metal ions. They stripped layered massive Ni(Im)2 ZIF to 2D Ni(Im)2 nanosheets of different thicknesses (5 nm, 15 nm, 65 nm, and 140 nm) and used them as CO2RR electrochemical catalysts. As can be seen in Figure 7b, their N2 adsorption/resolution isotherms are all type VI, which corresponds to a mesoporous structure. It provides a large specific surface area for the catalysts while more reactive sites are exposed, thus ensuring the high catalytic activity of this material. This catalyst was found to achieve 78.8% CO Faraday efficiency at −0.9 V (vs. RHE) with a stability of more than 14 h (Figure 7c).
Because a single porous structure is not sufficient for high performance in CO2 capture and electrocatalytic reduction, some researchers have attempted to create hierarchical porous materials through pore engineering strategies [60,61,62,63]. In recent years, the field of MOF has made great progress in the development of hierarchically porous materials [64,65,66,67]. In such hierarchical porous polymers, the inherent micropores of the framework contribute to a high surface area, while the mesopores and/or macropores provide channels for mass transfer and accommodate molecules larger than the micropores [68]. Martin et al. [69] reported the deposition of porous metal-organic skeletons on copper electrodes by electrosynthesis of ionic liquid templates [Cu2(L)] [H4L = 4,4′,4″,4‴-(1,4-phenylenebis(pyridine-4,2,6-triyl))tetrabenzoic acid]. The structure of Cu2(L) comprises a 3D network built around binuclear [Cu2(OOCR)4] paddlewheels with four bridging carboxylate ligands (Figure 8a,b). The porosity of this MOF was investigated by the N2 adsorption/desorption test at 77 K, and Cu2(L)-e exhibited a distribution between type I and type IV, indicating the presence of both mesopores and micropores. The distribution of micropores and mesopores in both materials was analyzed by Horvath Kawazoe and Barrett–Joyner–Halenda (BJH) methods, respectively (Figure 8c,d). By contrast, the reduction of micropores in Cu2(L)-e resulted in lower N2 uptake at low pressure, while the total pore volume of Cu2(R)-e produced by the template effect of the ionic liquid during electrolytic synthesis was 1.89 cm3 g−1, significantly larger than that of Cu2(L)-t (0.32 cm3 g−1), reflecting the presence of Cu2(L)-e-mediated pores. The presence of mesopores also facilitates the mass transfer process during the reaction. As a result, this electrode showed excellent activity for the electroreduction of CO2 to formic acid with a low starting potential of −1.45 V (vs. Ag/AgCl) and HCOOH Faraday efficiency reaching 90.5% at −1.80 V (vs. Ag/AgCl). Despite a similar micro- and mesoporous structure, the role of pore structure in the materials reported by Lan et al. [70] is quite different. The specific surface area of the material decreases instead after the introduction of PPy in the material, which is due to the encapsulation of PPy in the pores of MOF-545-Co. The formation of PPy occurs in the MOF channel, and after the formation of PPy, there will be a π-π interaction between the PPy molecules and TCPP molecules in the MOF structure similar to the host–object interaction, and this interaction will facilitate the charge transfer process during the reaction. Meanwhile, the results of CO2 adsorption measurements showed that the addition of PPy was quite beneficial for CO2 adsorption. As a result, the CO2 reduction efficiency is significantly enhanced.
Table 3. Summary of catalytic performance of PCP catalysts in CO2 electroreduction.
Table 3. Summary of catalytic performance of PCP catalysts in CO2 electroreduction.
CatalystsE/V vs. RHEMajor ProductsFE/%StructureRef.
Bi-MOF−0.9HCOOH92.2microporous[57]
NiPc-Ni(NH)4−0.7CO96.4microporous[58]
2D Ni(Im)2-5 nm−0.9CO78.8mesoporous[59]
Cu2(L)-e/Cu−1.2HCOOH90.5micro- and mesoporous[69]
PPy@MOF-545-Co−0.8CO98micro- and mesoporous[70]
MOF-NS-Cu−0.6HCOOH83.1mesoporous[71]
MOF-NS-Co−0.6CO98.7mesoporous[71]
PCN-222(Cu)−0.7HCOOH44.3mesoporous[72]
PCN-224(Cu)−0.7HCOOH34.1microporous[72]
Al2(OH)2TCPP-Co−0.7CO76mesoporous[73]
PcCu-O8-Zn−0.7CO88microporous mesoporous[74]
PCN-224-NH3−0.9CO60microporous[75]
ZIF-A-LD−1.1CO90.57microporous[76]
D-P-CoPc−0.6CO97macroporous[77]
Cu30%ZIF-8−1.0CH435.21microporous mesoporous[78]
p-CuL-4−0.37CH3COOH64microporous[79]

2.3. Covalent Organic Frameworks (COFs)

Covalent organic frameworks (COFs) are a class of reticulated porous polymeric materials in which the constructed organic blocks are connected by covalent bonds. COFs are characterized by their tunable porosity, which gives them a great accessible surface area and an abundance of active sites. It is due to this property that they have also been extensively studied in catalytic and energy storage systems [80,81]. In recent years, COFs have been widely recognized by researchers as an excellent material that can be used for electrocatalytic CO2 reduction reactions due to their rich pore structure that can facilitate the adsorption and desorption of gas molecules during the reaction. By consciously introducing functional units such as bipyridyl, porphyrin, and phthalocyanine in the COF backbone to link metal ions for electrocatalytic CO2 reduction reactions, they have been widely explored in CO2 reduction reactions [82,83,84].
Recently, fluorinated COFs have also been effective catalysts for electrocatalysis. CTFs are a class of COF materials consisting of three elements, C, N, and H, and aromatic triazine units connected by covalent bonds. Since CTFs usually have a large specific surface area, high nitrogen content, excellent stability, and electrical conductivity, CTFs have a wide range of applications in gas separation and storage, energy storage, and electrocatalysis [85,86,87,88,89,90]. We summarize the catalytic performance of COF catalysts in CO2 electroreduction in Table 4. Dai et al. [91] synthesized a fluorinated covalent triazine skeleton (F-CTF-1) with the low-temperature ionothermal method using 2,3,5,6-tetrafluoroterephthalonitrile as a monomer (Figure 9a,b). The BET specific surface area of this fluorinated COF-containing material was 367 m2 g−1 and mainly microporous (Figure 9c–e). The introduction of fluorine enhanced the hydrophobic and pro-CO2 properties of F-CTF-1 and improved the interaction between F-CTF-1 and CO2 molecules. As an outcome, F-CTF-1 exhibited great Faraday efficiency (95.7%) at −0.8 V (vs. RHE) and was used for the electroreduction of CO2 to CO. Also with a microporous structure, Yaghi and colleagues [92] synthesized a 2D COF (COF-366-Co) in which the cobalt porphyrin site could catalyze the conversion of CO2 to CO by CO2RR. They synthesized a model framework (COF-366-Co) by the imine condensation of 5,10,15,20-tetrakis[(4-aminophenyl)porphinato]cobalt[Co(TAP)] with 1,4-benzenedicarboxaldehyde (BDA). The porous COF material was evacuated by activation with supercritical carbon dioxide and heating to 100 °C for 18 h. The surface area was determined to be 1360 m2 g−1. The adsorption branch fitted using density functional theory (DFT) showed a relatively narrow pore size distribution (10 to 18 Å), consistent with the proposed model. When tested in an aqueous solution, COF-336-Co showed a high Faraday efficiency of 90% Co (FECO), which is 10% higher than that of the molecular cobalt porphyrin catalyst. It is therefore inferred that micropores would allow for a higher CO2 adsorption capacity in the framework, and the greater electrocatalytic activity and chemical availability of the catalytic cobalt porphyrin active site.
It was found that the difference in pore size in similar structures also affects the performance of COF catalysts. Recently, Huang et al. [93] constructed two catalysts, CuPcF8-CoNPc-COF and CuPcF8-CoPc-COF (Figure 10a), with similar pore structures. The adsorption curves of both COFs were classified as typical type I isotherms, which is characteristic of micropores (Figure 10c,d). The surface areas of CuPcF8-CoPc-COF and CuPcF8-CoNPc-COF were determined to be 376 and 452 m2 g−1, respectively. Correspondingly, their pore volumes were calculated to be 0.23 and 0.30 cm3 g−1. Non-local density flooding theory (NLDFT) was used to evaluate their pore sizes of 1.0 and 1.2 nm (Figure 6c,d). Meanwhile, the CO2 capacities of CuPcF8-CoPc-COF and CuPcF8-CoNPc-COF were 18 and 35 mg g−1, respectively. It can be seen that CuPcF8-CoNPc-COF has a larger pore size, specific surface area, and better CO2 capacity compared to CuPcF8-CoPc-COF, which is more favorable to improve the catalytic activity of CO2RR. Thus, when tested in 0.5 M CsHCO3, the FECO of CuPcF8-CoPc-COF reached a peak of 91% at −0.70 V (vs. RHE), while the FECO of CuPcF8-CoNPc-COF was as high as 97% at −0.62 V (vs. RHE).
Similarly, mesoporous structures are also present in COF materials, and Zeng et al. [94] formed COF on the surface of Mg/Al-LDH, and the obtained catalysts showed a layered and porous morphology with a thickness of about 36 nm and an abundance of active sites. The N2 adsorption isotherm at 77 K showed a BET surface area of 123 m2 g−1 for 2D-Co-COF500 (Figure 11a). The pore size distribution plot shows that the pore volume of 2D-CO-COF500 is 0.52 cm3 g−1, where most of the pores are between 25 and 80 nm in diameter (Figure 11b). Electrochemical tests were then performed on this material. The FE of CO was found to be 66.1% at −0.5 V, 84.0% at −0.6 V, 92.1% at −0.7 V, 96.5% at −0.8 V, 96.0 at −0.9 V, and finally 80.2% at −1.0 V (Figure 11c), while electrochemical tests on CO-COF500 revealed that the FEs of CO were all at the same test potential. Among them, 2D-Co-COF500 reached a maximum FE of 96.5% for CO at −0.8 V, and the catalytic performance was better than most of the current COF materials. To further investigate the catalyst activity, the CO current density (jCO) was tested. It was found that 2D-CO-COF500 had the highest current density of 17.9 mA cm−2 at a potential of −1.0 V (Figure 11d), while Co-COF500 also reached a maximum current density of 6.8 mA cm−2 at −0.9 V.
