Download PDF
Review  |  Open Access  |  9 Aug 2023

Synthesis strategies of metal-organic frameworks for CO2 capture

Views: 719 |  Downloads: 201 |  Cited:   2
Microstructures 2023;3:2023032.
10.20517/microstructures.2023.32 |  © The Author(s) 2023.
Author Information
Article Notes
Cite This Article

Abstract

The high consumption of fossil energy has led to increasing concentrations of carbon dioxide (CO2) in the atmosphere, making carbon capture and separation a research hotspot in this century. As novel porous materials, metal-organic frameworks (MOFs) are widely used for CO2 capture due to their unique structures and tunable properties. Currently, several relatively mature strategies have been applied to synthesize MOFs for CO2 capture. Herein, we investigate strategies for tuning the pore windows, pore sizes, open metal sites, and post-synthesis or pre-synthesis modifications of MOFs from the perspective of CO2 capture performance. Furthermore, we summarize the relevant CO2 capture technologies and research advances and describe the application of different strategies in the synthesis of CO2 capture-oriented MOFs.

Keywords

Metal-organic frameworks, carbon capture and separation, CO2 capture strategy

INTRODUCTION

Since the Industrial Revolution, the extraction and consumption of fossil fuels have caused a remarkable expansion of greenhouse gases in the atmosphere. The Earth's climate is undergoing major changes characterized by global warming, according to numerous authoritative studies[1-4]. Compared to other greenhouse gases, carbon dioxide (CO2) has a weaker greenhouse effect, but it has the highest proportion in the atmosphere, with its warming effect accounting for about 60% of the total warming effect among all greenhouse gases[5,6]. Therefore, it is urgent to work out an environment-friendly CO2 capture technology to alleviate the greenhouse effect[7,8].

In recent years, the development of carbon capture and separation (CCS) technology has attracted social attention. Adsorbents that can be used for CO2 capture include activated carbon, zeolite, alumina, metal oxides (CaO, MgO, K2O, Li2O), metal-organic frameworks (MOFs), and other surface-modified porous media[9-12]. Compared to traditional inorganic porous materials, MOFs have many advantages and show great application potential in CO2 adsorption and sequestration. One is the modifiability of its secondary building units (SBUs) and organic ligands. Through the modification of inorganic/organic nodes on functionality groups, the fine-tuning of the pore size and channel environment in MOFs can be precisely achieved, and then MOFs oriented to CO2 capture can be synthesized[13-15]. On the other hand, due to their high density of active sites, high stability, and rich topological structures, MOFs have distinct advantages, such as mild reaction conditions and easy adsorption and sequestration of CO2[16-18].

A large number of relevant papers and reviews have been published[19,20]. In contrast to the published reviews, this paper presents the current CO2 capture technology and related adsorbents. Then, the parameters for evaluating the performance of CO2 capture adsorbents are presented to achieve the best capture capacity and reduce costs and energy expenditure. In addition, recent advances in MOF-based CO2 capture methods and ways to improve the capture performance of the materials are explored. The strategies and methods described in this review will not only provide new propositions for the construction of CO2-oriented capture MOFs but also contribute to the mitigation of industrial consumption to achieve carbon neutrality and thus slow down the process of global warming.

CO2 CAPTURE

In September 2020, the “Address at the General Debate of the 75th Session of the United Nations General Assembly” and the “Address at the United Nations Biodiversity Summit” proposed that China should aim to achieve CO2 peak emissions by 2030 and carbon neutrality by 2060. In addition, many countries around the world have enacted policies or legislation to advance the goal of carbon neutrality. So far, 29 countries and regions have pledged to be carbon neutral[21]. The prevailing view is that CCS may be the cheapest technology to reduce emissions in the long term and will gain time to develop new technologies such as alternative energy sources[22,23].

CO2 capture technology

CCS technology refers to the separation of CO2 from industrial and energy-related production activities and its long-term sequestration in natural underground reservoirs in order to reduce CO2 emissions to the atmosphere[24,25]. There are three main components to CCS technology: (1) CO2 capture from fixed carbon sources (e.g., power plants); (2) CO2 compression and transport; and (3) CO2 storage.

In terms of the capture process, there are three main types of CO2 capture technologies: pre-combustion capture, post-combustion capture, and oxy-fuel combustion[26] [Figure 1]. Pre-combustion capture uses new gasification technology to convert fossil fuels into H2 and CO2, which are then captured before combustion. Compared with direct coal combustion, this technology has higher fossil fuel utilization, less waste disposal, and less water consumption. In addition, due to high CO2/H2 gas pressure and CO2 concentration, the energy consumption of pre-combustion CO2 capture is only 10%-16% of the total energy consumption, which is about half of the energy consumption of post-combustion capture. However, the high energy consumption of the fuel conversion step limits its further development[27-29]. Oxy-fuel combustion is based on burning fossil fuels in oxygen-rich gases to obtain high concentrations of CO2, which is direct sequestration and utilization. Due to the removal of N2 in the air, the volume of combustion gas is reduced, so the volume of flue gas produced after combustion is reduced to 1/5-1/3 of that of conventional coal-fired boilers. Meanwhile, the concentration of CO2 in the flue gas is high, making the cost of capturing CO2 lower. However, this technology is not ideal for CO2 capture due to its high energy consumption and investment costs[30,31]. Post-combustion capture means capturing CO2 in the flue gas from the combustion emissions. Since the partial pressure of CO2 after capture is low, it is necessary to pressurize the CO2 before storage, which increases the operating cost. Even so, post-combustion capture has the following advantages over other capture techniques: (1) A wide range of applications. Post-combustion capture technology can not only separate CO2 from flue gas but also capture NOX and SOX to achieve the purpose of denitrification and desulfurization; (2) Strong applicability. The trapping device is installed in the flue gas tail of the power plant, which has no influence on the existing power generation equipment and; (3) Relatively mature development and flexible operation[32-34]. Therefore, post-combustion capture is a common CO2 capture technology.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 1. Three basic routes (post-combustion capture, pre-combustion capture, and oxy-fuel combustion) of CO2 capture in industry.

In addition, CO2 capture technology can be classified based on different capture methods, including chemical absorption[35], membrane separation[36-38], biological immobilization[39], adsorption[40-43], and other approaches. Among them, the adsorption method is the most competitive CO2 capture technique.

CO2 capture adsorbent

Adsorption is the uptake of molecules or ions from the surrounding liquid or gas by the surface of a solid substance. The adsorption method has been widely studied by scholars for its advantages of low cost and simple operation[44]. However, its practical application requires the design and synthesis of an easily regenerated and durable adsorbent material. Generally speaking, a suitable CO2 adsorbent needs to have high selectivity and adsorption capacity, adequate adsorption/desorption kinetics, stability after several adsorption/desorption cycles, and good chemical and mechanical stability.

Among many CO2 adsorbents, porous solid adsorbents have good application prospects because of their low adsorption enthalpy and easy recycling. The traditional porous adsorbents for CO2 capture mainly include activated carbon, zeolite, silica gel, activated alumina, and so on[45-48]. Zeolite, a common solid adsorbent, is a porous silica-aluminate material with a pore size between 5 Å and 12 Å[49-51]. Among them, zeolite 13X is the most widely studied adsorbent, with a specific surface area of 726 m2 g-1, a pore volume of 0.25 cm3 g-1, and a CO2 adsorption capacity of 16.4 wt% (0.8 bar) at room temperature, which can be used as a benchmark for solid adsorbents[52,53].

Another class of solid adsorbents is carbon-based adsorbents, including activated carbon, carbon molecular sieves, and carbon nanotubes, which have been used for different forms of CO2 separation. The common activated carbon has an inorganic porous structure, high surface area, and high CO2 adsorption capacity, and it can be generated through resin and other pyrolysis processes, which is a low-cost and highly available adsorbent[54,55]. Even in an environment with water vapor, activated carbon can maintain its structure without being destroyed and has high adsorption stability. However, activated carbon has lower CO2 adsorption capacity at low pressures and poor adsorption selectivity for different gases due to the absence of an electric field generated by cations on the activated carbon surface. To sum up, these traditional adsorption technologies and adsorbents need to be further improved to enhance their ability of CO2 capture and separation.