For materials with multi-level pore structures with both micro- and mesopores, the large specific surface area and rich pore structure allows them to have a strong CO2 adsorption capacity, enhanced mass transfer, and more exposed reactive sites, ultimately leading to stronger CO2 reduction performance. Zhang et al. [95] used porphyrin-based covalent organic framework (Por-COF) nanosheets vertically anchored to carbon nanotubes (CNT) by covalent linkage for an efficient electrocatalytic CO2 reduction reaction. The porosity structure of Por-COF and MWCNT-Por-COF-M were investigated by the BET test at 77 K. The results indicate that all the samples exhibit type I adsorption isotherms and the pore size distributions are concentrated on 15 Å (Figure 12a,b). The electrochemical test results show that the MWCNT-Por-COF-M (M: Co, Ni, Fe) material has higher electrocatalytic performance and Faraday efficiency than Por-COF-M. Specifically, MWCNT-Por-COF-Co achieves the highest Faraday efficiency of 99.3% for CO at -0.6 V, and the highest current density of 18.77 mA cm−2 at −1.0 V and TOF of 70.6 s−1 (Figure 12c). The electrochemical test results indicate that the excellent electrocatalytic activity of MWCNT-Por-COF-Co is due to the effective electron transfer. In addition, they used MWCNT-Por-COF-Cu in a flow cell for a CO2RR test with electrolytes in 1.0 M KOH. The test results illustrate that MWCNT-Por-COF-Cu has a higher CO2 catalytic efficiency than MWCNT-Por-COF, the product is mainly CH4, and the Faraday efficiency can reach up to 71.2% (Figure 12d). Sun et al. [96] used tetraanhydrides of 2,3,9,10,16,17,23,24-octacarboxyphthalocyanine cobalt(II) (CoTAPc) as a junction to couple with 5,15-di(4-aminophenyl)-10,20-diphenylporphyrin (DAPor) or 5,15,10,20-tetrayl(4-aminophenyl)porphyrin (TAPor) by imination reaction to fabricate novel coupled phthalocyanine–porphyrin type 1:2 (CoPc-2H2Por) or type 1:1 (CoPc-H2Por) COFs. Both type 1:2 and type 1:1 COFs were found to exhibit over 90% maximum Faraday efficiency and high stability in CO2 electrocatalytic reduction reactions. Meanwhile, CoPc-2H2Por has a larger pore size and a more conjugated structure than CoPc-H2Por, leading to more efficiency in electron transfer, more adsorption and reaction of CO2, and rapid proton transfer during the reaction, which ultimately leads to higher reaction kinetics, resulting in better CO2RR activity than CoPc-H2Por.
Table 4. Summary of catalytic performance of COF catalysts in CO2 electroreduction.
Table 4. Summary of catalytic performance of COF catalysts in CO2 electroreduction.
Catalysts E/V vs. RHEMajor ProductsFE/%StructureRef.
COF-366-Cu (HS)−0.9CH452.4microporous[86]
F-CTF-1-275−0.8CO95.7microporous[87]
COF-366-Co−0.67CO90microporous[92]
CuPcF8-CoNPc-COF−0.62CO97microporous[93]
2D-Co-COF500−0.8CO96.5mesoporous[94]
MWCNT-PorCOF-Cu−1.0CH471.2micro- and mesoporous[95]
CoPc-2H2Por−0.55CO95micro- and mesoporous[96]
AAn-COF-Cu (NF)−0.9CH477microporous[97]
FN-CTF-400−0.8CH491.7microporous[98]
TPE-CoPor-COF−0.7CO95microporous[99]
3D-Por(Co/H)-COF−0.6CO92.4microporous[100]
COF@CoPor−0.6CO94.3microporous[101]
NiPc-TFPN COF−0.9CO99.8microporous[102]
CoPc-PI-COF-3−0.9CO96microporous[103]
COF-366-(OMe)2-Co@CNT−0.68CO93.6macroporous[104]
CTF-Cu−1.47C2H430.6micro- and mesoporous[105]

2.4. Nitrogen-Containing Polymers (N-Polymers)

There is also a class of porous polymeric materials called “metal-loaded ligand-active nitrogen-containing polymers” (N-polymers). This kind of N-polymer is characterized by the presence of nitrogen functional groups that complex metal cations or anchor metal nanoparticles within the polymer backbone (e.g., polyaniline and polypyrrole) or side chains (e.g., polyvinyl terbium pyridine and poly (4-vinyl pyridine)). The application of such materials in the field of CO2 electroreduction, however, dates back to three decades ago, when Wrighton’s [106] and Meyer’s groups [107] introduced N-polymers in the field of non-homogeneous CO2 electroreduction and found that such catalysts have excellent catalytic properties in CO2 conversion. So far, metal-doped N-polymer materials have been tested to synthesize polycarbon products up to propanol.
CO2 adsorption by porous polymers has been used as a competitive strategy for selective CO2 removal/capture [108,109,110]. Due to the rapid development of nitrogen-containing nanoporous polymeric materials, the application of N-polymers in CO2 capture has achieved remarkable success. Nitrogen-rich carbons are of great interest because of the contribution of N as a basic or polar site for effective interaction with CO2 [111,112]. Therefore, the preparation of nitrogen-rich carbon materials for CO2 capture is quite attractive and important in various applications such as CO2 capture. Due to the high porosity and relatively high nitrogen content, the adsorption of polyaniline (PANI) is very effective in both liquid and gas phases [113,114,115]. Park and his colleagues [116] prepared a porous carbon material using melamine and PANI with milder reaction conditions (≤650 °C). The prepared materials (MPCs) were found to have some CO2 adsorption capacity and therefore could be used for CO2 adsorption. It can be seen from Figure 13a that the adsorption capacity of PDCs for N2 increases with the increase in the pyrolysis temperature. Therefore, for N2 adsorption capacity, MPC-500 is the worst while MPC-650 is the best. The pore size distribution (PSD) of the material was also obtained from the isothermal curve shown in Figure 13a and shown in Figure 13b by nonlocal linear density functional theory (NLDFT). The adsorption capacities of the different materials prepared for CO2 and N2 are depicted from Figure 13c. As shown in Figure 13a, the adsorption of CO2 decreases according to the decrease in the pyrolysis temperature of the material. Thus, MPC-600 has the highest CO2 adsorption capacity and MPC-500 has the lowest CO2 adsorption capacity. Among them, MPC-550 showed superior CO2 selectivity, attributing this performance to the increase in the elemental N content of the material, the decrease in porosity, or the decrease in pore size and the increase in defects.
P4VP is an N-heteroaromatic polymer with a mesoporous structure that allows easy diffusion of small molecules and anions through the polymer network [117,118]. Nowadays, this N-polymer has been introduced into the electrochemical reduction of CO2 reactions, and Koper and colleagues [119] showed a significant improvement in the selectivity of polycrystalline Cu and Au electrodes for HCOOH at low overpotentials after chemical modification by using poly(4-vinyl pyridine) (P4VP) layers. Measurements showed that the hydrophobicity of the P4VP layer limited the mass transport of HCO3 and H2O, but had little effect on the mass transport of CO2 due to its mesoporous structure.
Meanwhile, g-C3N4 is also a common N-polymer material with a structure in which the C and N atoms are sp2-hybridized to form a π-type conjugated system with a highly delocalized domain. Due to its high stability no matter the extreme chemical environment or high temperature, proper pore structure, redox potential, and abundant functional groups on the surface, there are also many studies based on g-C3N4 materials reported for energy and environmental applications [120,121], especially for electrocatalytic CO2 reduction [122]. Zhao et al. [123] prepared a Au-CDots-C3N4 material for CO2RR that can efficiently convert CO2 to CO. Electrochemical tests showed that 4 wt% Au-CDots-C3N4 has optimal CO2 electrocatalytic performance, and this material can undergo CO2 electrocatalytic reactions at −0.5 V and the Faraday efficiency of CO reached a maximum of 79.8%. The surface area of 4 wt% Au-CDots-C3N4 was investigated by the BET method at 77K. Figure 14a demonstrates the BET plot of 4 wt% Au-CDots-C3N4, which is consistent with a typical type IV isotherm with an H3-type hysteresis loop, indicating a large number of mesopores in this material. The specific surface area of this material was tested to be 117 m2 g−1, which originates from the rich void structure unique to the C3N4 matrix. Additionally, the pore size distribution obtained by the BJH method test can be shown by Figure 14a. Hu et al. [124] tried to modify Cu and Ru onto g-C3N4 (CuxRuyCN) and use it as an electrode for electrocatalytic CO2 reduction, and it can be seen from the test results that Cu and Ru modification on g-C3N4 can be an effective way to improve its electrocatalytic performance for energy applications. As can be seen in Figure 14a, the surface area of the g-C3N4 sample decreases with increasing Cu content, which indicates that the increased Ru and tiny CuO particles partially block the mesopores. Therefore, this result indicates that the CuxRuyCN samples with proper Cu and Ru ratios show the best BET surface area, pore volume and pore size under optimized Cu and Ru decoration to attract reactant molecules and provide active sites for better electrocatalytic reactions.
To summarize, the influence of the structural characteristics of four different types of porous polymer materials, POPs, PCPs, COFs, and N-polymers, on CO2 capture and electrocatalytic reduction have been reviewed. Their structural features and advantages and disadvantages for CO2 capture and electrocatalytic reduction are further summarized in Table 5.