MOFs are the most representative new porous materials and promising adsorbents, which have attracted extensive attention in gas separation applications[56,57]. As a porous adsorbent with a framework structure formed by organic ligands bridging metal ion nodes, the MOF has a very high surface area, ultra-high porosity, flexibility of the porous structure, and diversity of surface functional groups due to the presence of organic ligands and is easy to be chemically modified. These advantages make MOFs exhibit great potential for CO2 capture and storage[58-60].

CO2 capture performance parameters

Compared with other solid materials, the greatest advantage of MOFs lies in their ability to precisely fine-tune their pore size through the modulation of ligand size. At the same time, the pore environment can be modified by modifying the functional groups on the metal ions (clusters) or ligands to enhance the interaction force with the guest molecules, thus achieving the separation of different gas molecules[61-64]. There have been several reviews on MOFs for gas separation applications, and they have made great progress in adsorptive separation applications in the past few years. However, there are still some limitations. To address the limitations of MOFs, a series of studies have been conducted in recent years on improving their adsorption and separation capacities, expanding new structures, novel functionalization pathways, and adopting hybrid systems and techniques[65]. In addition, studies exploring the adsorption mechanism of MOFs and the improvement of their adsorption capacity in moist environments have gradually become the focus of attention[14,66]. An adsorbent suitable for capturing CO2 from flue gas should consider the following performance parameters.

The adsorption capacity of CO2

This is a key criterion for evaluating the performance of solid sorbents and represents the quantity of sorbent required for a given load, as well as the adsorbent bed size, which is deemed to be a key factor in determining the energy requirement in the regeneration step. In addition, the amount of CO2 adsorbed is related to its partial pressure in the gas phase. The specific adsorption of CO2 indicates the ability of the MOF material to adsorb CO2

The selectivity of CO2

It represents the adsorption ratio of the adsorbent for CO2 to other gases (usually used for post-combustion capture and natural gas upgrading). The sorption selectivity of a gas mixture for CO2 can be estimated quantitatively by means of a single-component gas sorption isotherm or ideal adsorption solution theory (IAST). The CO2 selectivity of materials is primarily due to the following aspects (the relevant parameters are shown in Table 1): (1) Selectivity based on pore size screening (molecular sieve effect). Due to the different kinetic diameters of gas molecules of each component in the gas mixture, CO2 can be effectively separated from the gas mixture by precisely adjusting the pore size of the material; (2) Selectivity based on adsorption (thermodynamic separation), i.e., separation based on the interaction forces between different gas molecules and the material surface and; (3) Selectivity based on diffusion effects, i.e., to separate the gas mixture according to the different diffusion rates of different gases in the material[5,67].

Table 1

Physical properties of common gas molecules coexisting with CO2

Gas moleculesDynamic diameter (Å)Size (Å)Boiling point (K)Polarity (10-25 cm3)
N23.64-77.3515.3
H22.89-20.38-
H2O2.65-37314.8
CO23.33.18 × 3.33 × 5.36194.7529.11
C2H23.33.32 × 3.34 × 5.70188.433.3-39.3
CH43.763.83 × 3.94 × 4.10111.625.93

The adsorption enthalpy of CO2

Adsorption enthalpy is another key parameter for evaluating the performance of the storage of CO2 by physical adsorption. It represents the strength of the interaction between the adsorbent and the adsorbate molecules. Moreover, it is an indicator of the energy required to regenerate a solid adsorbent, and the magnitude of the adsorption enthalpy clearly affects the cost of the adsorbent regeneration process. If the adsorption enthalpy is too high, the material binds CO2 too strongly, requiring a large amount of energy to break the framework-CO2 interaction, thereby increasing the regeneration cost. Conversely, very low adsorption enthalpy is undesirable. While regeneration costs will be lower, the captured CO2 will be less pure, resulting in lower adsorption selectivity and a larger adsorption bed size[68-70]. Typically, the adsorption enthalpy of MOFs is in the range of 20-50 kJ mol-1, which is comparable to other physical adsorbents (such as zeolites). Table 2 shown some MOF-based sorbents with their CO2 uptake capacity and Qst.

Table 2

MOF-based sorbent and related CO2 uptake capacity at 298 K

MaterialsBET surface area (m2/g)Pressure (bar)Capacity (mmol g-1)Qst
(kJ mol-1)
Ref.
Fe-dbai1,28016.423.5[100]
Cu(adci)-28050.152.0127.5[110]
NKU-5211,10016.2141[119]
MUF-16(Mn)21412.3137[113]
MUF-16(Ni)20412.2532[113]
ZnDatzBdc30312.05-[125]
NJU-Bai521,9080.00040.01318.1[128]
NJU-Bai531,8440.00040.6417.5[128]

The experimental adsorption enthalpy (Qst) was applied to assess the strength of the bond between the adsorbent and the adsorbate and was defined as:

$$ \begin{equation} \begin{aligned} Q_{s t}=\mathrm{RT}^{2}\left(\frac{\partial \operatorname{lnp}}{\partial \mathrm{T}}\right) \end{aligned} \end{equation} $$

The adsorption enthalpy, Qst, is determined using the pure component isotherm fits using the Clausius-Clapeyron equation, where T(K) is the temperature, p(kPa) is the pressure, and R is the gas constant.

The stability of the CO2 adsorbent

In order to reduce operating costs and operational difficulties, solid adsorbents must be able to be used in a wide range of industrial environments and show good stability under fume conditions, under adsorption operating conditions, and during multi-cycle adsorption regeneration. Particular attention must be paid to the stability of the adsorbent in the presence of water vapor. In addition, chemical and mechanical stability are equally important[71].

The adsorption/desorption kinetics of CO2

Sorption-regeneration cycle times are largely dependent on the kinetic properties of the CO2 adsorption-desorption curves measured in breakthrough experiments. The ability to effectively reduce the cycle time and the amount of adsorbent, thereby lowering the cost of CO2 separation, is the first parameter to be considered in selecting the CO2 adsorbent. Adsorbents that adsorb and desorb CO2 in a relatively short period of time become the preferred choice.

The cost of adsorbent material

This is an essential ingredient in choosing an adsorbent material. Owing to the high preparation cost, many adsorbents with excellent adsorption properties are not successfully used in industry. Therefore, the preparation of materials with good adsorption properties at low cost is considered to be the main goal of researchers in the field of CO2 capture.

Based on the above elaboration of CO2 capture performance parameters, the quantitative assessment of the CO2 capture capacity of adsorbents mainly refers to the following data: (1) The adsorption capacity of CO2 at 15 kPa is greater than 50 STP cm3 g-1; (2) IAST selectivity (CO2/N2 = 15/85) is greater than 500; (3) Good thermal stability, the structure will not collapse before the temperature is higher than 300 °C; (4) Good chemical stability, the structure remains stable in most organic solvents; (5) Qst < 40 kJ mol-1.

Research progress of MOFs for CO2 capture

As novel porous materials, MOFs were initially proposed by Prof. Omar M. Yaghi and have been extensively researched by scientists[72-77]. They are formed from inorganic metal ions or clusters and organic ligands that are connected through coordination bonds with varying degrees of connectivity. In contrast to conventional inorganic porous materials, such as porous silicates and molecular sieves, MOFs possess a remarkably adaptable structure. Different structures and characteristics of MOFs are built by choosing metal nodes with varying activities and a diverse range of organic ligands[78-80]. Furthermore, they enable specific function-oriented compositions. By utilizing various trapping mechanisms, including molecular sieve separation, host-guest interaction, and kinetic diffusion, MOFs with distinct pore sizes ranging from micropore to mesopore can be conveniently synthesized by modifying the length of ligands or functionalizing inorganic nodes and ligands, resulting in MOFs with unique properties. As novel crystalline porous materials, MOFs possess significantly higher specific surface area and porosity compared to other porous materials (with a specific surface area of up to 10,000 m2 g-1 and porosity of 90%), providing substantial scope for the development in gas storage[81,82], adsorption and separation[83], catalysis[84], sensor[85], and other areas.