3. Conclusions and Perspectives

Global climate change caused by excessive CO2 emissions has become a major concern in modern society. CO2 capture and conversion can remove excess CO2 from the atmosphere and reduce it into products with high added value. Porous polymers have been applied as potential CO2 adsorbents and electroreduction catalysts due to their high surface area, designable functionalization, tunable porosity, and various reaction sites. This review has systematically summarized the influence of the structural characteristics of different porous materials on CO2 capture and electrocatalytic reduction. Four different types of porous polymers, including POPs, PCPs, COFs, and N-polymers, have been discussed in depth. Porous polymers can serve as a multifunctional platform to achieve the desired purpose. Their high specific surface area and porosity enable them to absorb and store CO2. The design of the structure can not only improve the selective adsorption of CO2, but also allow the porous polymer as a catalyst to capture and convert CO2 into high-value-added chemicals. The adsorption of CO2 on porous polymers mainly depends on pore filling, so the highly developed porous features determine the ability of CO2 capture. It is worth noting that the introduction of heteroatoms can effectively improve the affinity of porous polymers to CO2 compared with other gases. Heteroatoms act as binding sites of CO2 through dipole–quadrupole interaction. In addition, the functionalization of the surface CO2-philic groups helps to further improve the adsorption capacity and selectivity.
The role of pore structures in CO2 adsorption and electroreduction should be particularly emphasized. An abundant variety of building blocks with various functional groups are utilized for porous materials’ synthesis, which allows the pore structure and the chemical environment within the pore to be designed for a specific purpose. The microporous and ultramicroporous structures are considered to be preferred for the adsorption of CO2 molecules. The physical and chemical interactions between pore walls and CO2 molecules can be enhanced by fine-tuning the pore size and structure, which improves the catalytic efficiency of the electrocatalyst. The mesoporous structure is conducive to the increase in the accessibility of catalytic reaction sites, mass transfer, and the enrichment of reactants in the pores, which will accelerate the rate of the CO2 electrocatalytic reduction reaction. The hierarchical porous structure generally consists of a composite of micropores and mesopores. Although the surface area and pore volume of the polymers decrease as the mesopores increase, the disadvantage of single micropores that are not conducive to the diffusion of reactants and products at a single microporous pore size is overcome. Hierarchical porous polymers are promising candidates for expanding the electrocatalyst family. Effective electrocatalytic reactions occur at certain active sites. Therefore, it is of practical significance to create and regulate porous structures in order to expose more catalytic active sites.
Despite the above progresses achieved, there are still some great challenges to overcome in the design and synthesis of porous polymers towards highly efficient CO2 capture and catalytic reduction. Promising future research directions in this field are also prospected as follows.
Firstly, the specific surface area, pore size, and pore morphology of porous polymer materials are tuned by functionalized modification of the polymer backbone to achieve the result of improved catalytic performance, for example, functional elements of a semiconductor nature can be introduced into the backbone structure, thus enabling porous polymer materials to achieve and improve performance in electrocatalysis, energy storage, etc. However, for now, achieving specific tuning of porous polymer materials is still a great challenge. More attention should be paid to boosting the coordination of active sites and pore topology to prevent the separation of metal ions or mixtures while improving the electrocatalytic performance, so as to provide long-term stability. Hierarchical porous structures can be constructed by developing advanced synthetic methods to improve the adsorption and activation of electrocatalysts for CO2. The chemical environment within the pores can be regulated by introducing heteroatoms, substituents, and metal components to stabilize key reaction intermediates and promote the charge transfer properties of porous polymers.
Secondly, the mechanism of porous polymers in the electrocatalytic reduction of CO2 is not well understood, especially for the exploration of the pathway of polycarbonate product generation. In the current study, different reaction intermediates are detected corresponding to different reaction mechanisms, but most of them stay in proposing the mechanism, and the investigation of experimental evidence is not comprehensive. It is difficult to accurately predict the electrocatalytic performance of porous polymers only by theoretical calculation, which limits the development of novel advanced porous polymers for CO2 electrocatalysts. Development of cutting-edge in situ monitoring technologies could bring about promising solutions to this issue. In situ surface characterization techniques, such as XPS, TEM, XRD, and XAS, can provide clear information about the evolution of the active sites of the electrocatalyst and the intermediate species of key reactions, which will deepen the understanding of CO2 capture and catalytic reduction by porous polymers.
Third, the lab-scale research generally utilizes concentrated CO2 for testing, which is quite different from the atmospheric CO2 in practical situations. To facilitate the conversion of basic research achievements into products in industry, atmospheric samples are recommended to be used in basic research. In addition, most of the existing synthetic means are costly and have harsh reaction conditions, making it difficult to achieve low-cost mass production. We should focus on finding environmentally friendly, simple, and inexpensive co-production methods to promote the industrial application of porous polymer materials.
In conclusion, porous polymers are highly efficient CO2 capture and conversion materials with great development potential. There are still severe challenges including lack of accurate design of porous polymers with well-coordinated catalytic active sites and pore topology, insufficient cutting-edge in situ monitoring technologies, as well as impractical testing conditions and high production costs. Through reasonable structure design, in-depth research on electrocatalysis mechanisms, and reduction of production cost, the practicability of porous polymers can be further promoted.

Author Contributions

Conceptualization, H.W.; literature search, H.W. and G.W.; writing—original draft preparation, H.W., G.W., L.H. and B.G.; writing—review and editing, X.Y. and J.D.; supervision, X.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Hong Kong Polytechnic University (ZZLM, YY4V, YY5K).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sebos, I. Fossil fraction of CO2 emissions of biofuels. Carbon Manag. 2022, 13, 154–163. [Google Scholar] [CrossRef]
  2. Kajla, S.; Kumari, R.; Nagi, G.K. Microbial CO2 fixation and biotechnology in reducing industrial CO2 emissions. Arch. Microbiol. 2022, 204, 149. [Google Scholar] [CrossRef] [PubMed]
  3. Burger, F.A.; John, J.G.; Frölicher, T.L. Increase in ocean acidity variability and extremes under increasing atmospheric CO2. Biogeosciences 2020, 17, 4633–4662. [Google Scholar] [CrossRef]
  4. Qu, Y.; Jevrejeva, S.; Jackson, L.P.; Moore, J.C. Coastal Sea level rise around the China Seas. Glob. Planet. Chang. 2019, 172, 454–463. [Google Scholar] [CrossRef]
  5. Malati, M.A. Mitigation of CO2 greenhouse effect. Combined disposal and utilisation by photocatalysis. Energy Convers. Manag. 1996, 37, 1345–1350. [Google Scholar] [CrossRef]
  6. Zhao, Y.; Su, Q.; Li, B.; Zhang, Y.; Wang, X.; Zhao, H.; Guo, S. Have those countries declaring “zero carbon” or “carbon neutral” climate goals achieved carbon emissions-economic growth decoupling? J. Clean. Prod. 2022, 363, 132450. [Google Scholar] [CrossRef]