The ability of MOFs to capture CO2 from varying gas mixtures depends on their inherent properties and the attributes of the gas mixtures. Over the past two decades, diverse topological structures and function-oriented MOFs have been synthesized, leading to the formation of various branches of materials, including iso-reticular MOFs (IRMOFs), zeolitic imidazolate frameworks (ZIFs), materials of institute Lavoisier frameworks (MILs), and porous coordination networks (PCNs), etc.[86-89]. Different branches possess specific characteristics to cater to distinct applications. Currently, several relatively mature strategies [Figure 2] have been applied to synthesize CO2 capture-oriented MOFs[90-92].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 2. Schematic representation of MOF performance parameters and state-of-the-art CO2 capture strategies.

Functionalized modification strategy

The -NH2 functional group is widely utilized in various adsorbent materials due to its strong attraction of CO2 to the amine group, which gives the amine molecule higher adsorption and selectivity for CO2. Moreover, numerous polar functional groups, including halogen atoms, hydroxyl, carboxyl, cyano, and nitro, have been demonstrated to influence the adsorption ability of CO2 in MOFs[93-95].

Diamine-functionalized MOFs in the form of diamine-Mg2(dobpdc) (dobpdc4- = 4,4′-dioxidobiphenyl-3,3′-dicarboxylate) offer the most potential for carbon capture applications due to their adjustable, stair-like profiles for CO2 adsorption. In view of this, Dinakar et al. reported that MOFs containing dmen-Mg2(dobpdc) (dobpdc = 1,2-diamino-2-methylpropane) composition [Figure 3] can capture CO2 from coal-fired flue gas at moderate pressure[96]. Further, using Mg2(pc-dobpdc) (pc-dobpdc = 3,3′-dioxobiphenyl-4,4′-dicarboxylate) with higher structural symmetry to avoid sub-stability during CO2 adsorption, dmen-Mg2(pc-dobpdc) demonstrates carbon capture capabilities under similar conditions in a simulated coal-fired power plant, achieving a complete CO2 adsorption effect.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 3. Depiction of the formation of ammonium carbamate chains in diamine-Mg2(dobpdc) upon cooperative CO2 adsorption. Reproduced with permission from Dinakar et al.[96]. Copyright 2021 American Chemical Society.

It has been shown that some specific sites of MOFs can effectively capture CO2, including unsaturated metal sites (UMSs) and Lewis base sites (LBSs), such as amines, pyridines, sulfones, and amides. UMSs are capable of establishing potent electrostatic interactions with CO2 and have a high CO2 capture capacity. However, the ubiquitous water molecules tend to coordinate with UMSs in a competitive manner, resulting in a significant reduction in the ability of CO2 adsorption. Regarding LBSs, compared with -NH2, the amide functional group exhibits a robust attraction towards CO2 due to the existence of two binding sites, carbonyl (CO-) and amine (NH-), leading to superior CO2 adsorption and selectivity[97-99]. In addition, MOFs modified by amide functional groups tend to be more stable.

Fe-dbai (dbai = 5-(3,5-Dicarboxybenzoylamino) isophthalic acid) combines two specific functional sites, UMSs and amide functional groups. Its CO2 adsorption capacity is measured at 6.4 mmol g-1, while its CO2/N2 selectivity is 64 (298 K, 1 bar), surpassing multiple other reported MOFs[100]. Importantly, in the breakthrough experiments, the CO2 adsorption capacity of Fe-dbai at 60% RH (Relative Humidity: the percentage of water vapor pressure in air to the saturated water vapor pressure at the same temperature) was able to maintain 94% of its capacity under dry conditions. Molecular simulation results showed that the amide CO-group, with its electronegative properties, exhibits a strong affinity towards CO2 and enhances the interaction between Fe-UMS and CO2 [Figure 4]. The outstanding CO2 capture efficiency of Fe-dbai suggests its potential suitability for real-world implementation of CO2 capture.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 4. Adsorption behavior and experimental column breakthrough curves of Fe-dbai. Reproduced with permission from Tu et al.[100]. Copyright 2023 American Chemical Society.

Amino acid (AA)-modified MOFs also show great potential for CO2 capture applications. Modification of MOF-808 with 11 different AAs resulted in a series of MOF-808-AA structures [Figure 5]. Under fume conditions, MOF-808 functionalized with glycine and DL-lysine (MOF-808-Gly and MOF-808-DL-Lys) was observed to exhibit the greatest CO2 adsorption capacity. The increased CO2 capture efficiency in the presence of water was detected and analyzed by single-component adsorption isotherms, CO2/H2O dichotomy isotherms, and dynamic breakthrough measurements. This study enhances our comprehension of CO2 capture in MOFs by uncovering the mechanism in which amine groups, firmly attached to the MOF structure, generate molecules within the pores that facilitate the adsorption and desorption of CO2 at relatively low temperatures without requiring any heating[101].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 5. Structure of MOF-808-AA and structural schemes of the coordinatively loaded amino acids. Reproduced with permission from Lyu et al.[101]. Copyright 2022 American Chemical Society.

Among physisorption materials, anion-functionalized MOFs are novel porous materials composed of metal moieties, organic linkers, and inorganic anions[102,103]. In recent years, anionic-functionalized MOFs have gradually made their presence felt in the field of CO2 capture. The reticular design approach can effectively regulate the pore chemistry of anionic-functionalized MOFs by utilizing molecular building units. As previously reported by Bhatt et al., NbOFFIVE-1-Ni and SIFSIX-3-Cu exhibit CO2 adsorption capacities of 1.3 and 1.2 mmol g-1 at a low concentration of 400 ppm and 298 K, respectively[104]. Their study also shows that reducing the pore size of the porous adsorbent facilitates enhanced interaction between the CO2 molecules and the main framework, resulting in high CO2 uptake at lower pressures.

ZU-16-Co is an anion-functionalized MOF with fine-tuned pore chemistry featuring one-dimensional (1D) pores modified by enriched F atoms that can effectively trap CO2 at concentrations of 400-10,000 ppm. Highly organized Lewis basic sites of anions limited to the ultramicroporous pores substantially enhance the ability to bind CO2 [Figure 6]. This work clarifies the structure-function relationship of ZU-16-Co in capturing CO2 and demonstrates its suitability for decarbonization at low concentrations[105].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 6. (A) Schematic representation of the construction and the pore structure of ZU-16 (TIFSIX-3) materials with pyrazine linker; (B) CO2 adsorption isotherms on various anion-functionalized ultramicroporous materials at 298 K; (C and D) CO2/N2 (1/99, flow rate: 5 mL min-1) and CO2/CH4 (50/50, flow rate: 4 mL min-1) conducted on ZU-16-Co. Reproduced with permission from Zhang et al.[105]. Copyright 2021 Science China Press and Springer-Verlag GmbH Germany, part of Springer Nature.

Two polar sulfonated oxygen-rich 3D MOFs, {[Zn2(TPOM)(3,7-DBTDC)2] 7H2O·DMA}n (1) and {[Cd2(TPOM)(3,7-DBTDC)2]·6H2O·3DMF}n (2) (TPOM = tetrakis(4-pyridyloxymethylene)-methane), were synthesized by a solvothermal approach by Chakraborty et al.[106]. Structural diversity was achieved by changing the metal centers. Due to the strong quadrupole interaction between the sulfone moiety and CO2 molecules, the adsorption of CO2 on 1 and 2 is highly selective over the adsorption of N2 and CH4[106,107]. Furthermore, both frameworks exhibit high chemical and water stability and cyclic regeneration [Figure 7]. This work provides an effective route for the development of functionalized MOFs with high selectivity for CO2 capture.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 7. Single-crystal X-ray structure of 1 (A) and 2 (B). Reproduced with permission from Chakraborty et al.[106]. Copyright 2020 American Chemical Society.

The pore windows and pore sizes of MOFs are made up of both organic and inorganic structural blocks. Therefore, the pore size of MOFs can be adjusted by changing the type of organic and inorganic structural blocks, allowing the pore structure of MOFs to vary considerably in size, which is one of the key reasons for the wide variety of MOF materials currently available. Due to the wide range of pore sizes of MOFs, they show great potential for different kinds of gas capture applications[108,109]. Selective adsorption of CO2 can be achieved by using large linkers, short ligands, interpenetrating networks, and smaller metal molecules.