  7. Sanz-Pérez, E.S.; Murdock, C.R.; Didas, S.A.; Jones, C.W. Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, 11840–11876. [Google Scholar] [CrossRef]
  8. Zhu, P.; Wang, H. High-purity and high-concentration liquid fuels through CO2 electroreduction. Nat. Catal. 2021, 4, 943–951. [Google Scholar] [CrossRef]
  9. Zang, G.; Sun, P.; Yoo, E.; Elgowainy, A.; Bafana, A.; Lee, U.; Wang, M.; Supekar, S. Synthetic Methanol/Fischer–Tropsch Fuel Production Capacity, Cost, and Carbon Intensity Utilizing CO2 from Industrial and Power Plants in the United States. Environ. Sci. Technol. 2021, 55, 7595–7604. [Google Scholar] [CrossRef]
  10. Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.; Xie, B.; Toe, C.Y.; Zhu, X.; Wang, J.; et al. Industrial carbon dioxide capture and utilization: State of the art and future challenges. Chem. Soc. Rev. 2020, 49, 8584–8686. [Google Scholar] [CrossRef]
  11. Aminu, M.D.; Nabavi, S.A.; Rochelle, C.A.; Manovic, V. A review of developments in carbon dioxide storage. Appl. Energy 2017, 208, 1389–1419. [Google Scholar] [CrossRef] [Green Version]
  12. Bhanja, P.; Modak, A.; Bhaumik, A. Porous Organic Polymers for CO2 Storage and Conversion Reactions. ChemCatChem 2018, 11, 244–257. [Google Scholar] [CrossRef]
  13. Wang, S.; Yan, S.; Ma, X.; Gong, J. Recent advances in capture of carbon dioxide using alkali-metal-based oxides. Energy Environ. Sci. 2011, 4, 3805–3819. [Google Scholar] [CrossRef]
  14. Naguib, H.M.; Hou, G. Exploitation of Natural and Recycled Biomass Resources to Get Eco-friendly Polymer. J. Polym. Environ. 2023, 31, 533–540. [Google Scholar] [CrossRef]
  15. Hou, G.; Yan, Z.; Sun, J.; Naguib, H.M.; Lu, B.; Zhang, Z. Microstructure and mechanical properties of CO2-cured steel slag brick in pilot-scale. Constr. Build. Mater. 2021, 271, 121581. [Google Scholar] [CrossRef]
  16. Luo, Y.; Li, B.; Wang, W.; Wu, K.; Tan, B. Hypercrosslinked Aromatic Heterocyclic Microporous Polymers: A New Class of Highly Selective CO2Capturing Materials. Adv. Mater. 2012, 24, 5703–5707. [Google Scholar] [CrossRef]
  17. Singh, G.; Lakhi, K.S.; Kim, I.Y.; Kim, S.; Srivastava, P.; Naidu, R.; Vinu, A. Highly Efficient Method for the Synthesis of Activated Mesoporous Biocarbons with Extremely High Surface Area for High-Pressure CO2 Adsorption. ACS Appl. Mater. Interfaces 2017, 9, 29782–29793. [Google Scholar] [CrossRef]
  18. Cen, Q.; Fang, M.; Wang, T.; Majchrzak-Kucęba, I.; Wawrzyńczak, D.; Luo, Z. Thermodynamics and regeneration studies of CO2 adsorption on activated carbon. Greenh. Gases Sci. Technol. 2016, 6, 787–796. [Google Scholar] [CrossRef]
  19. Elfving, J.; Bajamundi, C.; Kauppinen, J.; Sainio, T. Modelling of equilibrium working capacity of PSA, TSA and TVSA processes for CO2 adsorption under direct air capture conditions. J. CO2 Util. 2017, 22, 270–277. [Google Scholar] [CrossRef]
  20. Álvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef]
  21. Jia, S.; Ma, X.; Sun, X.; Han, B. Electrochemical Transformation of CO2 to Value-Added Chemicals and Fuels. CCS Chem. 2022, 4, 3213–3229. [Google Scholar] [CrossRef]
  22. Popovic, S.; Smiljanic, M.; Jovanovic, P.; Vavra, J.; Buonsanti, R.; Hodnik, N. Stability and Degradation Mechanisms of Copper-Based Catalysts for Electrochemical CO2 Reduction. Angew. Chem. 2020, 59, 14736–14746. [Google Scholar] [CrossRef] [PubMed]
  23. Xiang, Z.; Cao, D.; Huang, L.; Shui, J.; Wang, M.; Dai, L. Nitrogen-Doped Holey Graphitic Carbon from 2D Covalent Organic Polymers for Oxygen Reduction. Adv. Mater. 2014, 26, 3315–3320. [Google Scholar] [CrossRef] [PubMed]
  24. Du, P.; Eisenberg, R. Catalysts made of earth-abundant elements (Co, Ni, Fe) for water splitting: Recent progress and future challenges. Energy Environ. Sci. 2012, 5, 6012–6021. [Google Scholar] [CrossRef]
  25. Wang, Y.-N.; Yang, Z.-J.; Yang, D.-H.; Zhao, L.; Shi, X.-R.; Yang, G.; Han, B.-H. FeCoP2 Nanoparticles Embedded in N and P Co-doped Hierarchically Porous Carbon for Efficient Electrocatalytic Water Splitting. ACS Appl. Mater. Interfaces 2021, 13, 8832–8843. [Google Scholar] [CrossRef]
  26. Cui, H.; Guo, Y.; Zhou, Z. Three-Dimensional Graphene-Based Macrostructures for Electrocatalysis. Small 2021, 17, e2005255. [Google Scholar] [CrossRef]
  27. Menga, D.; Ruiz-Zepeda, F.; Moriau, L.; Šala, M.; Wagner, F.; Koyutürk, B.; Bele, M.; Petek, U.; Hodnik, N.; Gaberšček, M.; et al. Active-Site Imprinting: Preparation of Fe–N–C Catalysts from Zinc Ion–Templated Ionothermal Nitrogen-Doped Carbons. Adv. Energy Mater. 2019, 9, 1902412. [Google Scholar] [CrossRef]
  28. Panda, D.; Kumar, E.A.; Singh, S.K. Introducing mesoporosity in zeolite 4A bodies for Rapid CO2 capture. J. CO2 Util. 2020, 40, 101223. [Google Scholar] [CrossRef]
  29. Zhang, Z.; Cano, Z.P.; Luo, D.; Dou, H.; Yu, A.; Chen, Z. Rational design of tailored porous carbon-based materials for CO2 capture. J. Mater. Chem. A 2019, 7, 20985–21003. [Google Scholar] [CrossRef]
  30. Singh, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2 capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar] [CrossRef]
  31. Yang, D.-H.; Tao, Y.; Ding, X.; Han, B.-H. Porous organic polymers for electrocatalysis. Chem. Soc. Rev. 2022, 51, 761–791. [Google Scholar] [CrossRef]
  32. Song, K.S.; Fritz, P.W.; Coskun, A. Porous organic polymers for CO2 capture, separation and conversion. Chem. Soc. Rev. 2022, 51, 9831–9852. [Google Scholar] [CrossRef]
  33. Côté, A.P.; Benin, A.I.; Ockwig, N.W.; O’Keeffe, M.; Matzger, A.J.; Yaghi, O.M. Porous, Crystalline, Covalent Organic Frameworks. Science 2005, 310, 1166–1170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Chakravarty, C.; Mandal, B.; Sarkar, P. Multifunctionalities of an Azine-Linked Covalent Organic Framework: From Nanoelectronics to Nitroexplosive Detection and Conductance Switching. J. Phys. Chem. C 2018, 122, 3245–3255. [Google Scholar] [CrossRef]
  35. Sick, T.; Hufnagel, A.G.; Kampmann, J.; Kondofersky, I.; Calik, M.; Rotter, J.M.; Evans, A.; Döblinger, M.; Herbert, S.; Peters, K.; et al. Oriented Films of Conjugated 2D Covalent Organic Frameworks as Photocathodes for Water Splitting. J. Am. Chem. Soc. 2018, 140, 2085–2092. [Google Scholar] [CrossRef] [Green Version]
  36. Vardhan, H.; Nafady, A.; Al-Enizi, A.M.; Ma, S. Pore surface engineering of covalent organic frameworks: Structural diversity and applications. Nanoscale 2019, 11, 21679–21708. [Google Scholar] [CrossRef] [PubMed]
  37. Ji, G.; Zhao, Y.; Liu, Z. Design of porous organic polymer catalysts for transformation of carbon dioxide. Green Chem. Eng. 2022, 3, 96–110. [Google Scholar] [CrossRef]
  38. Siegelman, R.L.; Kim, E.J.; Long, J.R. Porous materials for carbon dioxide separations. Nat. Mater. 2021, 20, 1060–1072. [Google Scholar] [CrossRef] [PubMed]
  39. Smith, C.L.; Clowes, R.; Sprick, R.S.; Cooper, A.I.; Cowan, A.J. Metal–organic conjugated microporous polymer containing a carbon dioxide reduction electrocatalyst. Sustain. Energy Fuels 2019, 3, 2990–2994. [Google Scholar] [CrossRef]
  40. Lu, C.; Yang, J.; Wei, S.; Bi, S.; Xia, Y.; Chen, M.; Hou, Y.; Qiu, M.; Yuan, C.; Su, Y.; et al. Atomic Ni Anchored Covalent Triazine Framework as High Efficient Electrocatalyst for Carbon Dioxide Conversion. Adv. Funct. Mater. 2019, 29, 1806884. [Google Scholar] [CrossRef]
  41. Eder, G.M.; Pyles, D.A.; Wolfson, E.R.; McGrier, P.L. A ruthenium porphyrin-based porous organic polymer for the hydrosilylative reduction of CO2 to formate. Chem. Commun. 2019, 55, 7195–7198. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, R.; Wang, X.; Weng, W.; Yao, Y.; Kidkhunthod, P.; Wang, C.; Hou, Y.; Guo, J. Proton/Electron Donors Enhancing Electrocatalytic Activity of Supported Conjugated Microporous Polymers for CO2 Reduction. Angew. Chem. 2022, 61, e202115503. [Google Scholar] [CrossRef]
  43. Jena, H.S.; Krishnaraj, C.; Schmidt, J.; Leus, K.; Van Hecke, K.; Van Der Voort, P. Effect of Building Block Transformation in Covalent Triazine-Based Frameworks for Enhanced CO2 Uptake and Metal-Free Heterogeneous Catalysis. Chemistry 2020, 26, 1548–1557. [Google Scholar] [CrossRef] [Green Version]
  44. Liu, Z.-W.; Han, B.-H. Ionic porous organic polymers for CO2 capture and conversion. Curr. Opin. Green Sustain. Chem. 2019, 16, 20–25. [Google Scholar] [CrossRef]
  45. Buyukcakir, O.; Je, S.H.; Talapaneni, S.N.; Kim, D.; Coskun, A. Charged Covalent Triazine Frameworks for CO2 Capture and Conversion. ACS Appl. Mater. Interfaces 2017, 9, 7209–7216. [Google Scholar] [CrossRef]
  46. Ren, S.-B.; Li, P.-X.; Stephenson, A.; Chen, L.; Briggs, M.E.; Clowes, R.; Alahmed, A.; Li, K.-K.; Jia, W.-P.; Han, D.-M. 1,3-Diyne-Linked Conjugated Microporous Polymer for Selective CO2 Capture. Ind. Eng. Chem. Res. 2018, 57, 9254–9260. [Google Scholar] [CrossRef]
  47. Song, N.; Wang, T.; Ma, T.; Li, J.; Yao, H.; Guan, S. Microporous polyimide networks with tunable micropore size constructed through side-chain engineering of linear precursors. Polymer 2022, 255, 125161. [Google Scholar] [CrossRef]
  48. Zhu, X.; Tian, C.; Wu, H.; He, Y.; He, L.; Wang, H.; Zhuang, X.; Liu, H.; Xia, C.; Dai, S. Pyrolyzed Triazine-Based Nanoporous Frameworks Enable Electrochemical CO2 Reduction in Water. ACS Appl. Mater. Interfaces 2018, 10, 43588–43594. [Google Scholar] [CrossRef] [PubMed]
  49. Tang, J.-K.; Zhu, C.-Y.; Jiang, T.-W.; Wei, L.; Wang, H.; Yu, K.; Yang, C.-L.; Zhang, Y.-B.; Chen, C.; Li, Z.-T.; et al. Anion exchange-induced single-molecule dispersion of cobalt porphyrins in a cationic porous organic polymer for enhanced electrochemical CO2 reduction via secondary-coordination sphere interactions. J. Mater. Chem. A 2020, 8, 18677–18686. [Google Scholar] [CrossRef]
  50. Soliman, A.B.; Haikal, R.R.; Hassan, Y.S.; Alkordi, M.H. The potential of a graphene-supported porous-organic polymer (POP) for CO2 electrocatalytic reduction. Chem. Commun. 2016, 52, 12032–12035. [Google Scholar] [CrossRef]
  51. Haikal, R.R.; Soliman, A.B.; Amin, M.; Karakalos, S.G.; Hassan, Y.S.; Elmansi, A.M.; Hafez, I.H.; Berber, M.R.; Hassanien, A.; Alkordi, M.H. Synergism of carbon nanotubes and porous-organic polymers (POPs) in CO2 fixation: One-pot approach for bottom-up assembly of tunable heterogeneous catalyst. Appl. Catal. B Environ. 2017, 207, 347–357. [Google Scholar] [CrossRef]
  52. Wu, P.; Li, Y.; Zheng, J.-J.; Hosono, N.; Otake, K.-I.; Wang, J.; Liu, Y.; Xia, L.; Jiang, M.; Sakaki, S.; et al. Carbon dioxide capture and efficient fixation in a dynamic porous coordination polymer. Nat. Commun. 2019, 10, 4362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Meng, M.-M.; Xi, J.-M.; Zhu, R.; Wang, L.-F.; Xu, Y.-K.; Liu, X.-G.; Zhang, R.; Lu, Z.-Z. Tubular porous coordination polymer for selective adsorption of CO2. Inorg. Chem. Commun. 2021, 132, 108798. [Google Scholar] [CrossRef]
  54. Inukai, M.; Tamura, M.; Horike, S.; Higuchi, M.; Kitagawa, S.; Nakamura, K. Storage of CO2 into Porous Coordination Polymer Controlled by Molecular Rotor Dynamics. Angew. Chem. 2018, 57, 8687–8690. [Google Scholar] [CrossRef]
  55. Kauffman, K.L.; Culp, J.T.; Allen, A.J.; Espinal, L.; Wong-Ng, W.; Brown, T.D.; Goodman, A.; Bernardo, M.P.; Pancoast, R.J.; Chirdon, D.; et al. Selective Adsorption of CO2 from Light Gas Mixtures by Using a Structurally Dynamic Porous Coordination Polymer. Angew. Chem. 2011, 50, 10888–10892. [Google Scholar] [CrossRef] [Green Version]
  56. Emerson, A.J.; Chahine, A.; Batten, S.R.; Turner, D.R. Synthetic approaches for the incorporation of free amine functionalities in porous coordination polymers for enhanced CO2 sorption. Coord. Chem. Rev. 2018, 365, 1–22. [Google Scholar] [CrossRef]
  57. Li, F.; Gu, G.H.; Choi, C.; Kolla, P.; Hong, S.; Wu, T.-S.; Soo, Y.-L.; Masa, J.; Mukerjee, S.; Jung, Y.; et al. Highly stable two-dimensional bismuth metal-organic frameworks for efficient electrochemical reduction of CO2. Appl. Catal. B Environ. 2020, 277, 119241. [Google Scholar] [CrossRef]
  58. Zhang, M.-D.; Si, D.-H.; Yi, J.-D.; Yin, Q.; Huang, Y.-B.; Cao, R. Conductive phthalocyanine-based metal-organic framework as a highly efficient electrocatalyst for carbon dioxide reduction reaction. Sci. China Chem. 2021, 64, 1332–1339. [Google Scholar] [CrossRef]
  59. Wu, J.-X.; Yuan, W.-W.; Xu, M.; Gu, Z.-Y. Ultrathin 2D nickel zeolitic imidazolate framework nanosheets for electrocatalytic reduction of CO2. Chem. Commun. 2019, 55, 11634–11637. [Google Scholar] [CrossRef]
  60. Zhao, X.; Pachfule, P.; Li, S.; Langenhahn, T.; Ye, M.; Schlesiger, C.; Praetz, S.; Schmidt, J.; Thomas, A. Macro/Microporous Covalent Organic Frameworks for Efficient Electrocatalysis. J. Am. Chem. Soc. 2019, 141, 6623–6630. [Google Scholar] [CrossRef] [PubMed]
  61. Yue, Y.; Qiao, Z.-A.; Fulvio, P.F.; Binder, A.J.; Tian, C.; Chen, J.; Nelson, K.M.; Zhu, X.; Dai, S. Template-Free Synthesis of Hierarchical Porous Metal–Organic Frameworks. J. Am. Chem. Soc. 2013, 135, 9572–9575. [Google Scholar] [CrossRef]
  62. Wang, J.; Tang, J.; Ding, B.; Malgras, V.; Chang, Z.; Hao, X.; Wang, Y.; Dou, H.; Zhang, X.; Yamauchi, Y. Hierarchical porous carbons with layer-by-layer motif architectures from confined soft-template self-assembly in layered materials. Nat. Commun. 2017, 8, 15717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Anderson, M.W.; Holmes, S.M.; Hanif, N.; Cundy, C.S. Hierarchical Pore Structures through Diatom Zeolitization. Angew. Chem. 2000, 39, 2707–2710. [Google Scholar] [CrossRef]
  64. Li, K.; Lin, S.; Li, Y.; Zhuang, Q.; Gu, J. Aqueous-Phase Synthesis of Mesoporous Zr-Based MOFs Templated by Amphoteric Surfactants. Angew. Chem. 2018, 57, 3439–3443. [Google Scholar] [CrossRef]
  65. Cai, G.; Jiang, H.-L. A Modulator-Induced Defect-Formation Strategy to Hierarchically Porous Metal-Organic Frameworks with High Stability. Angew. Chem. 2016, 56, 563–567. [Google Scholar] [CrossRef]
  66. Huang, H.; Li, J.-R.; Wang, K.; Han, T.; Tong, M.; Li, L.; Xie, Y.; Yang, Q.; Liu, D.; Zhong, C. An in situ self-assembly template strategy for the preparation of hierarchical-pore metal-organic frameworks. Nat. Commun. 2015, 6, 8847. [Google Scholar] [CrossRef] [Green Version]
  67. Shen, K.; Zhang, L.; Chen, X.; Liu, L.; Zhang, D.; Han, Y.; Chen, J.; Long, J.; Luque, R.; Li, Y.; et al. Ordered macro-microporous metal-organic framework single crystals. Science 2018, 359, 206–210. [Google Scholar] [CrossRef] [Green Version]
  68. Vattipalli, V.; Qi, X.; Dauenhauer, P.J.; Fan, W. Long Walks in Hierarchical Porous Materials due to Combined Surface and Configurational Diffusion. Chem. Mater. 2016, 28, 7852–7863. [Google Scholar] [CrossRef]
  69. Kang, X.; Li, L.; Sheveleva, A.; Han, X.; Li, J.; Liu, L.; Tuna, F.; McInnes, E.J.L.; Han, B.; Yang, S.; et al. Electro-reduction of carbon dioxide at low over-potential at a metal–organic framework decorated cathode. Nat. Commun. 2020, 11, 5464. [Google Scholar] [CrossRef] [PubMed]
  70. Xin, Z.; Liu, J.; Wang, X.; Shen, K.; Yuan, Z.; Chen, Y.; Lan, Y.-Q. Implanting Polypyrrole in Metal-Porphyrin MOFs: Enhanced Electrocatalytic Performance for CO2RR. ACS Appl. Mater. Interfaces 2021, 13, 54959–54966. [Google Scholar] [CrossRef] [PubMed]
  71. Zhou, Y.; Zheng, L.; Yang, D.; Yang, H.; Lu, Q.; Zhang, Q.; Gu, L.; Wang, X. Enhancing CO2 Electrocatalysis on 2D Porphyrin-Based Metal–Organic Framework Nanosheets Coupled with Visible-Light. Small Methods 2021, 5, 2000991. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, M.-J.; Cao, S.-M.; Feng, B.-Q.; Dong, B.-X.; Ding, Y.-X.; Zheng, Q.-H.; Teng, Y.-L.; Li, Z.-W.; Liu, W.-L.; Feng, L.-G. Revealing the structure–activity relationship of two Cu-porphyrin-based metal–organic frameworks for the electrochemical CO2-to-HCOOH transformation. Dalton Trans. 2020, 49, 14995–15001. [Google Scholar] [CrossRef]
  73. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal–Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhong, H.; Ghorbani-Asl, M.; Ly, K.H.; Zhang, J.; Ge, J.; Wang, M.; Liao, Z.; Makarov, D.; Zschech, E.; Brunner, E.; et al. Synergistic electroreduction of carbon dioxide to carbon monoxide on bimetallic layered conjugated metal-organic frameworks. Nat. Commun. 2020, 11, 1409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hou, Y.; Hu, X.-J.; Tong, H.-Y.; Huang, Y.-B.; Cao, R. Unraveling the relationship of the pore structures between the metal-organic frameworks and their derived carbon materials. Inorg. Chem. Commun. 2020, 114, 107825. [Google Scholar] [CrossRef]
  76. Dou, S.; Song, J.; Xi, S.; Du, Y.; Wang, J.; Huang, Z.-F.; Xu, Z.J.; Wang, X. Boosting Electrochemical CO2 Reduction on Metal–Organic Frameworks via Ligand Doping. Angew. Chem. 2019, 58, 4041–4045. [Google Scholar] [CrossRef]