A novel copper-based ultramicroporous MOF, Cu(adci)-2 (adci = 2-amino-4,5-dicyanoimidazole), was proposed by Jo et al.[110]. This MOF is a CO2 capture-oriented physical adsorbent synthesized by two strategies: performing aromatic amine functionalization and introducing ultramicropores. The Cu(adci)-2 structure has one-dimensional square channels where all of the auxiliary ligands, particularly the NH2 group at the 2 position of the imidazole ring, are oriented in the identical direction in each channel so that pairs of NH2 groups face away from each other along contrary sides of the channel walls. cu(adci)-2 shows higher CO2 adsorption capacity (2.01 mmol at 298 K and 15 kPa g-1) but a lower adsorption enthalpy at zero coverage (27.5 kJ mol-1). It showed good selectivity and easy regeneration under both dry and humid conditions [Figure 8].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 8. (A) Rietveld plot of desolvated Cu(adci)-2; (B) Perspective views of the refined structure of Cu(adci)-2 along the c (left) and b (right) axes; (C) Top and side views of a 1D channel of Cu(adci)-2. Reproduced with permission from Jo et al.[110]. Copyright 2022 American Chemical Society.

CALF-20 consists of 1,2,4-triazole-bridged zinc(II) ion layers supported by oxalate ionic pillars forming a 3D lattice and a 3D pore structure [Figure 9]. The crystallographically unique zinc center is five-coordinate with a distorted triangular bipyramidal geometry. The competitive separation on CALF-20 shows not only preferential physisorption of CO2 below 40% RH but also inhibition of water adsorption by CO2, which is confirmed by computational modeling. Furthermore, CALF-20 facilitates industrial-scale CO2 capture in a cost-effective and reliable manner[111]. This shows that the reasonable addition of anionic column bracing can effectively inhibit water adsorption, which, in turn, makes the adsorbent exhibit an excellent CO2 capture capacity.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 9. (A-C) Single-crystal structure of CALF-20; (D) Powder X-ray pattern simulated from the single-crystal structure (top) and obtained experimentally. Reproduced with permission from Lin et al.[111]. Copyright 2021 American Association for the Advancement of Science.

Efficient and sustainable CO2 capture can be achieved by porous physical adsorbents with high void fraction whose sizes and electrostatic potentials complementary to CO2 molecules[112]. Qazvini et al. proposed a strong, recoverable, and affordable adsorbent called MUF-16 [Figure 10][113]. MUF-16(Co), MUF-16(Ni), and MUF-16(Mn) were prepared by mixing 5-amino-m-m-phthalic acid (H2aip), a cheap, commercially available ligand, with cobalt (II), nickel (II), or manganese (II) salts in methanol. Through static adsorption curves, IAST, and density functional theory calculations, it is determined that the one-dimensional channel of MUF-16 can capture CO2 with high affinity while demonstrating weaker affinities for other competitive gases such as CH4, C2H2, C2H4, C2H6, C3H6, and C3H8. Therefore, MUF-16 has high CO2 adsorption selectivity. The selectivity of MUF-16 to CO2/CH4 and CO2/C2H2 is measured at 6,690 and 510, respectively.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 10. (A-D) Synthesis and structure of MUF-16 materials; (E) Volumetric adsorption (filled circles) and desorption (open circles) isotherms of CO2 at 293 K and for MUF-16 (black), MUF-16(Mn) (red), and MUF-16(Ni) (blue); (F) Adsorption enthalpy (Qst) calculated for CO2 binding to MUF-16 (black), MUF-16(Mn) (red), and MUF-16(Ni) (blue) as a function of CO2 uptake. Reproduced with permission from Qazvini et al.[113]. Copyright 2021 Springer Nature.

Open metal site modification strategy

Besides the functionalization of MOFs, metal sites have an important impact on enhancing the capacity and selectivity of CO2 relative to other gases. UMSs are usually partially positively charged, and these sites show an affinity for larger quadrupole moments and greater polarizability for CO2 compared to N2. It is these open sites that lead to high CO2 uptake. Furthermore, the difference in intensity between CO2 and other gas molecules is the driving force for CO2 capture[114-116].

MIL-101 has two dissimilar mesopores and distinct metal sites in a single local pore. Shin et al. demonstrated by adsorption isotherm combined with in situ crystallographic analysis that substrate-adsorbate interactions influence the initial adsorption and pore coalescence steps[117].

Unsaturated alkali metal sites have been reported to be anchored in MOFs by tetrazolium-based patterning to improve gas affinity[118]. In NKU-521 (NKU denotes Nankai University), Li et al. effectively embedded K+ cations into the trinuclear Co2+-tetrazolium coordination pattern[119]. The embedded K+ sites were exposed in the pores of NKU-521 by dehydration, and the Qst of CO2 increased to 41 kJ mol-1. The K+ cations actually act as gas traps and increase the CO2-framework affinity, as measured by the Qst, by 24%. Furthermore, the effect of unsaturated alkali metal sites in MOFs on hydrocarbon separation was investigated. IAST calculations and breakthrough experiments showed that the exposed K+ sites greatly improved the CO2 capture and separation performance of this MOF [Figure 11].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 11. Structural change of the embedded K+ from H2O coordinated state to the exposed state as CO2 traps for the preferential binding of CO2. Reproduced with permission from Li et al.[119]. Copyright 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

UMSs and adsorbent polarity play an important role in CO2 adsorption. MIL-88 was selected as the prototyping framework to verify the above strategy[120]. By introducing the C3 symmetry of the second ligand 2,4,6-tri(4-pyridyl)-1,3,5-triazine (tpt), the initial hexagonal 1D pore of the MIL-88 is split into numerous hexagonal cages of a certain length. As a direct result, MIL-88 loses metal vacancies and is replaced by split ligands. Moreover, the framework becomes more rigid, and there is no respiration effect before and after gas adsorption. The introduction of the second ligand (tpt) not only provides more action sites but also enhances the interaction area between objects. In addition, the group also modifies the terephthalic acid that constitutes the framework, introduces -OH and -NH2 groups, and increases the interaction sites between gas molecules and the framework. The CO2 adsorption test shows that CPM-33b with tpt and -OH introduced has the highest CO2 adsorption value compared to the existing MOFs without metal vacancies, which is comparable to the adsorption effect of MOF-74-Ni [Figure 12]. This study provides a new idea for MOFs to improve gas adsorption performance and is expected to achieve greater breakthroughs in fuel molecule adsorption.

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 12. (A and B) Illustration of pore space partition through symmetry-matching regulated ligand insertion; (C-F) Gas sorption study on compounds CPM-33a and CPM -33b. Reproduced with permission from Zhao et al.[120]. Copyright 2015 American Chemical Society.

Other functionalization

In addition to these common CO2 capture strategies, scientists have also explored other ways to efficiently capture CO2. In a typical Langmuir-type isotherm, bulk CO2 adsorbents cannot be regenerated during desorption because the adsorption gain decreases with increasing temperatures. Instead, adsorbents with S-shaped CO2 isotherms are preferable. Some flexible MOFs can display such S-shaped CO2 equivalent due to the structural switch from the CP (close-phase) state to the OP (open-phase) state[121-123]. Furthermore, as an individual guest molecule has its own gate opening pressure, this difference allows for high selectivity in gas mixtures, i.e., the target molecule can open the gate while others cannot[124]. Therefore, flexible MOFs with S-shaped isotherms can be used as potential carbon capture adsorbents with good operability and high selectivity for CO2.

ZnDatzBdc (Datz = 3,5-diamine-1,2,4-triazolate, Bdc = 1,4-benzenedicarboxylate) is a flexible MOF for highly selective capture of CO2 [Figure 13]. X-ray diffraction studies confirmed a convertible structure conversion between its OP and CP states. Importantly, ZnDatzBdc exhibits an S-shaped CO2 isotherm, yielding appreciable CO2 workability of 94.9 cm3 cm-3 under typical PVSA operations at 273 K, superior to most reported flexible MOFs[125].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 13. (A) Single-component CO2 isotherms on ZnDatzBdc up to 100 kPa and varied temperatures; (B) adsorption isotherms of CO2, N2, and CH4 up to 100 kPa at 273 and 298 K; (C) CO2 working capacities for the step-shaped isotherm of ZnDatzBdc and the simulated Langmuir isotherm from the CO2 adsorption data in the open phase, for a cycle of adsorption at 100 kPa and desorption at 20 kPa (273 K); (D) comparison of CO2 uptakes and working capacities of ZnDatzBdc and other flexible MOFs with S-shaped CO2 isotherms. Reproduced with permission from Peng et al.[125]. Copyright 2022 American Chemical Society.