  77. Wu, H.; Zeng, M.; Zhu, X.; Tian, C.; Mei, B.; Song, Y.; Du, X.-L.; Jiang, Z.; He, L.; Xia, C.; et al. Defect Engineering in Polymeric Cobalt Phthalocyanine Networks for Enhanced Electrochemical CO2 Reduction. ChemElectroChem 2018, 5, 2717–2721. [Google Scholar] [CrossRef]
  78. Ahmad, A.; Iqbal, N.; Noor, T.; Hassan, A.; Khan, U.A.; Wahab, A.; Raza, M.A.; Ashraf, S. Cu-doped zeolite imidazole framework (ZIF-8) for effective electrocatalytic CO2 reduction. J. CO2 Util. 2021, 48, 101523. [Google Scholar] [CrossRef]
  79. Guo, F.; Liu, B.; Liu, M.; Xia, Y.; Wang, T.; Hu, W.; Fyffe, P.; Tian, L.; Chen, X. Selective electrocatalytic CO2 reduction to acetate on polymeric Cu–L (L = pyridinic N and carbonyl group) complex core–shell microspheres. Green Chem. 2021, 23, 5129–5137. [Google Scholar] [CrossRef]
  80. Iqbal, R.; Yasin, G.; Hamza, M.; Ibraheem, S.; Ullah, B.; Saleem, A.; Ali, S.; Hussain, S.; Nguyen, T.A.; Slimani, Y.; et al. State of the art two-dimensional covalent organic frameworks: Prospects from rational design and reactions to applications for advanced energy storage technologies. Coord. Chem. Rev. 2021, 447, 214152. [Google Scholar] [CrossRef]
  81. Li, H.; Feng, X.; Shao, P.; Chen, J.; Li, C.; Jayakumar, S.; Yang, Q. Synthesis of covalent organic frameworks via in situ salen skeleton formation for catalytic applications. J. Mater. Chem. A 2019, 7, 5482–5492. [Google Scholar] [CrossRef]
  82. Shah, R.; Ali, S.; Raziq, F.; Ali, S.; Ismail, P.M.; Shah, S.; Iqbal, R.; Wu, X.; He, W.; Zu, X.; et al. Exploration of metal organic frameworks and covalent organic frameworks for energy-related applications. Coord. Chem. Rev. 2023, 477, 214968. [Google Scholar] [CrossRef]
  83. Lee, M.K.; Shokouhimehr, M.; Kim, S.Y.; Jang, H.W. Two-Dimensional Metal–Organic Frameworks and Covalent–Organic Frameworks for Electrocatalysis: Distinct Merits by the Reduced Dimension. Adv. Energy Mater. 2022, 12, 2003990. [Google Scholar] [CrossRef]
  84. Wang, Y.-R.; Ding, H.-M.; Ma, X.-Y.; Liu, M.; Yang, Y.-L.; Chen, Y.; Li, S.-L.; Lan, Y.-Q. Imparting CO2 Electroreduction Auxiliary for Integrated Morphology Tuning and Performance Boosting in a Porphyrin-based Covalent Organic Framework. Angew. Chem. 2022, 61, e202114648. [Google Scholar] [CrossRef]
  85. Zhu, H.-J.; Lu, M.; Wang, Y.-R.; Yao, S.-J.; Zhang, M.; Kan, Y.-H.; Liu, J.; Chen, Y.; Li, S.-L.; Lan, Y.-Q. Efficient electron transmission in covalent organic framework nanosheets for highly active electrocatalytic carbon dioxide reduction. Nat. Commun. 2020, 11, 497. [Google Scholar] [CrossRef] [Green Version]
  86. Yang, Y.-L.; Wang, Y.-R.; Gao, G.-K.; Liu, M.; Miao, C.; Li, L.-Y.; Cheng, W.; Zhao, Z.-Y.; Chen, Y.; Xin, Z.; et al. Self-assembly of single metal sites embedded covalent organic frameworks into multi-dimensional nanostructures for efficient CO2 electroreduction. Chin. Chem. Lett. 2022, 33, 1439–1444. [Google Scholar] [CrossRef]
  87. Zhou, T.; Zhao, Y.; Choi, J.W.; Coskun, A. Lithium-Salt Mediated Synthesis of a Covalent Triazine Framework for Highly Stable Lithium Metal Batteries. Angew. Chem. 2019, 58, 16795–16799. [Google Scholar] [CrossRef] [Green Version]
  88. Liu, M.; Guo, L.; Jin, S.; Tan, B. Covalent triazine frameworks: Synthesis and applications. J. Mater. Chem. A 2019, 7, 5153–5172. [Google Scholar] [CrossRef]
  89. Ghosh, I.; Khamrai, J.; Savateev, A.; Shlapakov, N.; Antonietti, M.; König, B. Organic semiconductor photocatalyst can bifunctionalize arenes and heteroarenes. Science 2019, 365, 360–366. [Google Scholar] [CrossRef] [Green Version]
  90. Bhadra, M.; Kandambeth, S.; Sahoo, M.K.; Addicoat, M.; Balaraman, E.; Banerjee, R. Triazine Functionalized Porous Covalent Organic Framework for Photo-organocatalytic EZ Isomerization of Olefins. J. Am. Chem. Soc. 2019, 141, 6152–6156. [Google Scholar] [CrossRef]
  91. Suo, X.; Zhang, F.; Yang, Z.; Chen, H.; Wang, T.; Wang, Z.; Kobayashi, T.; Do-Thanh, C.-L.; Maltsev, D.; Liu, Z.; et al. Highly Perfluorinated Covalent Triazine Frameworks Derived from a Low-Temperature Ionothermal Approach Towards Enhanced CO2 Electroreduction. Angew. Chem. 2021, 60, 25688–25694. [Google Scholar] [CrossRef]
  92. Lin, S.; Diercks, C.S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.M.; et al. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208–1213. [Google Scholar] [CrossRef] [Green Version]
  93. Yue, Y.; Cai, P.; Xu, K.; Li, H.; Chen, H.; Zhou, H.-C.; Huang, N. Stable Bimetallic Polyphthalocyanine Covalent Organic Frameworks as Superior Electrocatalysts. J. Am. Chem. Soc. 2021, 143, 18052–18060. [Google Scholar] [CrossRef] [PubMed]
  94. Miao, Q.; Lu, C.; Xu, Q.; Yang, S.; Liu, M.; Liu, S.; Yu, C.; Zhuang, X.; Jiang, Z.; Zeng, G. CoN2O2 sites in carbon nanosheets by template-pyrolysis of COFs for CO2RR. Chem. Eng. J. 2022, 450, 138427. [Google Scholar] [CrossRef]
  95. Dong, H.; Lu, M.; Wang, Y.; Tang, H.-L.; Wu, D.; Sun, X.; Zhang, F.-M. Covalently anchoring covalent organic framework on carbon nanotubes for highly efficient electrocatalytic CO2 reduction. Appl. Catal. B Environ. 2021, 303, 120897. [Google Scholar] [CrossRef]
  96. Yuan, J.; Chen, S.; Zhang, Y.; Li, R.; Zhang, J.; Peng, T. Structural Regulation of Coupled Phthalocyanine–Porphyrin Covalent Organic Frameworks to Highly Active and Selective Electrocatalytic CO2 Reduction. Adv. Mater. 2022, 34, 2203139. [Google Scholar] [CrossRef]
  97. Liu, M.; Wang, Y.-R.; Ding, H.-M.; Lu, M.; Gao, G.-K.; Dong, L.-Z.; Li, Q.; Chen, Y.; Li, S.-L.; Lan, Y.-Q. Self-assembly of anthraquinone covalent organic frameworks as 1D superstructures for highly efficient CO2 electroreduction to CH4. Sci. Bull. 2021, 66, 1659–1668. [Google Scholar] [CrossRef]
  98. Wang, Y.; Chen, J.; Wang, G.; Li, Y.; Wen, Z. Perfluorinated Covalent Triazine Framework Derived Hybrids for the Highly Selective Electroconversion of Carbon Dioxide into Methane. Angew. Chem. 2018, 57, 13120–13124. [Google Scholar] [CrossRef]
  99. Gong, L.; Gao, Y.; Wang, Y.; Chen, B.; Yu, B.; Liu, W.; Han, B.; Lin, C.; Bian, Y.; Qi, D.; et al. Efficient electrocatalytic carbon dioxide reduction with tetraphenylethylene- and porphyrin-based covalent organic frameworks. Catal. Sci. Technol. 2022, 12, 6566–6571. [Google Scholar] [CrossRef]
  100. Chi, S.-Y.; Chen, Q.; Zhao, S.-S.; Si, D.-H.; Wu, Q.-J.; Huang, Y.-B.; Cao, R. Three-dimensional porphyrinic covalent organic frameworks for highly efficient electroreduction of carbon dioxide. J. Mater. Chem. A 2022, 10, 4653–4659. [Google Scholar] [CrossRef]
  101. Zhai, L.; Yang, S.; Lu, C.; Cui, C.-X.; Xu, Q.; Liu, J.; Yang, X.; Meng, X.; Lu, S.; Zhuang, X.; et al. CoN5 Sites Constructed by Anchoring Co Porphyrins on Vinylene-Linked Covalent Organic Frameworks for Electroreduction of Carbon Dioxide. Small 2022, 18, 2200736. [Google Scholar] [CrossRef]
  102. Lu, M.; Zhang, M.; Liu, C.-G.; Liu, J.; Shang, L.-J.; Wang, M.; Chang, J.-N.; Li, S.-L.; Lan, Y.-Q. Stable Dioxin-Linked Metallophthalocyanine Covalent Organic Frameworks (COFs) as Photo-Coupled Electrocatalysts for CO2 Reduction. Angew. Chem. 2021, 60, 4864–4871. [Google Scholar] [CrossRef]
  103. Han, B.; Jin, Y.; Chen, B.; Zhou, W.; Yu, B.; Wei, C.; Wang, H.; Wang, K.; Chen, Y.; Chen, B.; et al. Maximizing Electroactive Sites in a Three-Dimensional Covalent Organic Framework for Significantly Improved Carbon Dioxide Reduction Electrocatalysis. Angew. Chem. 2022, 61, e202114244. [Google Scholar] [CrossRef]
  104. Lu, Y.; Zhang, J.; Wei, W.; Ma, D.-D.; Wu, X.-T.; Zhu, Q.-L. Efficient Carbon Dioxide Electroreduction over Ultrathin Covalent Organic Framework Nanolayers with Isolated Cobalt Porphyrin Units. ACS Appl. Mater. Interfaces 2020, 12, 37986–37992. [Google Scholar] [CrossRef]
  105. Ma, L.; Hu, W.; Mei, B.; Liu, H.; Yuan, B.; Zang, J.; Chen, T.; Zou, L.; Zou, Z.; Yang, B.; et al. Covalent Triazine Framework Confined Copper Catalysts for Selective Electrochemical CO2 Reduction: Operando Diagnosis of Active Sites. ACS Catal. 2020, 10, 4534–4542. [Google Scholar] [CrossRef]
  106. Stalder, C.J.; Chao, S.; Wrighton, M.S. Electrochemical reduction of aqueous bicarbonate to formate with high current efficiency near the thermodynamic potential at chemically derivatized electrodes. J. Am. Chem. Soc. 1984, 106, 3673–3675. [Google Scholar] [CrossRef]