Mechanochemistry at the chemical reaction level mainly refers to the application of mechanical energy to condensed substances, such as solids and liquids, by means of shear, abrasion, impact, and extrusion to induce changes in their structure and physicochemical properties and to induce chemical reactions. Unlike ordinary thermochemical reactions, mechanization, the driving force of the reaction, is mechanical energy rather than thermal energy; thus, the reaction can be completed without high temperature, high pressure, and other harsh conditions. It has been developing rapidly. This approach not only provides a green and energy-efficient route for chemical transformations but also offers more possibilities for the expanded preparation of materials[126]. Ultramicroporous MOFs with negatively charged anionic columns, such as SiF62- and TiF62-, often exhibit promising applications in gas separation and CO2 capture. As shown in Figure 14, a GeF62- pillared ultramicroporous MOF, ZU-36-Ni (also known as GeFSIX-3-Ni, GeFSIX = GeF62-), was prepared for the first time by sphere mill auxiliary conversion methodology by Zhang et al.[102]. The strong binding affinity of GeF62- and the comparatively long Ge-F bond distance (1.83 Å) lead to increased electronegativity and pore size contraction, resulting in high capacity and enhanced selectivity for trapping trace amounts of CO2 (1.07 mmol g-1 at 400 ppm).

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 14. (A and B) Adsorption isotherms of CO2, CH4, and N2 on ZU-36 materials at 298 K. (C) Comparison of the CO2 uptake on various materials at 400 ppm. (D) IAST selectivity of CO2/N2 (15/85) and CO2/CH4 (50/50) on ZU-36 materials; (E) Qst value of CO2 on ZU-36 materials. (F) DFT calculated binding site of CO2 in ZU-36-Ni. Reproduced with permission from Zhang et al.[102]. Copyright 2020 Royal Society of Chemistry.

A winning path to high levels of connectivity MOFs was proposed by Cairns et al. in 2007 by using metal-organic polyhedra (MOPs) as supramolecular building blocks (SBBs)[127]. NJU-Bai52 and NJU-Bai53, prepared by a pure-supramolecular-linker (PSL) approach, are two kinds of high-connected isomeric MOFs with a rare (3,3,6,6)-c topology[128]. Among them, TPBTM acts as a 12-connected supramolecular linker connected by the M3O cluster, forming two height-connected linkers. The equilibrium charge on these M3O clusters was delicately tuned from Cl- ions to monodentate hydroxide anions, resulting in a significant ~50-fold increase in the CO2 uptake for NJU-Bai53 (2.74 wt%) compared to NJU-Bai52 (0.74 wt%) at 298 K and 0.4 mbar. At 298 K and 0.15 bar, the CO2 uptake of NJU-Bai53 (7.67 wt%) was greatly increased compared to NJU-Bai52 (0.057 wt%), which is the highest among the reported amide-functionalized MOFs (AFMOFs). In addition, NJU-Bai53 exhibited higher selectivity and chemical stability [Figure 15].

Synthesis strategies of metal-organic frameworks for CO<sub>2</sub> capture

Figure 15. (A) TPBTM6- ligand self-assembles into PSL by strong π-π stacking and H-bonds, together with Fe3O clusters to construct NJU-Bai52, in which the P-isophthalates form metallamacrocycles and the C-isophthalates bridge these metallamacrocycles; (B) N2 adsorption and desorption isotherms at 77 K of NJU-Bai52 and NJU-Bai53; (C) CO2 adsorption isotherms at 298 K of NJU-Bai52 and NJU-Bai53; (D) CO2 adsorption enthalpies of NJU-Bai52 and NJU-Bai53; (E) breakthrough curves at 298 K of NJU-Bai52 and NJU-Bai53. Reproduced with permission from Song et al.[128]. Copyright 2019 American Chemical Society.

CONCLUSIONS

In conclusion, a great deal of CO2 capture research has been performed on MOFs in the past few decades. This is due to their special properties, such as large pore volume, high surface area, maintainability, structural diversity, etc. This review systematically describes several strategies that can be applied to design and synthesize CO2 capture-oriented MOFs, such as tuning the pore size window, functional group modification, and active site insertion. The emergence of these synthetic strategies offers enormous possibilities for the use of MOFs in practical applications in the area of CO2 capture and separation. Regardless of these strategies, it is imperative that certain characteristics of the original MOF are considered during the design phase, including the original functional groups, crystal structures, and acid and base properties of the MOF.

DECLARATIONS

Authors’ contributions

Conceptualization, investigation, and writing-original draft: Sun M

Editing: Wang X, Gao F, Xu M

Writing-review & editing, supervision, and funding acquisition: Fan W, Xu B, Sun D

Availability of data and materials

Not applicable.

Financial support and sponsorship

This work was supported by the National Natural Science Foundation of China (NSFC, Grant Nos. 22201305, 22275210), the Fundamental Research Funds for the Central Universities (22CX06024A, 23CX04001A), and the Outstanding Youth Science Fund Projects of Shandong Province (2022HWYQ-070).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2023.

REFERENCES

1. Sejas SA, Taylor PC, Cai M. Unmasking the negative greenhouse effect over the antarctic plateau. NPJ Clim Atmos Sci 2018;1:17.

2. Dong K, Dong X, Jiang Q, Zhao J. Assessing energy resilience and its greenhouse effect: a global perspective. Energy Econ 2021;104:105659.

3. Sullivan I, Goryachev A, Digdaya IA, et al. Coupling electrochemical CO2 conversion with CO2 capture. Nat Catal 2021;4:952-8.

4. Friedlingstein P, O'sullivan M, Jones MW, et al. Global carbon budget 2022. Earth Syst Sci Data 2022;14:4811-900.

5. Parekh A, Chaturvedi G, Dutta A. Sustainability analyses of CO2 sequestration and CO2 utilization as competing options for mitigating CO2 emissions. Sustain Energy Technol Assess 2023;55:102942.

6. Obama B. The irreversible momentum of clean energy. Science 2017;355:126-9.

7. Horike S, Kishida K, Watanabe Y, et al. Dense coordination network capable of selective CO2 capture from C1 and C2 hydrocarbons. J Am Chem Soc 2012;134:9852-5.

8. Wang Q, Bai J, Lu Z, Pan Y, You X. Finely tuning MOFs towards high-performance post-combustion CO2 capture materials. Chem Commun 2016;52:443-52.

9. Eddaoudi M, Sava DF, Eubank JF, Adil K, Guillerm V. Zeolite-like metal-organic frameworks (ZMOFs): design, synthesis, and properties. Chem Soc Rev 2015;44:228-49.

10. Wang S, Belmabkhout Y, Cairns AJ, et al. Tuning gas adsorption properties of zeolite-like supramolecular assemblies with gis topology via functionalization of isoreticular metal-organic squares. ACS Appl Mater Interfaces 2017;9:33521-7.

11. Zhou HC, Kitagawa S. Metal-organic frameworks (MOFs). Chem Soc Rev 2014;43:5415-8.

12. Guo B, Zhang J, Wang Y, Qiao X, Xiang J, Jin Y. Study on CO2 adsorption capacity and kinetic mechanism of CO2 adsorbent prepared from fly ash. Energy 2023;263:125764.

13. Pei J, Wang J, Shao K, et al. Engineering microporous ethane-trapping metal-organic frameworks for boosting ethane/ethylene separation. J Mater Chem A 2020;8:3613-20.

14. Gu XW, Wang JX, Wu E, et al. Immobilization of lewis basic sites into a stable ethane-selective MOF enabling one-step separation of ethylene from a ternary mixture. J Am Chem Soc 2022;144:2614-23.