  107. O’Toole, T.R.; Margerum, L.D.; Westmoreland, T.D.; Vining, W.J.; Murray, R.W.; Meyer, T.J. Electrocatalytic reduction of CO2 at a chemically modified electrode. J. Chem. Soc. Chem. Commun. 1985, 1416–1417. [Google Scholar] [CrossRef]
  108. Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J.M.; Balbuena, P.B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674–9754. [Google Scholar] [CrossRef]
  109. Pardakhti, M.; Jafari, T.; Tobin, Z.; Dutta, B.; Moharreri, E.; Shemshaki, N.S.; Suib, S.; Srivastava, R. Trends in Solid Adsorbent Materials Development for CO2 Capture. ACS Appl. Mater. Interfaces 2019, 11, 34533–34559. [Google Scholar] [CrossRef]
  110. Peng, A.-Z.; Qi, S.-C.; Liu, X.; Xue, D.-M.; Peng, S.-S.; Yu, G.-X.; Liu, X.-Q.; Sun, L.-B. N-doped porous carbons derived from a polymer precursor with a record-high N content: Efficient adsorbents for CO2 capture. Chem. Eng. J. 2019, 372, 656–664. [Google Scholar] [CrossRef]
  111. Xiong, L.; Wang, X.-F.; Li, L.; Jin, L.; Zhang, Y.-G.; Song, S.-L.; Liu, R.-P. Nitrogen-Enriched Porous Carbon Fiber as a CO2 Adsorbent with Superior CO2 Selectivity by Air Activation. Energy Fuels 2019, 33, 12558–12567. [Google Scholar] [CrossRef]
  112. Yoo, D.K.; An, H.J.; Khan, N.A.; Hwang, G.T.; Jhung, S.H. Record-high adsorption capacities of polyaniline-derived porous carbons for the removal of personal care products from water. Chem. Eng. J. 2018, 352, 71–78. [Google Scholar] [CrossRef]
  113. Qezelsefloo, E.; Khalili, S.; Jahanshahi, M.; Peyravi, M. Adsorptive removal of CO2 on Nitrogen-doped porous carbon derived from polyaniline: Effect of chemical activation. Mater. Chem. Phys. 2020, 239, 122304. [Google Scholar] [CrossRef]
  114. He, J.; To, J.W.F.; Psarras, P.C.; Yan, H.; Atkinson, T.; Holmes, R.T.; Nordlund, D.; Bao, Z.; Wilcox, J. Tunable Polyaniline-Based Porous Carbon with Ultrahigh Surface Area for CO2 Capture at Elevated Pressure. Adv. Energy Mater. 2016, 6, 1502491. [Google Scholar] [CrossRef]
  115. Park, J.M.; Woo, H.C.; Jhung, S.H. Effective CO2 adsorption at low pressure over nitrogen-enriched porous carbons, derived from melamine-loaded polyaniline. Chem. Eng. J. 2021, 412, 128641. [Google Scholar] [CrossRef]
  116. Groppo, E.; Uddin, M.J.; Zavorotynska, O.; Damin, A.; Vitillo, J.G.; Spoto, G.; Zecchina, A. Exploring the Chemistry of Electron-Accepting Molecules in the Cavities of the Basic Microporous P4VP Polymer by in situ FTIR Spectroscopy. J. Phys. Chem. C 2008, 112, 19493–19500. [Google Scholar] [CrossRef]
  117. Shieh, J.-J.; Chung, T.S. Gas permeability, diffusivity, and solubility of poly(4-vinylpyridine) film. J. Polym. Sci. Part B Polym. Phys. 1999, 37, 2851–2861. [Google Scholar] [CrossRef]
  118. Ye, C.; Raaijman, S.J.; Chen, X.; Koper, M.T.M. Enhanced Electrochemical CO2 Reduction to Formate on Poly(4-vinylpyridine)-Modified Copper and Gold Electrodes. ACS Appl. Mater. Interfaces 2022, 14, 45263–45271. [Google Scholar] [CrossRef]
  119. Qian, X.; Meng, X.; Sun, J.; Jiang, L.; Wang, Y.; Zhang, J.; Hu, X.; Shalom, M.; Zhu, J. Salt-Assisted Synthesis of 3D Porous g-C3N4 as a Bifunctional Photo- and Electrocatalyst. ACS Appl. Mater. Interfaces 2019, 11, 27226–27232. [Google Scholar] [CrossRef]
  120. Ren, Y.; Han, Q.; Zhao, Y.; Wen, H.; Jiang, Z. The exploration of metal-free catalyst g-C3N4 for NO degradation. J. Hazard. Mater. 2021, 404, 124153. [Google Scholar] [CrossRef]
  121. Zhou, Y.; Zhang, Y.; Xue, J.; Wang, R.; Yin, Z.; Ding, L.; Wang, H. Graphene oxide-modified g-C3N4 nanosheet membranes for efficient hydrogen purification. Chem. Eng. J. 2021, 420, 129574. [Google Scholar] [CrossRef]
  122. Wu, J.; Liu, M.; Sharma, P.P.; Yadav, R.M.; Ma, L.; Yang, Y.; Zou, X.; Zhou, X.-D.; Vajtai, R.; Yakobson, B.I.; et al. Incorporation of Nitrogen Defects for Efficient Reduction of CO2 via Two-Electron Pathway on Three-Dimensional Graphene Foam. Nano Lett. 2016, 16, 466–470. [Google Scholar] [CrossRef] [PubMed]
  123. Zhao, S.; Tang, Z.; Guo, S.; Han, M.; Zhu, C.; Zhou, Y.; Bai, L.; Gao, J.; Huang, H.; Li, Y.; et al. Enhanced Activity for CO2 Electroreduction on a Highly Active and Stable Ternary Au-CDots-C3N4 Electrocatalyst. ACS Catal. 2018, 8, 188–197. [Google Scholar] [CrossRef]
  124. Hu, C.; Liu, M.-T.; Sakai, A.; Yoshida, M.; Lin, K.-Y.A.; Huang, C.-C. Synergistic effect of Cu and Ru decoration on g-C3N4 for electrocatalytic CO2 reduction. J. Ind. Eng. Chem. 2022, 115, 329–338. [Google Scholar] [CrossRef]
Scheme 1. Porous materials for CO2 capture and electrocatalytic reduction.
Scheme 1. Porous materials for CO2 capture and electrocatalytic reduction.
Materials 16 01630 sch001
Figure 1. (a) N2−sorption isotherms at 77 K. (b) Pore size distribution. (c) Isosteric-heat values of the CO2 adsorption of LKK−CMP−1 [46]. Reprinted with permission from Ref. [46]. Copyright 2018 Copyright, American Chemical Society. (Washington, DC, USA). (d) The strategy for constructing the microporous polymer networks. (e) Pore size distribution of the polyimides. (f) CO2 adsorption isotherms for the crosslinked polyimides. (g) Isosteric-heat values of the CO2 adsorption of 6FA-PE-CL and 6FA-PEPH-CL [47]. Reprinted with permission from Ref. [47]. Copyright 2020 Copyright, Elsevier Ltd. (Amsterdam, The Netherlands).
Figure 1. (a) N2−sorption isotherms at 77 K. (b) Pore size distribution. (c) Isosteric-heat values of the CO2 adsorption of LKK−CMP−1 [46]. Reprinted with permission from Ref. [46]. Copyright 2018 Copyright, American Chemical Society. (Washington, DC, USA). (d) The strategy for constructing the microporous polymer networks. (e) Pore size distribution of the polyimides. (f) CO2 adsorption isotherms for the crosslinked polyimides. (g) Isosteric-heat values of the CO2 adsorption of 6FA-PE-CL and 6FA-PEPH-CL [47]. Reprinted with permission from Ref. [47]. Copyright 2020 Copyright, Elsevier Ltd. (Amsterdam, The Netherlands).
Materials 16 01630 g001
Figure 2. (a) The structure of triazine linking framework TTF-1. (b) N2−sorption isotherms on TTF-1 were recorded at 77 K and the corresponding pore size distribution [48]. Reprinted with permission from Ref. [48]. Copyright 2018 Copyright, American Chemical Society. (Washington, DC, USA).
Figure 2. (a) The structure of triazine linking framework TTF-1. (b) N2−sorption isotherms on TTF-1 were recorded at 77 K and the corresponding pore size distribution [48]. Reprinted with permission from Ref. [48]. Copyright 2018 Copyright, American Chemical Society. (Washington, DC, USA).
Materials 16 01630 g002
Figure 3. (a) The structure of POP−Py (n). (b) The structure of CoTCPP. (c) Faradaic efficiencies for carbon monoxide and hydrogen production on CoTCPP and POP−Py(n)/CoTCPP. (d) Representative chronoamperograms of CO2 electroreduction catalyzed by POP−Py(2)/CoTCPP at various potentials in a 0.5 M KHCO3 aqueous solution [49]. Reprinted with permission from Ref. [49]. Copyright 2020, Royal Society of Chemical (London, UK).
Figure 3. (a) The structure of POP−Py (n). (b) The structure of CoTCPP. (c) Faradaic efficiencies for carbon monoxide and hydrogen production on CoTCPP and POP−Py(n)/CoTCPP. (d) Representative chronoamperograms of CO2 electroreduction catalyzed by POP−Py(2)/CoTCPP at various potentials in a 0.5 M KHCO3 aqueous solution [49]. Reprinted with permission from Ref. [49]. Copyright 2020, Royal Society of Chemical (London, UK).
Materials 16 01630 g003
Figure 4. (a) The synthesis and structure of the PyPOP@G. (b) N2−sorption isotherms for the PyPOP, graphene, and PyPOP@G [50]. Reprinted with permission from Ref. [50]. Copyright 2016, THE SOCIETY, (London, UK ). (c) N2−sorption isotherms for the MWCNTs, POP, POP@MWCNT. (d) The Qst of CO2 into the POP and the POP@MWCNTs. (e) The electrode behavior of POP@MWCNT and POP-Cu@MWCNT [51]. Reprinted with permission from Ref. [51]. Copyright 2016, Elsevier B.V. (Amsterdam, The Netherlands).
Figure 4. (a) The synthesis and structure of the PyPOP@G. (b) N2−sorption isotherms for the PyPOP, graphene, and PyPOP@G [50]. Reprinted with permission from Ref. [50]. Copyright 2016, THE SOCIETY, (London, UK ). (c) N2−sorption isotherms for the MWCNTs, POP, POP@MWCNT. (d) The Qst of CO2 into the POP and the POP@MWCNTs. (e) The electrode behavior of POP@MWCNT and POP-Cu@MWCNT [51]. Reprinted with permission from Ref. [51]. Copyright 2016, Elsevier B.V. (Amsterdam, The Netherlands).