15. Lv XL, Feng L, Xie LH, et al. Linker desymmetrization: access to a series of rare-earth tetracarboxylate frameworks with eight-connected hexanuclear nodes. J Am Chem Soc 2021;143:2784-91.

16. Lei Z, Xue Y, Chen W, et al. MOFs-based heterogeneous catalysts: new opportunities for energy-related CO2 conversion. Adv Energy Mater 2018;8:1801587.

17. Cui X, Chen K, Xing H, et al. Pore chemistry and size control in hybrid porous materials for acetylene capture from ethylene. Science 2016;353:141-4.

18. Fu HR, Jiang YY, Luo JH, Li T. A robust heterometallic Cd(II)/Ba(II)-organic framework with exposed amino group and active sites exhibiting excellent CO2/CH4 and C2H2/CH4 separation. Chin J Struct Chem 2022;41:2203287-92. Available from: https://www.sciencedirect.com/science/article/abs/pii/S0254586123002180 [Last accessed on 11 Aug 2023].

19. Usman M, Iqbal N, Noor T, et al. Advanced strategies in metal-organic frameworks for CO2 capture and separation. Chem Rec 2022;22:e202100230.

20. Zhang X, Wang X, Fan W, Sun D. Pore-environment engineering in multifunctional metal-organic frameworks. Chin J Chem 2020;38:509-24.

21. Bhardwaj A, Kaur J, Wuest M, Wuest F. In situ click chemistry generation of cyclooxygenase-2 inhibitors. Nat Commun 2017;8:1.

22. Palluet A, Lique F. Fine-structure excitation of CCS by He: potential energy surface and scattering calculations. J Chem Phys 2023;158:044303.

23. Harvey S, Hopkins J, Kuehl H, O'brien S, Mateeva A. Quest CCS facility: time-lapse seismic campaigns. Int J Greenh Gas Control 2022;117:103665.

24. Li J, Ma Y, Mccarthy MC, et al. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord Chem Rev 2011;255:1791-823.

25. D’Alessandro DM, Smit B, Long JR. Carbon dioxide capture: prospects for new materials. Angew Chem Int Ed 2010;49:6058-82.

26. Fominykh S, Stankovski S, Markovic VM, Petrovic D, Osmanović S. Analysis of CO2 migration in horizontal saline aquifers during carbon capture and storage process. Sustain 2023;15:8912.

27. Kirchon A, Feng L, Drake HF, Joseph EA, Zhou HC. From fundamentals to applications: a toolbox for robust and multifunctional MOF materials. Chem Soc Rev 2018;47:8611-38.

28. Jansen D, Gazzani M, Manzolini G, Dijk EV, Carbo M. Pre-combustion CO2 capture. Int J Greenh Gas Control 2015;40:167-87.

29. Kheirinik M, Ahmed S, Rahmanian N. Comparative techno-economic analysis of carbon capture processes: pre-combustion, post-combustion, and oxy-fuel combustion operations. Sustain 2021;13:13567.

30. Maffei T, Khatami R, Pierucci S, Faravelli T, Ranzi E, Levendis YA. Experimental and modeling study of single coal particle combustion in O2/N2 and oxy-fuel (O2/CO2) atmospheres. Combust Flame 2013;160:2559-72.

31. Cau G, Tola V, Ferrara F, Porcu A, Pettinau A. CO2-free coal-fired power generation by partial oxy-fuel and post-combustion CO2 capture: techno-economic analysis. Fuel 2018;214:423-35.

32. Sircar S, Golden TC. Purification of hydrogen by pressure swing adsorption. Sep Sci Technol 2000;35:667-87.

33. Chao C, Deng Y, Dewil R, Baeyens J, Fan X. Post-combustion carbon capture. Renew Sustain Energ 2021;138:110490.

34. Dinca C, Slavu N, Badea A. Benchmarking of the pre/post-combustion chemical absorption for the CO2 capture. J Energy Inst 2018;91:445-56.

35. Na S, Hwang SJ, Kim H, Baek I, Lee KS. Modeling of CO2 solubility of an aqueous polyamine solvent for CO2 capture. Chem Eng Sci 2019;204:140-50.

36. Kárászová M, Zach B, Petrusová Z, et al. Post-combustion carbon capture by membrane separation, review. Sep Purif Technol 2020;238:116448.

37. Liu M, Nothling MD, Webley PA, Jin J, Fu Q, Qiao GG. High-throughput CO2 capture using PIM-1@MOF based thin film composite membranes. Chem Eng J 2020;396:125328.

38. Wang Z, Ren H, Zhang S, Zhang F, Jin J. Polymers of intrinsic microporosity/metal-organic framework hybrid membranes with improved interfacial interaction for high-performance CO2 separation. J Mater Chem A 2017;5:10968-77.

39. Witt A, Pozzi R, Diesch S, Hädicke O, Grammel H. New light on ancient enzymes - in vitro CO2 fixation by pyruvate synthase of desulfovibrio africanus and sulfolobus acidocaldarius. FEBS J 2019;286:4494-508.

40. Hefti M, Joss L, Bjelobrk Z, Mazzotti M. On the potential of phase-change adsorbents for CO2 capture by temperature swing adsorption. Faraday Discuss 2016;192:153-79.

41. Mesfer MK, Danish M, Fahmy YM, Rashid MM. Post-combustion CO2 capture with activated carbons using fixed bed adsorption. Heat Mass Transfer 2018;54:2715-24.

42. Gutierrez-ortega A, Nomen R, Sempere J, Parra J, Montes-morán M, Gonzalez-olmos R. A fast methodology to rank adsorbents for CO2 capture with temperature swing adsorption. Chem Eng J 2022;435:134703.

43. Rehman A, Farrukh S, Hussain A, Fan X, Pervaiz E. Adsorption of CO2 on amine-functionalized green metal-organic framework: an interaction between amine and CO2 molecules. Environ Sci Pollut Res Int 2019;26:36214-25.

44. Choi S, Drese JH, Jones CW. Adsorbent materials for carbon dioxide capture from large anthropogenic point sources. ChemSusChem 2009;2:796-854.

45. Yi H, Li F, Ning P, et al. Adsorption separation of CO2, CH4, and N2 on microwave activated carbon. Chem Eng J 2013;215:635-42.

46. Jung Y, Ko YG, Nah IW, Choi US. Designing large-sized and spherical CO2 adsorbents for highly reversible CO2 capture and low pressure drop. Chem Eng J 2022;427:131781.

47. Wahby A, Silvestre-albero J, Sepúlveda-escribano A, Rodríguez-reinoso F. CO2 adsorption on carbon molecular sieves. Microporous Mesoporous Mater 2012;164:280-7.

48. Li S, Gallucci F. CO2 capture and activation with a plasma-sorbent system. Chem Eng J 2022;430:132979.

49. Zhang Q, Gao S, Yu J. Metal sites in zeolites: synthesis, characterization, and catalysis. Chem Rev 2023;123:6039-106.

50. Su F, Lu C. CO2 capture from gas stream by zeolite 13X using a dual-column temperature/vacuum swing adsorption. Energy Environ Sci 2012;5:9021.

51. Hong SH, Jang MS, Cho SJ, Ahn WS. Chabazite and zeolite 13X for CO2 capture under high pressure and moderate temperature conditions. Chem Commun 2014;50:4927-30.

52. Shang J, Li G, Singh R, Xiao P, Liu JZ, Webley PA. Determination of composition range for “molecular trapdoor” effect in chabazite zeolite. J Phys Chem C 2013;117:12841-7.

53. Remy T, Peter SA, Van Tendeloo L, et al. Adsorption and separation of CO2 on KFI zeolites: effect of cation type and Si/Al ratio on equilibrium and kinetic properties. Langmuir 2013;29:4998-5012.

54. Fiuza RA Jr, Medeiros de Jesus Neto R, Correia LB, Carvalho Andrade HM. Preparation of granular activated carbons from yellow mombin fruit stones for CO2 adsorption. J Environ Manage 2015;161:198-205.

55. Xu D, Xiao P, Zhang J, et al. Effects of water vapour on CO2 capture with vacuum swing adsorption using activated carbon. Chem Eng J 2013;230:64-72.

56. Ghazvini M, Vahedi M, Najafi Nobar S, Sabouri F. Investigation of the MOF adsorbents and the gas adsorptive separation mechanisms. J Environ Chem Eng 2021;9:104790.

57. Liu J, Thallapally PK, McGrail BP, Brown DR, Liu J. Progress in adsorption-based CO2 capture by metal-organic frameworks. Chem Soc Rev 2012;41:2308-22.

58. Zheng D, Yu Q, Heng Y, Cheng PL. Recent advances in C2 gases separation and purification by metal-organic frameworks. Chin J Struct Chem 2022;41:2211031-44. Available from: https://www.sciencedirect.com/science/article/abs/pii/S0254586123002441 [Last accessed on 11 Aug 2023].

59. Wong-Foy AG, Matzger AJ, Yaghi OM. Exceptional H2 saturation uptake in microporous metal-organic frameworks. J Am Chem Soc 2006;128:3494-5.

60. Eddaoudi M, Kim J, Rosi N, et al. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science 2002;295:469-72.

61. Salahuddin U, Iqbal N, Noor T, et al. ZIF-67 Derived MnO2 doped electrocatalyst for oxygen reduction reaction. Catalysts 2021;11:92.

62. Zhu C, Zhao B, Takata M, Aoki Y, Habazaki H. Biomass derived porous carbon for superior electrocatalysts for oxygen reduction reaction. J Appl Electrochem 2023;53:1379-88.

63. Yaqoob L, Noor T, Iqbal N, Nasir H, Mumtaz A. Electrocatalytic performance of NiNH2BDC MOF based composites with rGO for methanol oxidation reaction. Sci Rep 2021;11:13402.

64. Usman M, Ali M, Al-Maythalony BA, et al. Highly efficient permeation and separation of gases with metal-organic frameworks confined in polymeric nanochannels. ACS Appl Mater Interfaces 2020;12:49992-50001.

65. Jiang H, Jia J, Shkurenko A, et al. Enriching the reticular chemistry repertoire: merged nets approach for the rational design of intricate mixed-linker metal-organic framework platforms. J Am Chem Soc 2018;140:8858-67.

66. Ming Y, Purewal J, Yang J, et al. Kinetic stability of MOF-5 in humid environments: impact of powder densification, humidity level, and exposure time. Langmuir 2015;31:4988-95.

67. Gu Y, Wang Y, Zhao S, et al. N-donating and water-resistant Zn-carboxylate frameworks for humid carbon dioxide capture from flue gas. Fuel 2023;336:126793.

68. Li JR, Sculley J, Zhou HC. Metal-organic frameworks for separations. Chem Rev 2012;112:869-932.

69. Zhang Z, Yao Z, Xiang S, Chen B. Perspective of microporous metal-organic frameworks for CO2 capture and separation. Energy Environ Sci 2014;7:2868.

70. Samanta A, Zhao A, Shimizu GKH, Sarkar P, Gupta R. Post-combustion CO2 capture using solid sorbents: a review. Ind Eng Chem Res 2012;51:1438-63.

71. Zhang J, Singh R, Webley PA. Alkali and alkaline-earth cation exchanged chabazite zeolites for adsorption based CO2 capture. Microporous Mesoporous Mater 2008;111:478-87.

72. Zhou C, Li H, Qin H, et al. Defective UiO-66-NH2 monoliths for optimizing CO2 capture performance. Chem Eng J 2023;467:143394.

73. Patel HA, Byun J, Yavuz CT. Carbon dioxide capture adsorbents: chemistry and methods. ChemSusChem 2017;10:1303-17.

74. Yu H, Li B, Liu S, et al. Three new copper(II) coordination polymers constructed from isomeric sulfo-functionalized phthalate tectonics: synthesis, crystal structure, photocatalytic and proton conduction properties. J Solid State Chem 2021;294:121860.

75. O’Keeffe M, Yaghi OM. Deconstructing the crystal structures of metal-organic frameworks and related materials into their underlying nets. Chem Rev 2012;112:675-702.

76. Stock N, Biswas S. Synthesis of metal-organic frameworks (MOFs): routes to various MOF topologies, morphologies, and composites. Chem Rev 2012;112:933-69.

77. Zhu QL, Xu Q. Metal-organic framework composites. Chem Soc Rev 2014;43:5468-512.

78. Bai Y, Dou Y, Xie LH, Rutledge W, Li JR, Zhou HC. Zr-based metal-organic frameworks: design, synthesis, structure, and applications. Chem Soc Rev 2016;45:2327-67.

79. Kalmutzki MJ, Hanikel N, Yaghi OM. Secondary building units as the turning point in the development of the reticular chemistry of MOFs. Sci Adv 2018;4:eaat9180.

80. Moghadam PZ, Li A, Wiggin SB, et al. Development of a cambridge structural database subset: a collection of metal-organic frameworks for past, present, and future. Chem Mater 2017;29:2618-25.

81. Farrusseng D, Aguado S, Pinel C. Metal-organic frameworks: opportunities for catalysis. Angew Chem Int Ed 2009;48:7502-13.

82. Morris RV, Ruff SW, Gellert R, et al. Identification of carbonate-rich outcrops on Mars by the Spirit rover. Science 2010;329:421-4.

83. Suh MP, Park HJ, Prasad TK, Lim DW. Hydrogen storage in metal-organic frameworks. Chem Rev 2012;112:782-835.

84. He Y, Zhou W, Qian G, Chen B. Methane storage in metal-organic frameworks. Chem Soc Rev 2014;43:5657-78.

85. Masoomi MY, Morsali A, Dhakshinamoorthy A, Garcia H. Mixed-metal MOFs: unique opportunities in metal-organic framework (MOF) functionality and design. Angew Chem Int Ed 2019;58:15188-205.

86. Ahmad A, Khan S, Tariq S, Luque R, Verpoort F. Self-sacrifice MOFs for heterogeneous catalysis: synthesis mechanisms and future perspectives. Mater Today 2022;55:137-69.

87. Small LJ, Henkelis SE, Rademacher DX, et al. Near-zero power MOF-based sensors for NO2 detection. Adv Funct Mater 2020;30:2006598.

88. Hausdorf S, Baitalow F, Böhle T, Rafaja D, Mertens FO. Main-group and transition-element IRMOF homologues. J Am Chem Soc 2010;132:10978-81.

89. Titi HM, Marrett JM, Dayaker G, et al. Hypergolic zeolitic imidazolate frameworks (ZIFs) as next-generation solid fuels: unlocking the latent energetic behavior of ZIFs. Sci Adv 2019;5:eaav9044.

90. Yang S, Li X, Zeng G, et al. Materials Institute Lavoisier (MIL) based materials for photocatalytic applications. Coord Chem Rev 2021;438:213874.

91. Yang J, Yang K, Zhu X, et al. Band engineering of non-metal modified polymeric carbon nitride with broad spectral response for enhancing photocatalytic CO2 reduction. Chem Eng J 2023;461:141841.

92. Zhang G, Wei G, Liu Z, Oliver SRJ, Fei H. A robust sulfonate-based metal-organic framework with permanent porosity for efficient CO2 capture and conversion. Chem Mater 2016;28:6276-81.

93. Liu Y, Wang ZU, Zhou H. Recent advances in carbon dioxide capture with metal-organic frameworks. Greenhouse Gas Sci Technol 2012;2:239-59.

94. Quílez-bermejo J, Melle-franco M, San-fabián E, Morallón E, Cazorla-amorós D. Towards understanding the active sites for the ORR in N-doped carbon materials through fine-tuning of nitrogen functionalities: an experimental and computational approach. J Mater Chem A 2019;7:24239-50.

95. Serra-crespo P, Ramos-fernandez EV, Gascon J, Kapteijn F. Synthesis and characterization of an amino functionalized MIL-101(Al): separation and catalytic properties. Chem Mater 2011;23:2565-72.

96. Dinakar B, Forse AC, Jiang HZH, et al. Overcoming metastable CO2 adsorption in a bulky diamine-appended metal-organic framework. J Am Chem Soc 2021;143:15258-70.

97. Lu W, Sculley JP, Yuan D, Krishna R, Wei Z, Zhou H. Polyamine-tethered porous polymer networks for carbon dioxide capture from flue gas. Angew Chem Int Ed 2012;51:7480-4.

98. Khan J, Iqbal N, Asghar A, Noor T. Novel amine functionalized metal organic framework synthesis for enhanced carbon dioxide capture. Mater Res Express 2019;6:105539.

99. Mcdonald TM, D'alessandro DM, Krishna R, Long JR. Enhanced carbon dioxide capture upon incorporation of N,N’-dimethylethylenediamine in the metal-organic framework CuBTTri. Chem Sci 2011;2:2022-8.

100. Tu S, Yu L, Liu J, et al. Efficient CO2 capture under humid conditions on a novel amide-functionalized Fe-soc metal-organic framework. ACS Appl Mater Interfaces 2023;15:12240-7.

101. Lyu H, Chen OI, Hanikel N, et al. Carbon dioxide capture chemistry of amino acid functionalized metal-organic frameworks in humid flue gas. J Am Chem Soc 2022;144:2387-96.

102. Zhang Z, Ding Q, Peh SB, et al. Mechano-assisted synthesis of an ultramicroporous metal-organic framework for trace CO2 capture. Chem Commun 2020;56:7726-9.

103. Peng YL, Pham T, Li P, et al. Robust ultramicroporous metal-organic frameworks with benchmark affinity for acetylene. Angew Chem Int Ed 2018;57:10971-5.

104. Bhatt PM, Belmabkhout Y, Cadiau A, et al. A fine-tuned fluorinated MOF addresses the needs for trace CO2 removal and air capture using physisorption. J Am Chem Soc 2016;138:9301-7.

105. Zhang Z, Ding Q, Cui J, Cui X, Xing H. High and selective capture of low-concentration CO2 with an anion-functionalized ultramicroporous metal-organic framework. Sci China Mater 2021;64:691-7.

106. Chakraborty G, Das P, Mandal SK. Polar sulfone-functionalized oxygen-rich metal-organic frameworks for highly selective CO2 capture and sensitive detection of acetylacetone at ppb level. ACS Appl Mater Interfaces 2020;12:11724-36.

107. Lin RB, Chen D, Lin YY, Zhang JP, Chen XM. A zeolite-like zinc triazolate framework with high gas adsorption and separation performance. Inorg Chem 2012;51:9950-5.

108. Lin R, Xiang S, Zhou W, Chen B. Microporous metal-organic framework materials for gas separation. Chem 2020;6:337-63.

109. Fan W, Wang X, Zhang X, et al. Fine-tuning the pore environment of the microporous Cu-MOF for high propylene storage and efficient separation of light hydrocarbons. ACS Cent Sci 2019;5:1261-8.

110. Jo D, Lee SK, Cho KH, Yoon JW, Lee UH. An Amine-functionalized ultramicroporous metal-organic framework for postcombustion CO2 capture. ACS Appl Mater Interfaces 2022;14:56707-14.

111. Lin JB, Nguyen TTT, Vaidhyanathan R, et al. A scalable metal-organic framework as a durable physisorbent for carbon dioxide capture. Science 2021;374:1464-9.

112. Oschatz M, Antonietti M. A search for selectivity to enable CO2 capture with porous adsorbents. Energy Environ Sci 2018;11:57-70.

113. Qazvini OT, Babarao R, Telfer SG. Selective capture of carbon dioxide from hydrocarbons using a metal-organic framework. Nat Commun 2021;12:197.

114. Chowdhury P, Mekala S, Dreisbach F, Gumma S. Adsorption of CO, CO2 and CH4 on Cu-BTC and MIL-101 metal organic frameworks: Effect of open metal sites and adsorbate polarity. Microporous Mesoporous Mater 2012;152:246-52.

115. Kökçam-Demir Ü, Goldman A, Esrafili L, et al. Coordinatively unsaturated metal sites (open metal sites) in metal-organic frameworks: design and applications. Chem Soc Rev 2020;49:2751-98.

116. Lim D, Chyun SA, Suh MP. Hydrogen storage in a potassium-ion-bound metal-organic framework incorporating crown ether struts as specific cation binding sites. Angew Chem Int Ed 2014;126:7953-6.

117. Shin SR, Cho HS, Lee Y, et al. In Situ mapping and local negative uptake behavior of adsorbates in individual pores of metal-organic frameworks. J Am Chem Soc 2021;143:20747-57.

118. Chen K, Kang YS, Zhao Y, Yang JM, Lu Y, Sun WY. Cucurbit[6]uril-based supramolecular assemblies: possible application in radioactive cesium cation capture. J Am Chem Soc 2014;136:16744-7.

119. Li N, Chang Z, Huang H, et al. Specific K+ binding sites as CO2 traps in a porous MOF for enhanced CO2 selective sorption. Small 2019;15:e1900426.

120. Zhao X, Bu X, Zhai QG, Tran H, Feng P. Pore space partition by symmetry-matching regulated ligand insertion and dramatic tuning on carbon dioxide uptake. J Am Chem Soc 2015;137:1396-9.

121. Oh JM, Venters CC, Di C, et al. U1 snRNP regulates cancer cell migration and invasion in vitro. Nat Commun 2020;11:1.

122. Balogun HA, Bahamon D, Almenhali S, Vega LF, Alhajaj A. Are we missing something when evaluating adsorbents for CO2 capture at the system level? Energy Environ Sci 2021;14:6360-80.

123. Wang Q, Ke T, Yang L, et al. Cover picture: separation of Xe from Kr with record selectivity and productivity in anion-pillared ultramicroporous materials by inverse size-sieving. Angew Chem Int Ed 2020;59:3341.

124. Chen Y, Qiao Z, Lv D, et al. Efficient adsorptive separation of C3H6 over C3H8 on flexible and thermoresponsive CPL-1. Chem Eng J 2017;328:360-7.

125. Peng J, Liu Z, Wu Y, Xian S, Li Z. High-performance selective CO2 capture on a stable and flexible metal-organic framework via discriminatory gate-opening effect. ACS Appl Mater Interfaces 2022;14:21089-97.

126. Jiang Y, Tan P, Qi S, et al. Cover picture: metal-organic frameworks with target-specific active sites switched by photoresponsive motifs: efficient adsorbents for tailorable CO2 capture. Angew Chem Int Ed 2019;58:6457.

127. Cairns AJ, Perman JA, Wojtas L, et al. Supermolecular building blocks (SBBs) and crystal design: 12-connected open frameworks based on a molecular cubohemioctahedron. J Am Chem Soc 2008;130:1560-1.

128. Song X, Zhang M, Chen C, et al. Pure-supramolecular-linker approach to highly connected metal-organic frameworks for CO2 capture. J Am Chem Soc 2019;141:14539-43.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Sun M, Wang X, Gao F, Xu M, Fan W, Xu B, Sun D. Synthesis strategies of metal-organic frameworks for CO2 capture. Microstructures 2023;3:2023032. http://dx.doi.org/10.20517/microstructures.2023.32

AMA Style

Sun M, Wang X, Gao F, Xu M, Fan W, Xu B, Sun D. Synthesis strategies of metal-organic frameworks for CO2 capture. Microstructures. 2023; 3(4): 2023032. http://dx.doi.org/10.20517/microstructures.2023.32

Chicago/Turabian Style

Sun, Meng, Xiaokang Wang, Fei Gao, Mingming Xu, Weidong Fan, Ben Xu, Daofeng Sun. 2023. "Synthesis strategies of metal-organic frameworks for CO2 capture" Microstructures. 3, no.4: 2023032. http://dx.doi.org/10.20517/microstructures.2023.32

ACS Style

Sun, M.; Wang X.; Gao F.; Xu M.; Fan W.; Xu B.; Sun D. Synthesis strategies of metal-organic frameworks for CO2 capture. Microstructures. 2023, 3, 2023032. http://dx.doi.org/10.20517/microstructures.2023.32

About This Article

© The Author(s) 2023. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
719
Downloads
201
Citations
2
Comments
0
4

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 12 clicks
Like This Article 4 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Microstructures
ISSN 2770-2995 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/