Materials 16 01630 g004
Figure 5. (a) N2 adsorption/desorption curves of Bi−MOF; (b) CO2 adsorption curves of Bi-MOF and Bi sheets; (c) Faraday efficiency of HCOOH; (d) structure of Bi-MOF [57]. Reprinted with permission from Ref. [57]. Copyright 2020 Copyright Elsevier (Amsterdam, The Netherlands).
Figure 5. (a) N2 adsorption/desorption curves of Bi−MOF; (b) CO2 adsorption curves of Bi-MOF and Bi sheets; (c) Faraday efficiency of HCOOH; (d) structure of Bi-MOF [57]. Reprinted with permission from Ref. [57]. Copyright 2020 Copyright Elsevier (Amsterdam, The Netherlands).
Materials 16 01630 g005
Figure 6. (a) Structure of NiPc−Ni(NH)4; (b) N2 adsorption/desorption isotherms of NiPc−Ni(NH)4; (c) CO2 adsorption isotherms of NiPc−Ni(NH)4 [58]. Reprinted with permission from Ref. [58]. Copyright 2021 Copyright Springer (Berlin, Germany).
Figure 6. (a) Structure of NiPc−Ni(NH)4; (b) N2 adsorption/desorption isotherms of NiPc−Ni(NH)4; (c) CO2 adsorption isotherms of NiPc−Ni(NH)4 [58]. Reprinted with permission from Ref. [58]. Copyright 2021 Copyright Springer (Berlin, Germany).
Materials 16 01630 g006
Figure 7. (a) Structure of 2D Ni(Im)2; (b) N2 adsorption/desorption isotherms; (c) Potential−dependent Faradaic efficiencies of CO [59]. Reprinted with permission from Ref. [59]. Copyright 2019 Copyright Royal Society of Chemistry (London, UK).
Figure 7. (a) Structure of 2D Ni(Im)2; (b) N2 adsorption/desorption isotherms; (c) Potential−dependent Faradaic efficiencies of CO [59]. Reprinted with permission from Ref. [59]. Copyright 2019 Copyright Royal Society of Chemistry (London, UK).
Materials 16 01630 g007
Figure 8. (a,b) Views of the crystal structure of [Cu2(L)] (CCDC1531622); (c,d) micropore and mesopore size distribution. (Cu2(L)−t (black lines) and Cu2(L)−e (red lines)) [69]. Reprinted with permission from Ref. [69]. Copyright 2020 Copyright Nature (Oxford, England).
Figure 8. (a,b) Views of the crystal structure of [Cu2(L)] (CCDC1531622); (c,d) micropore and mesopore size distribution. (Cu2(L)−t (black lines) and Cu2(L)−e (red lines)) [69]. Reprinted with permission from Ref. [69]. Copyright 2020 Copyright Nature (Oxford, England).
Materials 16 01630 g008
Figure 9. (a) Schematic diagram of the synthesis of F-CTF made by ionothermal method under the catalysis of Zn(NTf2)2. (b) Schematic diagram of the reaction for Zn(NTf2)2−catalyzed production of TFPT using PFBN. (c) N2 adsorption/desorption curves of F−CTF−1−T; (d,e) pore distribution [91]. Reprinted with permission from Ref. [91]. Copyright 2021 Copyright Wiley-VCH (Wehenheim, Germany).
Figure 9. (a) Schematic diagram of the synthesis of F-CTF made by ionothermal method under the catalysis of Zn(NTf2)2. (b) Schematic diagram of the reaction for Zn(NTf2)2−catalyzed production of TFPT using PFBN. (c) N2 adsorption/desorption curves of F−CTF−1−T; (d,e) pore distribution [91]. Reprinted with permission from Ref. [91]. Copyright 2021 Copyright Wiley-VCH (Wehenheim, Germany).
Materials 16 01630 g009
Figure 10. (a) Structure of CuPcF8−CoNPc−COF and CuPcF8−CoPc−COF; N2 adsorption/desorption isotherms of (b)CuPcF8−CoPc-COF and (c) CuPcF8−CoNPc−COF [93]. Reprinted with permission from Ref. [93]. Copyright 2021 Copyright American Chemical Society (Washington, DC, USA).
Figure 10. (a) Structure of CuPcF8−CoNPc−COF and CuPcF8−CoPc−COF; N2 adsorption/desorption isotherms of (b)CuPcF8−CoPc-COF and (c) CuPcF8−CoNPc−COF [93]. Reprinted with permission from Ref. [93]. Copyright 2021 Copyright American Chemical Society (Washington, DC, USA).
Materials 16 01630 g010
Figure 11. (a) BET plot of 2D−Co−COF500; (b) pore distribution; (c) the FEs of CO; (d) the current density of CO [94]. Reprinted with permission from Ref. [94]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Figure 11. (a) BET plot of 2D−Co−COF500; (b) pore distribution; (c) the FEs of CO; (d) the current density of CO [94]. Reprinted with permission from Ref. [94]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Materials 16 01630 g011
Figure 12. (a) BET plot of MWCNT−Por−COF-Cu at 77K; (b) Pore distribution of MWCNT−Por−COF−Co and Por−COF; (c) Faraday efficiency potentials in the range of −0.5 to −1.0 V for CO; (d) Products obtained by electrocatalysis of MWCNT−Por−COF−Cu [95]. Reprinted with permission from Ref. [95]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Figure 12. (a) BET plot of MWCNT−Por−COF-Cu at 77K; (b) Pore distribution of MWCNT−Por−COF−Co and Por−COF; (c) Faraday efficiency potentials in the range of −0.5 to −1.0 V for CO; (d) Products obtained by electrocatalysis of MWCNT−Por−COF−Cu [95]. Reprinted with permission from Ref. [95]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Materials 16 01630 g012
Figure 13. (a) N2 adsorption/desorption isotherms of MPCs and PC-550; (b) pore distribution of PC-550 and MPC-550; (c) CO2 adsorption isotherms of MPCs and PC-550 [116]. Reprinted with permission from Ref. [116]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Figure 13. (a) N2 adsorption/desorption isotherms of MPCs and PC-550; (b) pore distribution of PC-550 and MPC-550; (c) CO2 adsorption isotherms of MPCs and PC-550 [116]. Reprinted with permission from Ref. [116]. Copyright 2022 Copyright Elsevier (Amsterdam, The Netherlands).
Materials 16 01630 g013
Figure 14. (a) BET plot of 4 wt% Au−CDots−C3N4; (b) BET surface area (black), pore diameter (green), and pore volume (red) of the CuxRuyCN [124]. Reprinted with permission from Ref. [124]. Copyright 2021 Copyright Elsevier (Amsterdam, The Netherlands).
Figure 14. (a) BET plot of 4 wt% Au−CDots−C3N4; (b) BET surface area (black), pore diameter (green), and pore volume (red) of the CuxRuyCN [124]. Reprinted with permission from Ref. [124]. Copyright 2021 Copyright Elsevier (Amsterdam, The Netherlands).
Materials 16 01630 g014
Table 1. Reaction mechanism of electrocatalytic CO2 reduction.
Table 1. Reaction mechanism of electrocatalytic CO2 reduction.
Cathodic ReactionProductsE°/V
CO2 + 2H+ + 2e→HCOOHHCOOH−0.61
CO2 + 2H+ + 2e→CO + H2OCO−0.53
CO2 + 6H+ + 6e→CH3OH + H2OCH3OH−0.38
CO2 + 8H+ + 8e→CH4 + 2H2OCH4−0.24
2CO2 + 8H+ + 8e→CH3COOH + 2H2OCH3COOH−0.30
2CO2 + 12H+ + 12e→C2H4 + 4H2OC2H4−0.34
2CO2 + 12H+ + 12e→CH3CH2OH + 3H2OCH3CH2OH−0.33
2CO2 + 14H+ + 14e→C2H6 + 4H2OC2H6−0.27
2H+ + 2e→H2H2−0.42
Table 5. Advantages and disadvantages of porous polymer materials for CO2 capture and electrocatalytic reduction.
Table 5. Advantages and disadvantages of porous polymer materials for CO2 capture and electrocatalytic reduction.
MaterialsStructural FeaturesAdvantagesDisadvantages
POPsA 2D or 3D network structure expanded by organic groups or fragments generated by chemical polymerization
  • Large surface area
  • Controllable functionality
  • Tunable porosity
  • Good thermal stability
  • High cost
  • Poor porosity
PCPsA solid crystal material formed by connecting metal ions or metal clusters with organic complexes.
  • Large specific surface area
  • Adjustable pore surface function
  • Poor thermal stability
  • High cost
COFsA kind of organic building blocks connected together by covalent bonds to form a porous skeleton with periodic structure.
  • Large surface areas
  • Abundant active sites
  • Diverse synthetic strategies
  • High cost
  • Poor porosity
  • Difficulty in controlling crystallization
N-polymersA kind of material with nitrogen functional groups in the main chain or side chain of polymer, which can complex or anchor metal particles.
  • Simple preparation methods
  • Cost efficiency
  • Adjustable pore structures
  • Poor conductivity
  • Poor mechanical stability
  • Difficulty in large-scale preparation
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Wang, G.; Hu, L.; Ge, B.; Yu, X.; Deng, J. Porous Polymer Materials for CO2 Capture and Electrocatalytic Reduction. Materials 2023, 16, 1630. https://doi.org/10.3390/ma16041630

AMA Style

Wang H, Wang G, Hu L, Ge B, Yu X, Deng J. Porous Polymer Materials for CO2 Capture and Electrocatalytic Reduction. Materials. 2023; 16(4):1630. https://doi.org/10.3390/ma16041630

Chicago/Turabian Style

Wang, Hui, Genyuan Wang, Liang Hu, Bingcheng Ge, Xiaoliang Yu, and Jiaojiao Deng. 2023. "Porous Polymer Materials for CO2 Capture and Electrocatalytic Reduction" Materials 16, no. 4: 1630. https://doi.org/10.3390/ma16041630

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop