Next Article in Journal
The Effect of Green Roofs and Green Façades in the Pedestrian Thermal Comfort of a Mediterranean Urban Residential Area
Previous Article in Journal
Optimization of Vibrating Mesh Nebulizer Air Inlet Structure for Pulmonary Drug Delivery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Advances on Nitrogen-Doped Carbon Materials in COx Hydrogenation

1
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing 211816, China
2
School of Energy Science and Engineering, Nanjing Tech University, Nanjing 211816, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2023, 14(10), 1510; https://doi.org/10.3390/atmos14101510
Submission received: 30 August 2023 / Revised: 19 September 2023 / Accepted: 27 September 2023 / Published: 29 September 2023
(This article belongs to the Section Air Pollution Control)

Abstract

:
The excessive consumption of fossil fuels has resulted in massive carbon emissions and serious ecological and environmental crises. Therefore, achieving the efficient utilization of waste carbon sources is considered as an important pathway to addressing the aforementioned issues in the context of carbon neutrality. Developing and designing suitable catalyst materials has become the key to converting COx into valuable platform chemicals and value-added liquid fuels (e.g., CO, CH4, CH3OH, and C2+ hydrocarbons). A moderate interaction between nitrogen-doped carbon materials and active metals is more favorable for the progress of the COx hydrogenation reaction compared to traditional metal oxide carriers. In this work, we comprehensively summarize the synthesis methods of N-doped carbon materials and the relevant research progress in the field of COx hydrogenation. In addition, a general assessment of carbon-based catalysts for COx hydrogenation reactions, concerning the support and metal properties, the activity and product selectivity, and their interactions is systematically discussed. Finally, this review discusses the roles of N-doped carbon materials, the current challenges, and future development directions.

1. Introduction

With the acceleration of industrialization, the demands for energy-related industrial chemicals and fuel is rapidly increasing. However, the production of the aforementioned chemicals and fuels is heavily reliant on traditional fossil energy (e.g., coal, oil, and natural gas), causing considerable CO2 emissions and a serious energy crisis. To mitigate the greenhouse effect and address the energy crisis, C1 chemistry has become an important and attractive research field in chemistry and catalysis [1,2,3]. Therefore, the conversion of C1 molecules (e.g., CO and CO2) into high-value chemicals has attracted widespread attention. The hydrogenation of the C1 molecule via the Fischer–Tropsch synthesis (FTS) pathway is considered as one of the promising directions for developing sustainable and clean alternative energy. Therefore, it is necessary to vigorously promote in-depth research on CO2 conversion and Fischer–Tropsch synthesis.
FTS reactions are regarded as the key to producing clean fuels and high-value-added chemicals via the syngas-mediated pathway [4], which can be produced from coal, biomass, and natural gas in the existing industry [1,5]. The essential steps in the FTS process are the adsorption and activation of CO and H2, as well as C-C coupling. Generally, the obtained products via FTS reactions follow the Anderson–Schulz–Flory (ASF) distribution [6]. The derived gasoline, diesel, and aviation kerosene can effectively reduce reliance on fossil fuels and provide possibilities for developing alternative energy sources. In addition, the generated low-carbon olefins and high-carbon alcohols have tremendous application prospects. Currently, the combustion of fossil fuels and human activities have led to a rapid increase in CO2 concentrations in the atmosphere. Hence, the raw materials of the FTS reactions have extended to the atmospheric CO2 to achieve their efficient utilization and mitigate climate change. Therefore, researchers have also conducted extensive studies in related fields. In this work, COx is used to replace CO2 and CO, and we review the state-of-the-art research progress in the field of COx hydrogenation, which poses a major challenge for carbon emissions reduction. Therefore, researchers have extensively explored CO2 capture, storage, and utilization technologies [7,8,9]. Currently, CO2 has become a strategic chemical raw material for transport fuels and various chemicals [10,11], and the diversification of its products (CH4, C2–C4, C5+, etc.) has led to extensive research on this reaction. CO2 hydrogenation is a concept for comprehensive carbon resource utilization with broad prospects. It can not only reduce a series of environmental issues caused by the greenhouse effect but also reduce excessive reliance on fossil fuels in social development [12,13].
The catalyst performances (e.g., catalytic activity, targeted product selectivity, and stability) of COx hydrogenation can be regulated by adjusting the catalyst composition and structure [14,15], support materials [16], and additives. Among the aforementioned factors, the specific surface area, pore structure, and interaction with the active components of the support materials significantly affect the dispersion of active components, the adsorption/desorption properties of reactants and products, the catalytic activity, product selectivity, and catalyst stability [17]. Particularly, the strong interaction between traditional metal oxide supports (SiO2, Al2O3, TiO2, MgO) and active components hinders their reduction, resulting in poor catalytic performance. For instance, the strong metal-support interaction between cobalt and Al2O3, SiO2, and TiO2 leads to the formation of Co2SiO2, Co2AlO4, or CoTiO4, respectively, which are reducible only at high temperatures [18]. Compared to traditional supports, carbon materials (activated carbon (AC), carbon nanotubes (CNTs), carbon nanofibers (CNFs), carbon spheres (CSs), etc.) have weak metal-support interactions, which might overcome these drawbacks [19,20]. And the advantages of carbon materials, such as their diverse morphological structures, high specific surface area, easily controllable surface properties through doping, and good stability, make them widely used as catalyst supports [21].
Inert carbon carriers can be modified through subtle structural modifications, thereby affecting catalytic performance. In this case, doping the carbon matrix with heteroatoms has become one of the ways to regulate catalytic systems [22]. In recent years, metals supported on nitrogen-doped carbon materials have shown excellent performance as heterogeneous catalysts. Nitrogen-doped carbon materials have origins dating back to the 1960s, but their use as catalyst supports has only emerged recently, leading to significant advancements in heterogeneous catalysis. The major advantage of nitrogen-doped carbon materials lies in the ability to modulate the interactions between the support and the metals through electron interactions induced by introducing heteroatoms on the carbon material surface [23,24]. Additionally, doping nitrogen atoms introduces vacancies and defects on the carbon material surface. These defects can serve as anchoring sites for active metals, enhancing the stability of the catalyst. N is the most explored element among various dopants (B, P, F, S), and nitrogen-doped carbon materials also exhibit a wide range of potential applications [25,26]. The introduction of N atoms can effectively anchor the metal components, increase the dispersion of active components, and reduce their agglomeration. Moreover, with the incorporation of the N element, alkaline sites are provided, which is favorable for the adsorption/activation of weakly acidic CO2. The N element also plays a role in weakening the C-O bond and enhancing the C-C bond, resulting in excellent catalytic performance in COx hydrogenation reactions [27,28,29].
Many precious metals and non-precious metals are currently loaded onto nitrogen-doped carbon materials and widely applied in the catalysis field (photocatalysis, electrocatalysis, thermocatalysis) [30,31]. Substantial efforts and progress have been made for FTS or CO2 hydrogenation over the past few decades [32,33,34,35,36,37]. They have systematically summarized the application of efficient catalysts in COx hydrogenation reactions. However, none of them have summarized the application of nitrogen-doped carbon materials as carriers in COx hydrogenation reactions, implying that a comprehensive overview on this emerging topic would have great merit. The surface physical and chemical properties of carbon materials can be changed after the introduction of nitrogen atoms, which is beneficial to make the desired products more dominant. Therefore, it is necessary to summarize the recent research progress in this field to provide feasible suggestions for the design of efficient catalysts. In this review, we first introduce the preparation methods of nitrogen-doped carbon materials and the key factors influencing the N species structure. Second, the application of nitrogen-doped carbon materials as catalyst supports in COx hydrogenation is discussed, summarizing the role of N species in the reaction (including strategies for regulating the proportion of nitrogen species, the impact of material surface properties, and a preliminary exploration of the reaction mechanism). Finally, it discusses the potential challenges in this field to provide some guidance for the development of advanced N-doped carbon materials.

2. N-doped Carbon Material and Preparation Methods

2.1. N Type of Doping Species

The nitrogen species in nitrogen-doped carbon materials can be mainly divided into four types: pyridinic nitrogen, pyrrolic nitrogen, graphitic nitrogen, and pyridinic nitrogen oxide. Characterization techniques are crucial for elucidating the physicochemical properties of materials and understanding their catalytic mechanisms. The characterization techniques for elucidating the role of nitrogen-doped carbon materials in COx hydrogenation reactions can be broadly categorized as follows [22,38]: The type of nitrogen species can be distinguished using XPS spectra. The pyridinic N peak appears around 398.6 eV, the pyrrolic N peak appears around 400.5 eV, and the graphitic N peak appears around 401.3 eV. Additionally, pyridinic N, pyrrolic N, and graphitic N species can also be distinguished using XANES and EELS spectra.
Based on the current experimental and theoretical research results on COx hydrogenation, the first three nitrogen species have been widely applied in the relevant fields due to their superior catalytic abilities.
In general, pyridinic nitrogen is located in the edge or defective hexagonal carbon layers. Furthermore, some studies have proposed that a pyridinic nitrogen atom is bonded to two sp2-hybridized carbon atoms. Plus, the pyridinic nitrogen atom provides a p for the π system. The increasing electron density promotes the nucleation and growth of nanoparticles on nitrogen-doped carbon supports, and then enhance the interaction between the metal and the support [39]. As for pyrrolic nitrogen, the atom is incorporated into a pentagonal ring, which can provide two p electrons to the π electron system. Due to the unstable pentagonal structure, pyrrolic nitrogen species can gradually transform into graphitic nitrogen species when the temperature is higher than 800 °C [40]. However, considering the geometric effects, the anchoring abilities of pyridinic nitrogen and pyrrolic nitrogen are greater than that of graphitic nitrogen [41].
The physicochemical properties (e.g., acidity/basicity, wettability, charge transfer rate, and electronic structure) of carbon materials would change after the introduction of nitrogen species. More specifically, pyridinic nitrogen species have unhybridized lone pair electrons, which provide a strong electron-donating ability and effective Lewis base sites. Nitrogen-doped carbon catalysts that are rich in pyridinic nitrogen species can enhance the adsorption of CO2 (Lewis acid) in CO2 hydrogenation reactions, improving catalytic performances. However, the alkalinity of the carbon material doped that it is doped with relatively weak due to the weak electron donation capability of graphite nitride species [42]. For example, Feng et al. [43] revealed that the introduction of nitrogen species into carbon materials can provide metal anchoring sites on material surfaces and improve the electronic properties of transition metal species using theoretical calculations. In terms of wettability, carbon materials with sp2 hybridized structures have strong hydrophilicity, and the introduction of nitrogen species can increase the proportion of sp2 hybridized structures in the material, thereby enhancing its hydrophilicity [44]. Additionally, the introduction of nitrogen species can generate functional groups on the surface of carbon materials, improving the interaction between the carbon material surface and reactant molecules or active species through dipole-dipole interactions, hydrogen bonding, or covalent bonding [45]. Moreover, the introduction of nitrogen species can significantly influence the charge distribution in carbon materials by utilizing the difference in electronegativity. Generally, untreated carbon materials have weak charges on their surfaces. However, the introduction of nitrogen species can greatly affect the charge distribution in carbon materials, resulting in the generation of vacancies or defect sites in the carbon structure. According to the work of Su et al. [46], it was found that the pyrrolic nitrogen species formed during the preparation of Fe-based nitrogen-doped carbon catalysts can introduce certain defect sites in the material, effectively improving the coordination environment of the Fe species and enhancing its ORR reaction properties. Furthermore, due to the delocalization effect, doped nitrogen atoms can alter the electronic structure of carbon materials at nitrogen doping sites, which leads to local charge accumulation and changes in the local electronic state density. This effectively promotes electron transfer reactions and an acceleration of the adsorption and/or activation of reactant molecules [47].

2.2. Preparation Methods of Nitrogen-Doped Carbon Materials

Nitrogen-doped carbon materials (i.e., nitrogen-doped graphene, nitrogen-doped porous carbon, and nitrogen-doped carbon nanofibers) chemically bond or incorporate nitrogen atoms into the carbon material framework. Given the similar size of nitrogen and carbon atoms, the substitution of nitrogen atoms for carbon atoms causes less damage to the carbon material framework, which maintains the stability of the carbon materials [48,49]. Therefore, nitrogen-doped carbon materials have been widely applied in the field of catalysis. Typically, nitrogen-doped carbon materials can be prepared through post-doping and in situ doping methods. Meanwhile, their different types and morphologies can be achieved by leveraging disparities in the precursor types, preparation processes, and the introduction of nitrogen components.

2.2.1. Post-Doping Method

The post-doping method, as a classical method for preparing nitrogen-doped carbon materials, has captured considerable interest among researchers owing to its straightforward operational nature. It can be used to yield the nitrogen contents and proportions of nitrogen species by varying the reaction, reaction time, nitrogen precursor species, and types of carbon materials. To be specific, it involves a reaction between pre-synthesized carbon materials and nitrogen precursor species (such as ammonia, urea, dicyandiamide, melamine, etc.) to prepare nitrogen-doped carbon materials under high temperatures. Typically, this method requires a long-term treatment under high temperatures to ensure the successful introduction of nitrogen species. Herein, the nitrogen contents and proportions of nitrogen species can be changed by varying the reaction temperature, reaction time, nitrogen precursor species, and types of carbon materials.
The treatment temperature is considered as a major factor influencing the nitrogen content of nitrogen-doped carbon materials. For example, when treating graphene oxide and activated carbon in an NH3 atmosphere between 500 and 600 °C, the nitrogen content is highest, and is more than twice the nitrogen content of samples calcined at temperatures above 800 °C [50,51]. Wang et al. also found that as the pyrolysis temperature increased from 500 °C to 800 °C [52], the nitrogen content remarkably decreased from 19.8 to 2.08%. It can be seen that the pyrolysis temperature in an ammonia atmosphere has a significant impact on the nitrogen content of the material. In addition to the nitrogen content, the treatment temperature also has a considerable influence on the types of nitrogen species. Wang’s research revealed that as the treatment temperature increased, the unstable structure of pyrrolic nitrogen resulted in a reduction in its content, and there was also a decreasing trend in the content of pyridinic nitrogen.
Compared to conventional high-temperature pyrolysis methods, the microwave-assisted method can reduce the preparation time for nitrogen-doped carbon materials to even be as short as a few seconds [29]. It also conforms to the current trend of environmental friendliness and has gradually gained attention in recent years [53]. In the microwave method, the appropriate nitrogen precursors (e.g., urea, ammonium hydroxide, and ammonium carbonate) generally have advantages such as low cost, being environmentally friendly, and low pyrolysis temperatures. For example, Ali et al. [54] used urea as the nitrogen precursor and synthesized a mixture of carbon nanotubes and urea under microwave conditions. The nitrogen-doped carbon material was obtained in about 10 min at a power of 700 W, where the nitrogen content was close to 7%, and the nitrogen species included pyridine, pyrrole, and various other groups.

2.2.2. In Situ Doping Method

The in situ doping method can be regarded as a one-step synthesis approach for nitrogen-doped carbon materials. In this method, the nitrogen source and carbon source precursor materials are added simultaneously during the preparation process, and then they are treated under a high-temperature condition in an inert atmosphere or via hydrothermal carbonization. Notably, the specific morphologies or structures can be flexibly adjusted by selecting different precursors and preparation conditions in this method [55,56,57]. Common precursors include polyaniline (PANI), polypyrrole (PPy), urea, biomass-based materials [58], and nitrogen-containing metal–organic framework materials (such as a zeolitic imidazolate framework) [59]. In the in situ doping method, preparation parameters such as the carbonization temperature and the type of nitrogen source precursor play crucial roles in determining the types of nitrogen species. For instance, the pyrolysis temperature significantly affects the nitrogen content, the ratio of the different nitrogen species, and the graphitization degree of the carbon material [60]. Therefore, synthesizing nitrogen-doped carbon materials with specific structures poses significant challenges in in situ synthesis.
N-doped carbon materials derived from biomass have attracted significant attention due to the unique pore structure generated during their preparation process. Geng et al. [61] utilized spirulina as a precursor for nanoporous biochar preparation via in situ nitrogen–oxygen heteroatom co-doping through a pre-carbonization coupling pyrolytic activation process. The effects of the activation temperature and the KOH mass ratio on the carbon properties were investigated. The results indicated that both the temperature and the amount of activator significantly affect the physical properties of the material, such as the pore distribution and the specific surface area. Su et al. [62] used birch bark as the carbon source for biochar preparation for the first time. N-doped porous biochar was synthesized with one-step co-activation pyrolysis by adding an activator (NaOH) and a nitrogen source (urea) during the preparation process. The results indicated that nitrogen doping not only provides more reactive binding sites on the surface of carbon materials, but also enhances the reactivity of biochar by adjusting the electronic structure of the material framework. Liu et al. [63] prepared novel N-doped biochar via the one-step pyrolysis of algal sludge without external nitrogen sources. High-temperature pyrolysis leads to a more enriched pore structure on the material surface, which is more favorable for the adsorption and decomposition of sulfadiazine. Additionally, high-temperature pyrolysis also increases the defects in the material. Glaydson et al.’s research also indicates that N doping significantly alters the pore structure of the material (primarily mesopores) and increases the surface defects of the material [64]. The enriched pore structure enables the loading of a large number of active components, and the increased defects also play a role in anchoring the active species. Therefore, nitrogen-doped carbon materials hold great potential in the field of catalyst supports. Among the various carbon sources, Pinus sylvestris is advantageous for producing biochar due to its high lignin content (25.5%). Wang et al. [65] used Pinus sylvestris as a carbon source and mixed it with urea (nitrogen source), then calcinated the mixture in a tube furnace under a N2 atmosphere. The obtained materials exhibited a pore structure with both micropores and mesopores at different pyrolysis temperatures. However, as the pyrolysis temperature increased from 500 °C to 700 °C, the pore size, pore volume, and specific surface area of the materials all showed an increasing trend, indicating the significant impact of the pyrolysis temperature on the final material structure. Lignin, as the second most abundant biomass resource in the plant kingdom (after cellulose), is often used as a carbon source for synthesizing nitrogen-doped carbon materials. Zhong et al. [66] mixed a solution containing lignin and melamine, and then pyrolyzed them in a tube furnace, generating materials with a thin carbon-sheet-like, three-dimensional structure. In addition to the aforementioned nitrogen sources, organic compounds such as polypyrrole and polyaniline contain nitrogen elements themselves, making direct pyrolysis easier compared to adding external nitrogen sources. For instance, Sudhakaran et al. [67] used polypyrrole to synthesize nitrogen-doped carbon nanotubes (NCNTs) via pyrolysis in a N2 atmosphere. It can be observed that these carbon nanotubes were composed of amorphous carbon, and nitrogen elements were uniformly distributed on the carbon nanotubes. Li et al. [68] used self-synthesized nanofiber polyaniline as a precursor, and mixed it with KOH (an activating agent). The nitrogen-doped porous carbon fibers were generated by pyrolyzing the sample in a N2 atmosphere. It was found that the large pores between the nanofibers served as efficient gas transport channels.
Zeolitic imidazolate framework materials (ZIFs) possess a high specific surface area, a large pore volume, and a rich microstructure. Therefore, they have been important precursors for the preparation of nitrogen-doped carbon materials, and have been a research hotspot in recent years. Based on the advantages of the ZIF precursors, nitrogen-doped carbon materials with different structures, such as nanosheets [55], nanoparticles [56], and core-shell structures [69], have been designed. For example, Zhao et al. [55] synthesized nanosheet-like ZIF-8 using water as a solvent, and the material still maintained its sheet-like structure after pyrolysis. Zhang et al. [70] synthesized ZnCo-ZIF materials with particle sizes of around 10 nm by adjusting the preparation parameters, and found that nitrogen-doped carbon nanosheets could be prepared during further carbonization. On the basis of the encapsulation effect of MOF materials on metal components, Shi et al. [69] successfully prepared nitrogen-doped carbon nanotubes coated with ZnO in a core–shell structure (ZnO@NCNT), and a further etching of ZnO with an acid solution formed hollow nitrogen-doped carbon nanofibers with diameters ranging from 70 to 190 nm. In addition, considering MOFs as precursors, as well as the unique interaction between metal species and organic species, the preparation of supported single-atom catalysts can be realized, and the structural properties of metal sites can be controlled. Zheng et al. [71] used a two-dimensional layered bimetallic Zn/Fe-MOF as a precursor and prepared Fe-supported nitrogen-doped carbon catalysts with a single-atom/single-atom pair Fe-N4 structure. The loading amount of the Fe species was nearly 4%, and the nitrogen content reached 10%.
In summary, the current methods for preparing nitrogen-doped carbon materials have their respective advantages, but there are still challenges that are difficult to control. These issues include a low sample yield, the inability to precisely construct nitrogen species, and the stringent preparation conditions that may damage the carbon material or the support material’s structure. Therefore, developing milder synthesis conditions while achieving an efficient and controllable preparation of catalytically active components is currently a key focus in the research on nitrogen-doped carbon materials.

3. Catalytic Application of Nitrogen-Doped Carbon Materials in CO Hydrogenation

With the advancement of industrialization, fossil energy is becoming increasingly depleted, and developing alternative clean energy is becoming urgent. Fischer–Tropsch synthesis (FTS) is the essential process of converting syngas into high-value-added hydrocarbons. Studies have found that the interaction between different supports (such as alumina, silica, and carbon material) and metals is crucial for FTS reactions. Carbon supports, with their high specific surface area, excellent thermal stability, and weak metal–support interactions [5], have been widely used. Introducing nitrogen atoms into carbon materials can further regulate the metal–support interaction, promoting the dispersion of the active components [72]. In recent years, the nitrogen doping of carbon materials has become a promising strategy to improve catalytic performance. Through a review of the current research literature, we find that nitrogen-doped carbon materials are mainly applied in the production of long-chain hydrocarbons in FTS reactions. The following content is a brief description of the aforementioned application, which is classified according to the main active components, Co and Fe.

3.1. Cobalt-Based Catalysts

Co-based catalysts have been widely used in FTS reactions due to their high C5+ selectivity, high single-pass conversion rate, and low CO2 selectivity. However, the strong interaction between the traditional metal oxide supports and the active components leads to a decrease in the utilization of the active phases [73]. Recently, nitrogen-doped carbon materials have shown excellent catalytic performance in FTS due to the formation of suitable and favorable metal–support interactions [74].
Kafrani et al. [75] impregnated Co and Ru onto a nitrogen-doped graphene aerogel (NGA) and carbon nanotubes (CNTs). By comparing the selectivity of the two catalysts, it was found that when an NGA was used as the support, the product tended to generate methane due to enhanced H2 adsorption on the material surface, resulting in an increased H/C ratio. However, CNTs without nitrogen doping showed higher selectivity towards C5+ hydrocarbons. Although the C5+ selectivity of the NGA was not as high as that of the CNTs, the high dispersion and reduction rate of the active components in the N-doped materials significantly improved their catalytic performance compared to the CNTs. Wang et al. [76] prepared three-dimensional nitrogen-doped graphene aerogel catalysts with different Co loadings using a hydrothermal method, and investigated the effects of the Co content and Co phase on the catalytic performances. The results showed that when Co participated in the gelation of the aerogel, an excessive Co loading resulted in encapsulation within the material, reducing the reaction activity. Moreover, the phase change in active Co species was found to affect the activity of the catalyst. In more detail, as the reaction progressed, the Co phase in the in situ synthesized catalyst transformed from the fcc phase to the hcp phase, resulting in an increase in the catalytic activity with the extension of the reaction time (as shown in Figure 1). By changing the Co loading method (in situ/post-treatment), the authors found that no phase transition from the fcc to the hcp phase was observed in the post-treated catalyst, indicating no increase in catalytic activity. As mentioned earlier, different treatment conditions can have impacts on the structure of the materials. For example, Zhan et al. [77] used a pyrolysis-oxidation–reduction approach to treat the materials, resulting in the formation of a polyhedron structure consisting of nitrogen-doped nanotubes embedded with hollow CoNPs (as shown in Figure 2). In addition, they also studied the effects of different treatment methods on the catalyst structure and the relevant catalytic performances. The H-Co@NCNHP catalyst exhibited a CO conversion of 16.4% and a C5+ selectivity of 90.4% at 250 °C, 2 MPa, and 28,400 mLgcat−1h−1. The catalyst was not deactivated after 120 h of operation. H-Co NPs that underwent the oxidation–reduction process exhibited a stronger CO/CHx adsorption ability compared to S-Co NPs that did not undergo this process. Additionally, DFT calculations showed that the electronic structure of the H-Co NPs enhanced the CO dissociation adsorption and reduced the energy barrier for C-C coupling. This undoubtedly enhanced the reactivity and the possibility of long-chain hydrocarbon formation.
In addition to conventional carbon nanotubes as supports, carbon nanosheets and carbon nanospheres are also commonly used for N-modification. Taghavi et al. [78] investigated the effect of N-functionalization on FTS by using pure carbon nanosheets and nitrogen-doped carbon nanosheets as supports and loading Co on their surfaces. The characterization results showed that N doping increased the surface defects of the material, providing more anchor points for a better dispersion of Co and smaller active phase particles. Chernyak et al. [79] conducted experiments using pristine/oxidized/N-doped carbon nanosheets as supports and loading Co. By treating the carbon nanosheets using different pathways, the author found that Co/GNFox (oxidized nanosheets) exhibited the highest anti-sintering capability. N doping and oxidation treatment significantly reduced the particle size of the active component. Meanwhile, they reduced the catalyst’s specific surface area and pore size, which hindered the transfer and diffusion of products and impeded the formation of C5+. Although the C5+ selectivity decreased, the CO conversion was greatly improved. Dlamini et al. [80] utilized N-doped hollow carbon spheres as supports, which promoted both the dispersion and prevention of Co sintering. Cheng et al. [81] constructed N sites on carbon nanospheres and adjusted the N content by changing the pyrolysis temperature. Anchoring the N enhanced the dispersion of Co, and increased the exposure of active sites on the catalyst surface. The electron-rich Co facilitated the adsorption and dissociation of CO, which led to the generation of more C* on the catalyst surface. Then, CO-TPD showed that the catalyst with the highest pyrrolic N content had the highest CO2 desorption temperature (as shown in Figure 3), indicating stronger Co-C bonding, which increased the abundance of C* species and was beneficial for C-C coupling. The author considered the N-anchoring ability in the following order: pyrrolic N > pyridinic N > graphitic N.
To address the issue of the strong interaction between the traditional support, Al2O3, and the metal, Co, Guo et al. [82] coated the surface of Al2O3 with a layer of N-doped carbon shell. The nitrogen-doped carbon (NC) layer on the outer surface of the Al2O3 weakened the interaction between the metal and the support, making it more difficult for the Co to enter the lattice of the Al2O3 and facilitating its reduction. Additionally, the incorporation of N enhanced the reducibility of the catalyst. Particularly, the pyrrolic N, which increased the electron density on the surface of Co, led to strengthened C-Co bonds and weakened C=O bonds.
The summary of previous experimental research for cobalt-based catalysts is presented in Table 1.

3.2. Iron-Based Catalysts

Fe-based catalysts have been applied for CO hydrogenation due to their low cost and excellent FTS reaction activity. Carbon materials have been found to facilitate the carburization of Fe, promoting the formation of more active iron carbide phases [88]. Introducing nitrogen atoms into carbon materials can achieve similar effects to that of K promoters, namely increasing the surface basicity of the material and modifying the electron density of the active components. Therefore, the catalytic performance of the Fischer–Tropsch synthesis (FTS) reaction is further enhanced [89].
Tian et al. [90] utilized ammonium citrate ferrite to simultaneously load Fe and dope N into activated carbon. Raman spectroscopy revealed that the graphitization degree of the activated carbon remained unchanged before and after N doping, indicating that N entered the Fe lattice rather than the carbon material itself, thereby enhancing the electron transfer between the N and the Fe. To investigate the influences of different active phases encapsulated in a carbon shell on catalytic performances, Lee et al. [91] prepared three different catalysts (i.e., Fe@C/NPC, Fe3C@C/NPC, Fe5C2@C/NPC), as shown in Figure 4. The results showed that Fe5C2@C/NPC has high loading and dispersion, achieving a CO conversion of 96.4%, an FTS activity of 4.3 × 10−4 molCOgFe−1s−1, and a hydrocarbon yield of 4.71 gHCgcat−1h−1. The encapsulation of the carbon shell reduced the catalyst sintering and significantly enhanced the stability. N atoms adjusted the interface interaction, and increased the catalyst’s surface alkalinity. Then, the CO adsorption and dissociation were enhanced. The advantages of carbon materials were also demonstrated in Tang’s study [92] on inhibiting the migration of Fe species and maintaining their high dispersion during the reaction. Furthermore, the increase in the graphitization degree of the carbon material enhanced the electron transfer from N atoms to Fe. It was also shown that N species strengthened the Fe-C bond and weakened the C-O bond. The stable N-doped carbon shell structure enabled the catalyst to achieve a gasoline selectivity of nearly 50%. This is mainly because the thickness of the carbon shell hindered the outward diffusion of olefins, increasing the possibility of further hydrogenation and enhancing the selectivity towards long-chain hydrocarbons. In addition, N can act as an anchoring site for active metals, facilitating dispersion. Liu et al. [89] used an impregnation method to load 40 wt% Fe onto nitrogen-rich mesoporous carbon as an efficient catalyst for FTS. The results showed that the high pore volume of mesoporous carbon and the anchoring effect of N achieved the dispersion of high loading Fe (as shown in Figure 5).
MOFs have abundant porosity and an adjustable functionality, which are widely used as precursors for nitrogen doping [93]. ZIF-8 is a typical example, where, upon thermal decomposition at temperatures above 900 °C, all the Zn can be evaporated to generate pure N-doped carbon material [94]. Furthermore, ZIF-8 can be divided into two categories, cubic and nanosheet forms, providing more choices for catalyst preparation [95]. Zhao et al. [55] prepared N-doped nanosheets through the thermal decomposition of ZIF-8 nanosheets and used them as catalyst supports (as shown in Figure 6). Compared with traditional activated carbon and cubic ZIF-8, the advantage of nanosheets lies in their higher degree of graphitization, which is beneficial for electron transfer. In more detail, the CO conversion significantly increased from 7.7% (AC) and 22.1% (cubic) to 45.9%. XPS characterization analysis indicated that pyrrole N has a stronger electron-donating capacity than pyridine N and graphite N. TEM images (as shown in Figure 7) of the catalyst after use revealed that N anchoring reduced the sintering of the active phases and slowed down the deactivation, highlighting the indispensable role of N doping in catalyst stability. Furthermore, according to the work of Zhao et al. [96], it was found that the thermal decomposition temperature significantly affected the total N content and the proportion of N species in the materials. Increasing the thermal decomposition temperature within a reasonable range can improve the proportion of pyrrolic N, providing a stronger electron-donating ability to Fe.
The microwave-assisted preparation of N-doped carbon materials has attracted extensive attention from researchers due to its extremely short preparation time. Guo et al. [29] used microwave treatment instead of traditional calcination, greatly reducing the preparation time of NC materials (within seconds). The authors found that by adjusting the microwave treatment time, the particle size of the catalytic active component could be changed. In more detail, increasing the treatment time resulted in a decrease in the particle size from 120 nm to 10 nm. The ammonia treatment time can modify the total amount and configuration of N. With a 12 h ammonia treatment and a 10 s microwave treatment, the catalyst exhibited a CO conversion of 97.2% and a C5+ selectivity of 40% at 320 °C and 2 MPa. Park et al. [97] prepared a catalyst consisting of carbon nanotubes, N-doped reduced graphene oxide, and nano-sized Fe particles through microwave-assisted and CO activation. Under a space velocity of 210 NL gcat−1h−1, the STY value reached 4.4 × 10−3 molCOgFe−1s−1, and the CO conversion reached 80.5%. The high activity is attributed to the unique structure of the catalyst, which possesses a high surface area and a three-dimensional pore system that enhances mass transfer.
The uncontrollable nature of N species in N-doped carbon materials makes it difficult for researchers to elucidate the specific role of individual N species. In response to the complexity of N species in current N-doped carbon materials, Liu et al. [98] prepared carbon nanotubes containing only pyrrolic N through a simple hydrothermal method (as shown in Figure 8). It was found that the content of pyrrolic N was positively correlated with C5+ selectivity, which was mainly because pyrrolic N enhanced the alkalinity of the catalyst surface and the electron density around the active components. More specifically, the addition of N can enhance the interaction between the active metal and the support, and it also increases the reduction temperature of Fe. The formation of iron carbide is closely related to the degree of reduction, and a high degree of reduction favors the formation of iron carbide. Therefore, N doping increases the selectivity of C5+ hydrocarbons. A similar conclusion can be proven in the work of Lama et al. [99]. They modified CNTs with N atoms, and found that N doping stabilized smaller Fe particles (nanoscale) and could even stabilize atomic-level Fe species. The formation of iron carbide was accelerated (rapid activation capability), which could be attributed to the formation of an electron junction between the Fe species and the NC layer, enhancing the binding between CO and Fe.
In summary, although the active components (Fe/Co) may differ, N-doped carbon materials share many similarities in the mechanism of the CO hydrogenation reaction. N species have a significant impact on the surface acidity/basicity and the electronic structure around the active components of the catalyst. Furthermore, N-doped carbon materials can provide abundant anchoring sites, reducing the sintering of active components during the reaction process.
The summary of previous experimental research for iron-based catalysts is presented in Table 2.

4. Catalytic Application of Nitrogen-Doped Carbon Materials in CO2 Hydrogenation

In order to achieve global carbon neutrality, the efficient utilization of waste CO2 has been a concern in both industry and academia. Generally, CO2 can be treated as an important carbon source, and can be converted into valuable long-chain hydrocarbons with the aid of green H2 [107,108]. The obtained long-chain hydrocarbons can be used as an alternative to traditional fossil fuels owing to their high volumetric energy density. The products from CO2 hydrogenation exist in the form of C-H and C-C chemical bonds, known for their high volumetric energy density. These products are not only core components in the industry but also the foundation of modern society. Consequently, research in this field has garnered significant attention in recent years. Below, we will analyze and summarize the different performances of CO2 hydrogenation catalysts, as well as reasons for using them, from the perspective of different products.

4.1. CO2 Hydrogenation to CO over Nitrogen-Doped Carbon Materials

The reverse water–gas shift (RWGS) reaction involves the conversion of CO2 and H2 into CO and H2O [109]. The produced CO can be used as a feedstock for FTS to synthesize liquid fuels and high-value-added chemicals. Li et al. [110] loaded different metals (Fe, Co, Ni, and Cu) onto N-doped carbon materials derived from ZIF-8, and found that the reactivity order was Ni = Fe > Co > Cu (as shown in Figure 9). The experiment result shows that pyridinic N and metal carbides can capture and enrich CO2, and the number of pyridinic N sites correlates positively with the alkaline sites on the catalyst surface. Moreover, DFT calculations indicate that CO* tends to desorb rather than further hydrogenate, resulting in a high CO selectivity (as shown in Figure 10). Bruno et al. [111] covered SiO2 with a layer of N-doped carbon, which led to a higher CO selectivity compared with catalysts without the N-doped carbon layer. This can be attributed to a weakened CO adsorption on the Ni-NC surface. However, catalysts without the N-doped carbon layer exhibited better CH4 selectivity, indicating that the presence of the NC carbon layer can significantly alter product selectivity. In other words, the multilinear binding of CO with Ni on the surface of pure Ni catalysts is stronger than the linear binding of CO with Ni on the NC-Ni surface, making pure Ni more prone to promoting the formation of methane, while NC-Ni is more prone to resulting in CO generation. Li et al. [112] prepared catalysts with single atomic Co-N4 and Co nanoparticle structures. It was found that the atomic-level Co had a weaker CO adsorption, resulting in a preference for CO as the product. The nanoparticles had a stronger H2 activation ability than the single atom, leading to a higher tendency for CH4 production (selectivity of 99%). Considering that the anchoring effect of N promotes the formation of atomic-level Co (as shown in Figure 11), the quantity of N atoms plays a crucial role in the formation of single atoms. An insufficient N content leads to the formation of nanoparticles. Additionally, the electron structure of Co is adjusted to form Con+ species, optimizing the adsorption and desorption of intermediate products. Besides the commonly used Ni- and Fe-based catalysts, Mo-based catalysts have also been developed for the RWGS reaction. Zhang et al. [113] used N-doped graphene to support various metal single atoms and performed systematic DFT calculations, revealing that the VIB group metals exhibited significant advantages in the RGWS reaction. Particularly, Mo can provide electrons to CO2 molecules, facilitating the spontaneous generation of CO and O*, and decreasing the hydrogenation barrier of O* to 1.10 eV. Jiang et al. [114] used Mo as the active component in the RWGS reaction. The prepared catalyst had a MoN3 structure, which could effectively adsorb CO2 and convert it to CO. Under a low H2 partial pressure, the catalyst showed a conversion of around 30% at 500 °C and a low H2 partial pressure, with close to a 100% CO selectivity. DFT calculations indicated that the transition state energy during CO generation was at most 0.82 eV, much lower than the 2.16 eV required for CH4 formation, making it more favorable for CO selectivity. Compared to Mo/C, Mo/NC has fewer stable CO adsorption sites, making CO easier to desorb. The performance data of catalysts in the published literature are summarized in Table 3.

4.2. CO2 Hydrogenation to CH4 over Nitrogen-Doped Carbon Materials

Converting CO2 to CH4 through hydrogenation is also known as the Sabatier reaction. Methane is characterized by its easy storage and transportation, and it can be used as a fuel and a precursor for the production of other chemicals [115,116]. Stasi et al. [117] proved the potential of N doping to enhance the performance of catalysts for CO2 hydrogenation to methane. The authors introduced Ce and N separately into Ni-based catalysts and compared their catalytic effects. Both additives increased the catalyst’s surface alkalinity. While the effect of N was not as good as that of Ce, an increase in the N content showed an upward trend in the conversion rate and selectivity. In order to investigate the influence of N doping on the reaction, David et al. [27] directly introduced Fe and N into the carrier through CVD to maximize the interaction between the particles and the carrier. The catalyst doped with N showed an increased CO2 conversion rate and CH4 selectivity, but a significant decrease in CO selectivity. The TPD characterization data showed that N-doped carbon materials exhibited an increased adsorption strength for CO2 and CO, possibly due to the dipole moment similarity between C-N and CO2/CO. However, the absence of a dipole moment in hydrocarbon products made them more prone to desorption, leading to a decrease in the selectivity of long-chain hydrocarbons. To verify the effect of N, Na was introduced as an additive to mask the effect of N. The product distribution exhibited catalytic effects opposite to the N-doped carbon materials, validating the authors’ assumption. To elucidate the advantages of carbon materials over traditional metal oxide supports, Roldan et al. [118] compared Ru catalysts supported on different carriers (CNT, CNF, NCNF, Al2O3). Among all the carbon supports, CNF performed the worst. The carbon materials exhibited stronger CO2 adsorption compared to Al2O3. And with the incorporation of N into carbon materials, the interaction between the CO2 chemical adsorption intermediates and the nitrogen atoms was enhanced.
Wu et al. [119] prepared non-metal-loaded nitrogen-doped carbon materials for CO2 hydrogenation and compared the reaction results before and after N doping. The results showed that the introduction of N increased the surface Lewis basic sites and enhanced the CO2 adsorption. The catalyst activity (an increased CO2 conversion rate, increased CH4 selectivity, and a lowered reaction starting temperature) and the number of pyridinic N sites showed a significant positive linear relationship, while pyrrolic N and graphitic N did not exhibit this relationship. Wang et al. [120] also demonstrated that the addition of N would result in CO2 concentration gradients around the active components. They prepared a Ni-based catalyst supported on N-doped nanotubes, which efficiently converted CO2 selectively into methane. Under a high gas hourly space velocity of 120,000 mLg−1 h−1, the STY of methanol in the product was 5 molCH4gNi−1h−1. Moreover, the catalyst exhibited a significantly enhanced resistance to coke deposition. Wang et al. [65] used urea as the N source, and red pine as the C source to prepare materials through in situ thermolysis, with the metal Ru loaded as the methanation catalyst. The catalyst with the highest pyridinic N content achieved a CO conversion rate of 92% and a CH4 selectivity of 99.7% under the reaction conditions of 380 °C, 1 MPa, and H2/CO2 = 4. An XPS analysis revealed a higher content of zero-valent metal in the N-doped material, indicating that N altered the electronic state of the active metal, thereby affecting its reducibility. The increased zero-valent Ru improved the H2 dissociation, leading to an increased CH4 selectivity. Anchoring sites on the carrier reduced the aggregation and leaching of active sites during the reaction. In contrast to other researchers, the total N content in the material increased with the increasing pyrolysis temperature, possibly due to the favorable combination of N atoms and carbon materials at high temperatures. Gholampour et al. [56] investigated N-doped carbon materials derived from two different structures (nanosheets/cube) of ZIF-67 for methanation. It was found that ZIF-67-L (nanosheets) had smaller Co particles and a higher number of N species (especially pyridinic N, which was believed to exhibit the strongest Lewis basicity and have a unique effect on CO2 adsorption). The catalyst showed good performance in a 60 h stability test. Liliana et al. [121] compared the catalytic performance of Ni/CNT and Ni/CNT-N. The N-doped CNT exhibited a stronger interaction with the metal, Ni, resulting in an atomic-level dispersion of Ni in the catalyst. The specific metal surface area of CNT-N was 3.5 times higher than that of CNT, indicating better dispersion of the active metal, Ni. Gödde et al. [122] achieved a CO2 conversion rate of 51.4% and a CH4 selectivity of 95.8% (as shown in Table 4) by adjusting the loaded Ni content on NCNT. Despite the high loading of Ni, the anchoring effect of N kept the overall Ni particle size below 10 nm, providing a large accessible Ni metal surface area, even at high loading.

4.3. CO2 Hydrogenation to C2+ over Nitrogen-Doped Carbon Materials

With the in-depth study of CO2 hydrogenation, single-carbon atom compounds (e.g., CO and CH4) have been widely reported. In recent years, C2+ compounds have gained attention from researchers due to their potential conversion into high-value compounds such as olefins and aromatics [123]. Liu et al. [124] used nitrogen-doping materials derived from the pyrolysis of ZIF-8 as the carrier of the Fe-based catalyst, and the experimental results showed that the carbon carrier containing N species exhibited a better C2+ olefin selectivity and C5+ selectivity compared to traditional carbon carriers. Dong et al. [57] prepared a Fe-Co bimetallic N-doped catalyst using different pyrolysis temperatures. With the increase in the pyrolysis temperature, the total N content decreased, and the proportion of pyridinic N decreased (as shown in Figure 12). Generally, the pyrolysis temperature also has a significant influence on the carbonization degree of N-doped carbon carriers, the particle size, and the carburization degree of active components. Zhang et al. [125] addressed the issue of the poor activity of catalysts for CO2 hydrogenation to olefins by confining Fe3O4-FeCx heterojunction active sites within N-doped graphene shells (as shown in Figure 13). Under reaction conditions of 3.0 MPa, 320 °C, and GHSV = 4800 mLgcat−1h−1, the catalyst achieved a CO2 conversion of 54.5% (as shown in Table 5) and a C2=-C4= selectivity of 65.63%. The role of N-MOC is to enrich alkaline sites and enhance CO2 adsorption while inhibiting H2 adsorption and dissociation. FeNx formed after the interaction between the carrier and Fe can be converted into active sites of FeCx, enhancing the activity of the FTS reaction. Moreover, the FeCx particles in the N-doped catalyst were smaller, which is beneficial for structure-sensitive FTS reactions. At the same time, the shell formed by N-MOC prevented the contact between the active sites and water, avoiding the sintering of the heterojunction, and increasing the stability of the catalyst (with a good stability over 100 h of operation). Peng et al. [126] prepared a catalyst with a nitrogen-doping layer coated on a Co-Fe alloy, which achieved over a 50% C2+ selectivity and more than a 30% selectivity of ethylene and propylene under reaction conditions of 300 °C and 5 bar. Theoretical calculations showed that the NC carrier significantly altered the electronic density of the Co-Fe alloy, favoring the formation of C-C bonds.

4.4. CO2 Hydrogenation to CH3OH over Nitrogen-Doped Carbon Materials

Methanol, as one of the potential products of CO2 hydrogenation, is widely used for solvents, fuels, and platform chemicals. Limleamthong et al. [127] used nitrogen-doped porous carbon as a support, as well as loading Cu-Zn/Ce. The catalyst showed a space–time yield of 512 mgMeOH gcat−1 h−1 at a reaction temperature of 250 °C and pressure of 1.5 MPa. Sun et al. [128] synthesized catalysts dominated by different N species through different treatment methods and discussed the role of different N species in CO2 hydrogenation to methanol. The results showed that pyridinic N could lower the reduction temperature of CuO, and inhibit the growth of Cu particles. Moreover, pyridinic N could promote the dispersion of Cu and enhance the dissociation adsorption of H2. Similar conclusions were obtained in the study of Deerattrakul et al. [129]. Ma et al. [130] prepared CuZnAl catalysts with different contents of N-doped graphene via co-precipitation for methanol synthesis. The results indicated that there was an optimal value for the content of N-doped graphene (10 wt%). Initially, as the nitrogen-doping carbon material’s content increased, it led to the generation of oxygen vacancies, promoting the reduction of Cu species. As the NC content continued to increase, the H2 adsorption was enhanced and a reduction was facilitated. However, with the further increase in the nitrogen-doping carbon materials, the Cu sites could be covered, making the Cu reduction more difficult. Donphai et al. [131] used PBZ-derived C materials and doped them with different N sources (i.e., ammonia, urea, and arginine). Among the three N sources, arginine led to a decrease in the pore structure and specific surface area of the catalyst, which was not favorable for the loading of active components. Additionally, the low N content resulted in a decrease in CO2 adsorption. The other two N sources also slightly altered the physical structure. Thermogravimetric analysis showed an enhanced thermal stability in the N-doped materials. However, the three different N sources did not have a significant impact on the reaction results. The performance data of catalysts in the published literature are summarized in Table 6.

4.5. CO2 Hydrogenation to CHOOH over Nitrogen-Doped Carbon Materials

Formic acid is an important chemical which possesses a good hydrogen storage capacity (by weight and volume). Therefore, it is considered to be an ideal hydrogen storage material [132], and a feasible CO2 conversion and utilization solution. Yang et al. [133] prepared a nitrogen-doped, hollow, carbon-supported PdAg alloy catalyst for formic acid production. The changes in the thermal decomposition temperature affected the material’s specific surface area (which decreased as the temperature increased and the carbon shell became thinner), the dispersion of active components (an increased particle size with higher temperature), N content (which decreased with higher temperature), and the formation of Pd-N bonds. The introduction of N increased CO2 adsorption and H2 dissociation, while reducing the reaction activation energy. The decrease in carbon material defects upon the addition of the active phase indicated their combination, highlighting the important role of the N anchoring effect. Ahn et al. [134] loaded Ru onto N-doped mesoporous hollow carbon spheres. Through cyclic tests, it was found that Ru would leach out, but the amount of Ru leaching was reduced after N doping, benefiting from the strong binding energy between the Ru species and the pyridinic N and pyrrolic N sites, thereby enhancing the catalyst’s stability. Catalysts with a higher content of graphitic N exhibited serious leaching in cyclic tests, suggesting that the interactions between pyridinic N, pyrrolic N, and Ru were stronger than those with graphitic N. DFT calculations showed that the adsorption energy of Ru and pyridinic N was lower than that of graphitic N, indicating greater stability of Ru-pyridinic N. However, experimental results showed an improved catalytic performance with an increasing proportion of graphitic N species, implying that Ru-graphitic N might be the active site for the reaction. Jaleel et al. [135] found through calculations that among the various modes of N and Ru interaction, the bipyridine N site had the most stable binding with Ru (as shown in Figure 14), which could reduce the loss of Ru3+ during the reaction. Additionally, nitrogen doping increased the electron density of Ru. Zhang et al. [136] incorporated main group metals (B, Al, Ga) into N-doped materials and designed catalysts with a frustrated Lewis acid/base pair structure. Simulation calculations revealed that all three catalysts were favorable for the activation and dissociation of H2. Particularly, the catalysts designed with Al and Ga displayed a lower activation energy for the rate-determining step, indicating their potential as promising catalysts.
Compared to CO hydrogenation, the study of N-doped carbon materials in CO2 hydrogenation started later. The surface alkalinity of nitrogen-doped carbon materials is stronger compared to traditional supports, making it easier for CO2 (Lewis acid) to adsorb. Moreover, the anchoring effect of N species themselves allows researchers to load a larger amount of active components on the carbon support, with high dispersion. N-doped carbon materials have unique advantages compared to traditional supports. The performance data of catalysts in the published literature are summarized in Table 7.

5. Concluding Remarks

In the face of the dual crisis of energy and environment, researchers have conducted extensive research on COx hydrogenation conversion and utilization. The advantage of carbon materials over traditional carriers lies in the weaker metal–carrier interaction. Furthermore, the confinement effect of carbon materials allows for the easy separation of active components, effectively preventing their aggregation and leaching during the reaction process. The role of N doping in carbon materials is as follows: nitrogen serves as an anchoring site for the better dispersion of active components, leading to a higher utilization of metals. The anchoring effect can also prevent the sintering of active components during catalytic reactions. The electron-donating effect of N atoms increases the surface alkalinity of the catalyst, facilitating the adsorption of CO2. It can be observed that N-doped carbon materials not only increase the adsorption of the reactants (CO2) but also promote their dissociation. Compared to traditional metal oxide carriers, nitrogen-doped carbon materials have unique advantages.
According to current research findings, there are several challenges in the application of nitrogen-doped carbon materials in COx hydrogenation reactions. Firstly, the main challenge in this field is the high cost of nitrogen-doped carbon material synthesis, which hinders its large-scale application in COx hydrogenation. Developing simple and cost-effective synthesis methods is crucial. Secondly, the diversity of nitrogen species in nitrogen-doped carbon materials hinders the understanding of the specific roles of certain nitrogen species. Thirdly, carbon materials have poor stability in oxidative atmospheres, posing difficulties in catalyst preparation and regeneration. Fourth, among the numerous synthesis methods, there is a question of how to select the most favorable method for the reaction. It is particularly important to compare the effects of different doping methods on the reaction while controlling the physical and chemical properties of the materials to be similar. Lastly, nitrogen atoms on the surface of nitrogen-doped carbon materials may be lost under harsh reaction conditions, leading to the sintering of the active components.
Currently, many investigations have been performed in this regard, but many issues of COx hydrogenation to value-added products have not been solved yet.
  • Methods to improve the stability of nitrogen atoms on carbon materials and increase the nitrogen content need to be explored in relation to the material itself. On the one hand, nitrogen atoms serve as effective anchoring sites for active components, but the quantity of nitrogen atoms tends to decrease with the increasing pyrolysis temperature of the material. On the other hand, high pyrolysis temperatures often enhance the graphitization of carbon materials, facilitating electron transfer. These two processes seem contradictory and it appears difficult to achieve both simultaneously. If the interaction between nitrogen atoms and carbon materials can be strengthened (adding functional groups to the surface of the material), allowing nitrogen atoms to remain well-preserved on the surface during high-temperature pyrolysis, it would contribute to further enhancing the catalytic activity of the catalyst. Currently, there are limitations in our understanding of the role of nitrogen species. Nitrogen-doped carbon materials prepared using the current doping methods have complex nitrogen species, and it is difficult to precisely control the content of the corresponding components. This greatly restricts our understanding of the specific roles of nitrogen species (such as which of pyridine N and pyrrole N has the stronger electron-giving ability and the stronger anchoring ability), and causes significant confusion in explorations of the synthesis and mechanisms of subsequent catalytic materials. It is challenging to introduce a specific nitrogen species onto the surface of traditional carbon materials. Therefore, it is crucial to modify the surface of carbon materials or synthesize them through the pyrolysis of raw materials containing specific nitrogen species. Carbon materials with specific nitrogen species can help us to understand the active sites of reactions and provide important support for speculating reaction pathways.
  • With the rapid development of porous carbon technology, nitrogen-doped carbon materials with a tunable pore size and hierarchical pore structure can be developed. N-doped carbon materials with adjustable microporous and mesoporous structures can be prepared, featuring a layered structure that connects mesopores and macropores. This facilitates the investigation of the pore structure’s effect on the reaction. Micro-porous structures can spatially restrict the distribution of active components and promote the formation of nanoscale particles. At the same time, the layered structure is beneficial for improving the mass and heat transfer of the catalyst, which has a positive effect on enhancing the reaction activity and improving the product distribution.
  • A comprehensive insight into active species and their interactions, along with the interactions among the active centers with support, are of paramount significance for catalyst design and tailoring. Therefore, in situ characterization techniques (such as XRD, TEM, and IR spectroscopy) should be employed to distinguish active sites. Additionally, analysis methods such as XPS and DFT calculations are needed to identify the major surface species and explain the elusive reaction mechanisms of COx hydrogenation to value-added products. Moreover, in situ structural and surface studies can be conducted under working conditions to validate the proposed mechanisms.
The purpose of this review is to summarize the recent research on the application of N-doped carbon materials in COx hydrogenation. It aims to provide insights for the future development of N-doped carbon materials.

Author Contributions

Conceptualization, C.D.; methodology, R.G.; validation, L.Z. and L.W.; formal analysis, R.G.; investigation, C.D.; resources, C.Z.; writing—original draft preparation, R.G.; writing—review and editing, C.Z.; visualization, K.H., L.X. and X.C.; supervision, C.Z.; project administration, C.Z.; funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Next Generation Carbon Upcycling Project” (Project No. 2017M1A2A2043133) through the National Research Foundation (NRF) funded by the Ministry of Science and ICT, Republic of Korea. We also appreciate the Natural Science Foundation of Jiangsu Province (BK20200694, 20KJB530002, 21KJB480014), the Jiangsu Specially-Appointed Professors Program, and the open program of the State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (2021-K32).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are presented in the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The authors declare that they have no known competing financial interest or personal relationship that could have appeared to influence the work reported in this paper.

Abbreviations

FTSFischer–Tropsch synthesis
RWGSReverse water–gas shift
ASFAnderson–Schulz–Flory
CNTCarbon nanotube
NCNTNitrogen-doped carbon nanotube
CNFCarbon nanofiber
NCNFNitrogen-doped carbon nanofiber
ZIFsZeolitic imidazolate framework materials
PANIPolyaniline
PPyPolypyrrole
STYSpace–time yield
CVDChemical vapor deposition
DFTDensity functional theory

References

  1. Martínez, L.; Merino, P.; Santoro, G.; Martínez, J.I.; Katsanoulis, S.; Ault, J.; Mayoral, Á.; Vázquez, L.; Accolla, M.; Dazzi, A.; et al. Metal-catalyst-free gas-phase synthesis of long-chain hydrocarbons. Nat. Commun. 2021, 12, 5937. [Google Scholar] [CrossRef] [PubMed]
  2. Bao, J.; Yang, G.; Yoneyama, Y.; Tsubaki, N. Significant Advances in C1 Catalysis: Highly Efficient Catalysts and Catalytic Reactions. ACS Catal. 2019, 9, 3026–3053. [Google Scholar] [CrossRef]
  3. Cui, W.-G.; Zhang, G.-Y.; Hu, T.-L.; Bu, X.-H. Metal-organic framework-based heterogeneous catalysts for the conversion of C1 chemistry: CO, CO2 and CH4. Coord. Chem. Rev. 2019, 387, 79–120. [Google Scholar] [CrossRef]
  4. Claeys, M. Cobalt gets in shape. Nature 2016, 538, 44–45. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, Y.; Wei, J.; Duyar, M.S.; Ordomsky, V.V.; Khodakov, A.Y.; Liu, J. Carbon-based catalysts for Fischer–Tropsch synthesis. Chem. Soc. Rev. 2021, 50, 2337–2366. [Google Scholar] [CrossRef] [PubMed]
  6. Mainali, K.; Mood, S.H.; Pelaez-Samaniego, M.R.; Sierra-Jimenez, V.; Garcia-Perez, M. Production and applications of N-doped carbons from bioresources: A review. Catal. Today 2023, 423, 114248. [Google Scholar] [CrossRef]
  7. Tackett, B.M.; Gomez, E.; Chen, J.G. Net reduction of CO2 via its thermocatalytic and electrocatalytic transformation reactions in standard and hybrid processes. Nat. Catal. 2019, 2, 381–386. [Google Scholar] [CrossRef]
  8. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef]
  9. Burkart, M.D.; Hazari, N.; Tway, C.L.; Zeitler, E.L. Opportunities and Challenges for Catalysis in Carbon Dioxide Utilization. ACS Catal. 2019, 9, 7937–7956. [Google Scholar] [CrossRef]
  10. Orege, J.I.; Wei, J.; Han, Y.; Yang, M.; Sun, X.; Zhang, J.; Amoo, C.C.; Ge, Q.; Sun, J. Highly stable Sr and Na co-decorated Fe catalyst for high-valued olefin synthesis from CO2 hydrogenation. Appl. Catal. B Environ. 2022, 316, 121640. [Google Scholar] [CrossRef]
  11. Zheng, T.; Zhang, M.; Wu, L.; Guo, S.; Liu, X.; Zhao, J.; Xue, W.; Li, J.; Liu, C.; Li, X.; et al. Upcycling CO2 into energy-rich long-chain compounds via electrochemical and metabolic engineering. Nat. Catal. 2022, 5, 388–396. [Google Scholar] [CrossRef]
  12. Wang, X.; Zhang, J.; Chen, J.; Ma, Q.; Fan, S.; Zhao, T.S. Effect of preparation methods on the structure and catalytic performance of Fe–Zn/K catalysts for CO2 hydrogenation to light olefins. Chin. J. Chem. Eng. 2018, 26, 761–767. [Google Scholar] [CrossRef]
  13. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling Carbon Dioxide through Catalytic Hydrogenation: Recent Key Developments and Perspectives. ACS Catal. 2020, 10, 11318–11345. [Google Scholar] [CrossRef]
  14. van Ravenhorst, I.K.; Vogt, C.; Oosterbeek, H.; Bossers, K.W.; Moya-Cancino, J.G.; van Bavel, A.P.; van der Eerden, M.J.; Vine, D.; de Groot, F.M.F.; Meirer, F.; et al. Capturing the Genesis of an Active Fischer–Tropsch Synthesis Catalyst with Operando X-ray Nanospectroscopy. Angew. Chem. Int. Ed. 2018, 57, 11957–11962. [Google Scholar] [CrossRef]
  15. Meng, G.; Sun, J.; Tao, L.; Ji, K.; Wang, P.; Wang, Y.; Sun, X.; Cui, T.; Du, S.; Chen, J. Ru1Con Single-Atom Alloy for Enhancing Fischer–Tropsch Synthesis. ACS Catal. 2021, 11, 1886–1896. [Google Scholar] [CrossRef]
  16. Hernández Mejía, C.; van der Hoeven, J.E.S.; de Jongh, P.E.; de Jong, K.P. Cobalt–Nickel Nanoparticles Supported on Reducible Oxides as Fischer–Tropsch Catalysts. ACS Catal. 2020, 10, 7343–7354. [Google Scholar] [CrossRef]
  17. Okonye, L.U.; Yao, Y.; Hildebrandt, D.; Meijboom, R. Contributing to energy sustainability: A review of mesoporous material supported catalysts for Fischer–Tropsch synthesis. Sustain. Energy Fuels 2021, 5, 79–107. [Google Scholar] [CrossRef]
  18. Galvis, H.M.T.; de Jong, K.P. Catalysts for Production of Lower Olefins from Synthesis Gas: A Review. ACS Catal. 2013, 3, 2130–2149. [Google Scholar] [CrossRef]
  19. Zitolo, A.; Goellner, V.; Armel, V.; Sougrati, M.-T.; Mineva, T.; Stievano, L.; Fonda, E.; Jaouen, F. Identification of catalytic sites for oxygen reduction in iron- and nitrogen-doped graphene materials. Nat. Mater. 2015, 14, 937–942. [Google Scholar] [CrossRef]
  20. Zhu, C.; Zhang, M.; Huang, C.; Zhong, L.; Fang, K. Carbon-encapsulated highly dispersed FeMn nanoparticles for Fischer–Tropsch synthesis to light olefins. New J. Chem. 2018, 42, 2413–2421. [Google Scholar] [CrossRef]
  21. Galvis, H.M.T.; Bitter, J.H.; Khare, C.B.; Ruitenbeek, M.; Dugulan, A.I.; de Jong, K.P. Supported Iron Nanoparticles as Catalysts for Sustainable Production of Lower Olefins. Science 2012, 335, 835–838. [Google Scholar] [CrossRef] [PubMed]
  22. He, L.; Weniger, F.; Neumann, H.; Beller, M. Synthesis, Characterization, and Application of Metal Nanoparti cles Supported on Nitrogen-Doped Carbon: Catalysis beyond Electrochemistry. Angew. Chem. Int. Ed. 2016, 55, 12582–12594. [Google Scholar] [CrossRef] [PubMed]
  23. Rangraz, Y.; Heravi, M.M.; Elhampour, A. Recent Advances on Heteroatom-Doped Porous Carbon/Metal Materials: Fascinating Heterogeneous Catalysts for Organic Transformations. Chem. Rec. 2021, 21, 1985–2073. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, X.; Bai, Y.; Liu, M.; Li, Y.; Yang, H.; Wang, X.; Wu, C. Untangling the respective effects of heteroatom-doped carbon materials in batteries, supercapacitors and the ORR to design high performance materials. Energy Environ. Sci. 2021, 14, 2036–2089. [Google Scholar] [CrossRef]
  25. Wang, H.; Maiyalagan, T.; Wang, X. Review on Recent Progress in Nitrogen-Doped Graphene: Synthesis, Characterization, and Its Potential Applications. ACS Catal. 2012, 2, 781–794. [Google Scholar] [CrossRef]
  26. Wood, K.N.; O’Hayre, R.; Pylypenko, S. Recent progress on nitrogen/carbon structures designed for use in energy and sustainability applications. Energy Environ. Sci. 2014, 7, 1212–1249. [Google Scholar] [CrossRef]
  27. Williamson, D.L.; Herdes, C.; Torrente-Murciano, L.; Jones, M.D.; Mattia, D. N-Doped Fe@CNT for Combined RWGS/FT CO2 Hydrogenation. ACS Sustain. Chem. Eng. 2019, 7, 7395–7402. [Google Scholar] [CrossRef]
  28. Guo, L.; Zhang, P.; Cui, Y.; Liu, G.; Wu, J.; Yang, G.; Yoneyama, Y.; Tsubaki, N. One-Pot Hydrothermal Synthesis of Nitrogen Functionalized Carbonaceous Material Catalysts with Embedded Iron Nanoparticles for CO2 Hydrogenation. ACS Sustain. Chem. Eng. 2019, 7, 8331–8339. [Google Scholar] [CrossRef]
  29. Guo, L.; Guo, Z.; Liang, J.; Yong, X.; Sun, S.; Zhang, W.; Sun, J.; Zhao, T.; Li, J.; Cui, Y.; et al. Quick microwave assembling nitrogen-regulated graphene supported iron nanoparticles for Fischer-Tropsch synthesis. Chem. Eng. J. 2022, 429, 132063. [Google Scholar] [CrossRef]
  30. de Almeida Ribeiro, R.S.; Monteiro Ferreira, L.E.; Rossa, V.; Lima, C.G.; Paixao, M.W.; Varma, R.S.; de Melo Lima, T. Graphitic Carbon Nitride-Based Materials as Catalysts for the Upgrading of Lignocellulosic Biomass-Derived Molecules. ChemSusChem 2020, 13, 3992–4004. [Google Scholar] [CrossRef]
  31. Zhao, M.; Feng, J.; Yang, W.; Song, S.; Zhang, H. Recent Advances in Graphitic Carbon Nitride Supported Single-Atom Catalysts for Energy Conversion. ChemCatChem 2021, 13, 1250–1270. [Google Scholar] [CrossRef]
  32. Ye, R.-P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. CO2 hydrogenation to high-value products via heterogeneous catalysis. Nat. Commun. 2019, 10, 5698. [Google Scholar] [CrossRef] [PubMed]
  33. Martínez-Vargas, D.X.; Sandoval-Rangel, L.; Campuzano-Calderon, O.; Romero-Flores, M.; Lozano, F.J.; Nigam, K.D.P.; Mendoza, A.; Montesinos-Castellanos, A. Recent Advances in Bifunctional Catalysts for the Fischer–Tropsch Process: One-Stage Production of Liquid Hydrocarbons from Syngas. Ind. Eng. Chem. Res. 2019, 58, 15872–15901. [Google Scholar] [CrossRef]
  34. Khodakov, A.Y.; Chu, W.; Fongarland, P. Advances in the Development of Novel Cobalt Fischer–Tropsch Catalysts for Synthesis of Long-Chain Hydrocarbons and Clean Fuels. Chem. Rev. 2007, 107, 1692–1744. [Google Scholar] [CrossRef]
  35. Mota, F.M.; Kim, D.H. From CO2methanation to ambitious long-chain hydrocarbons: Alternative fuels paving the path to sustainability. Chem. Soc. Rev. 2019, 48, 205–259. [Google Scholar] [CrossRef]
  36. Nie, X.; Li, W.; Jiang, X.; Guo, X.; Song, C. Chapter Two—Recent advances in catalytic CO2 hydrogenation to alcohols and hydrocarbons. In Advances in Catalysis; Song, C., Ed.; Academic Press: Cambridge, MA, USA, 2019; Volume 65, pp. 121–233. [Google Scholar]
  37. Zang, G.; Sun, P.; Elgowainy, A.A.; Bafana, A.; Wang, M. Performance and cost analysis of liquid fuel production from H2 and CO2 based on the Fischer-Tropsch process. J. CO2 Util. 2021, 46, 101459. [Google Scholar] [CrossRef]
  38. Xu, H.; Ma, L.; Jin, Z. Nitrogen-doped graphene: Synthesis, characterizations and energy applications. J. Energy Chem. 2018, 27, 146–160. [Google Scholar] [CrossRef]
  39. Ham, K.; Shin, D.; Lee, J. The Role of Lone-Pair Electrons in Pt–N Interactions for the Oxygen Reduction Reaction in Polymer Exchange Membrane Fuel Cells. ChemSusChem 2020, 13, 1751–1758. [Google Scholar] [CrossRef]
  40. Wu, G.; Mack, N.H.; Gao, W.; Ma, S.; Zhong, R.; Han, J.; Baldwin, J.K.; Zelenay, P. Nitrogen-Doped Graphene-Rich Catalysts Derived from Heteroatom Polymers for Oxygen Reduction in Nonaqueous Lithium–O2 Battery Cathodes. ACS Nano 2012, 6, 9764–9776. [Google Scholar] [CrossRef]
  41. Li, R.; Wang, D. Understanding the structure-performance relationship of active sites at atomic scale. Nano Res. 2022, 15, 6888–6923. [Google Scholar] [CrossRef]
  42. Li, O.L.; Prabakar, K.; Kaneko, A.; Park, H.; Ishizaki, T. Exploration of Lewis basicity and oxygen reduction reaction activity in plasma-tailored nitrogen-doped carbon electrocatalysts. Catal. Today 2019, 337, 102–109. [Google Scholar] [CrossRef]
  43. Feng, H.; Ma, J.; Hu, Z. Nitrogen-doped carbon nanotubes functionalized by transition metal atoms: A density functional study. J. Mater. Chem. 2010, 20, 1702–1708. [Google Scholar] [CrossRef]
  44. Wang, X.; Pan, H.; Lin, Q.; Wu, H.; Jia, S.; Shi, Y. One-Step Synthesis of Nitrogen-Doped Hydrophilic Mesoporous Carbons from Chitosan-Based Triconstituent System for Drug Release. Nanoscale Res. Lett. 2019, 14, 259. [Google Scholar] [CrossRef] [PubMed]
  45. Yin, L.-C.; Liang, J.; Zhou, G.-M.; Li, F.; Saito, R.; Cheng, H.-M. Understanding the interactions between lithium polysulfides and N-doped graphene using density functional theory calculations. Nano Energy 2016, 25, 203–210. [Google Scholar] [CrossRef]
  46. Su, P.; Huang, W.; Zhang, J.; Guharoy, U.; Du, Q.; Sun, Q.; Jiang, Q.; Cheng, Y.; Yang, J.; Zhang, X.; et al. Fe atoms anchored on defective nitrogen doped hollow carbon spheres as efficient electrocatalysts for oxygen reduction reaction. Nano Res. 2021, 14, 1069–1077. [Google Scholar] [CrossRef]
  47. Chen, S.; Qi, P.; Chen, J.; Yuan, Y. Platinum nanoparticles supported on N-doped carbon nanotubes for the selective oxidation of glycerol to glyceric acid in a base-free aqueous solution. RSC Adv. 2015, 5, 31566–31574. [Google Scholar] [CrossRef]
  48. Inagaki, M.; Toyoda, M.; Soneda, Y.; Morishita, T. Nitrogen-doped carbon materials. Carbon 2018, 132, 104–140. [Google Scholar] [CrossRef]
  49. Jeon, I.; Noh, H.; Baek, J. Nitrogen-Doped Carbon Nanomaterials: Synthesis, Characteristics and Applications. Chem. Asian J. 2020, 15, 2282–2293. [Google Scholar] [CrossRef]
  50. Zhang, L.-S.; Liang, X.-Q.; Song, W.-G.; Wu, Z.-Y. Identification of the nitrogen species on N-doped graphene layers and Pt/NG composite catalyst for direct methanol fuel cell. Phys. Chem. Chem. Phys. 2010, 12, 12055–12059. [Google Scholar] [CrossRef]
  51. Luo, W.; Wang, B.; Heron, C.G.; Allen, M.J.; Morre, J.; Maier, C.S.; Stickle, W.F.; Ji, X. Pyrolysis of Cellulose under Ammonia Leads to Nitrogen-Doped Nanoporous Carbon Generated through Methane Formation. Nano Lett. 2014, 14, 2225–2229. [Google Scholar] [CrossRef]
  52. Wang, H.; Liu, X.; Xu, G.; Guo, Z.; Zhang, Y. In situ synthesis of Fe-N-C catalysts from cellulose for hydrogenation of nitrobenzene to aniline. Chin. J. Catal. 2019, 40, 1557–1565. [Google Scholar] [CrossRef]
  53. Wang, H.; Chen, C.; Chen, Y.; Wan, H.; Dong, L.; Guan, G. Construction of ultramicropore-enriched N-doped carbons for CO2 capture: Self-decomposition of polyethyleneimine-based precursor to promote pore formation and surface polarity. J. Environ. Chem. Eng. 2021, 9, 105046. [Google Scholar] [CrossRef]
  54. Ali, M.; Riaz, R.; Anjum, A.S.; Sun, K.C.; Li, H.; Ahn, S.; Jeong, S.H.; Ko, M.J. Microwave-assisted ultrafast in-situ growth of N-doped carbon quantum dots on multiwalled carbon nanotubes as an efficient electrocatalyst for photovoltaics. J. Colloid Interface Sci. 2021, 586, 349–361. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, Q.; Huang, S.; Han, X.; Chen, J.; Wang, J.; Rykov, A.; Wang, Y.; Wang, M.; Lv, J.; Ma, X. Highly active and controllable MOF-derived carbon nanosheets supported iron catalysts for Fischer-Tropsch synthesis. Carbon 2021, 173, 364–375. [Google Scholar] [CrossRef]
  56. Gholampour, N.; Zhao, Y.; Devred, F.; Sassoye, C.; Casale, S.; Debecker, D.P. CO2 Methanation over Cobalt Nanoparticles Embedded in ZIF-L–Derived Porous Carbon. ChemCatChem 2023, 15, e202201338. [Google Scholar] [CrossRef]
  57. Dong, Z.; Zhao, J.; Tian, Y.; Zhang, B.; Wu, Y. Preparation and Performances of ZIF-67-Derived FeCo Bimetallic Catalysts for CO2 Hydrogenation to Light Olefins. Catalysts 2020, 10, 455. [Google Scholar] [CrossRef]
  58. Matsagar, B.M.; Yang, R.-X.; Dutta, S.; Ok, Y.S.; Wu, K.C.-W. Recent progress in the development of biomass-derived nitrogen-doped porous carbon. J. Mater. Chem. A 2021, 9, 3703–3728. [Google Scholar] [CrossRef]
  59. Cheng, N.; Ren, L.; Xu, X.; Du, Y.; Dou, S.X. Recent Development of Zeolitic Imidazolate Frameworks (ZIFs) Derived Porous Carbon Based Materials as Electrocatalysts. Adv. Energy Mater. 2018, 8, 1801257. [Google Scholar] [CrossRef]
  60. Lee, S.-H.; Hwang, Y.-M.; Byun, T.-S.; Ko, J.-H.; Roh, J.-S. Effect of heating rate, temperature, and residence time during graphitization on the mechanical and electrical properties of isotropic graphite blocks. Carbon 2023, 208, 443–451. [Google Scholar] [CrossRef]
  61. Geng, Y.; Wang, J.; Chen, X.; Wang, Q.; Zhang, S.; Tian, Y.; Liu, C.; Wang, L.; Wei, Z.; Cao, L.; et al. In Situ N,O-Dually Doped Nanoporous Biochar Derived from Waste Eutrophic Spirulina for High-Performance Supercapacitors. Nanomaterials 2023, 13, 2431. [Google Scholar] [CrossRef]
  62. Su, Y.; Shi, Y.; Jiang, M.; Chen, S. One-Step Synthesis of Nitrogen-Doped Porous Biochar Based on N-Doping Co-Activation Method and Its Application in Water Pollutants Control. Int. J. Mol. Sci. 2022, 23, 14618. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, C.; Chen, Z.; Kang, R.; Wang, J.; Lu, Q.; Wang, T.; Tian, D.; Xu, Y.; Wang, Z.; Ding, H. N-Rich Algal Sludge Biochar for Peroxymonosulfate Activation toward Sulfadiazine Removal. Coatings 2023, 13, 431. [Google Scholar] [CrossRef]
  64. Preparation of highly porous nitrogen-doped biochar derived from birch tree wastes with superior dye removal performance. Colloids Surf. A Physicochem. Eng. Asp. 2023, 669, 131493. [CrossRef]
  65. Wang, X.; Liu, Y.; Zhu, L.; Li, Y.; Wang, K.; Qiu, K.; Tippayawong, N.; Aggarangsi, P.; Reubroycharoen, P.; Wang, S. Biomass derived N-doped biochar as efficient catalyst supports for CO2 methanation. J. CO2 Util. 2019, 34, 733–741. [Google Scholar] [CrossRef]
  66. Zhong, L.; Qiu, X.; Yang, S.; Sun, S.; Chen, L.; Zhang, W. Supermolecule-regulated synthesis strategy of general biomass-derived highly nitrogen-doped carbons toward potassium-ion hybrid capacitors with enhanced performances. Energy Storage Mater. 2023, 61, 102887. [Google Scholar] [CrossRef]
  67. Sudhakaran, M.; Raju, R.; Youk, J.H. Polypyrrole-derived N-doped CNT nanocomposites decorated with CoNi alloy nanoparticles for high-performance supercapacitor electrodes. Appl. Surf. Sci. 2023, 619, 156796. [Google Scholar] [CrossRef]
  68. Li, Y.; Wang, Y.; Chen, B.; Wang, L.; Yang, J.; Wang, B. Nitrogen-doped hierarchically constructed interconnected porous carbon nanofibers derived from polyaniline (PANI) for highly selective CO2 capture and effective methanol adsorption. J. Environ. Chem. Eng. 2022, 10, 108847. [Google Scholar] [CrossRef]
  69. Shi, P.-C.; Yi, J.-D.; Liu, T.-T.; Li, L.; Zhang, L.-J.; Sun, C.-F.; Wang, Y.-B.; Huang, Y.-B.; Cao, R. Hierarchically porous nitrogen-doped carbon nanotubes derived from core–shell ZnO@zeolitic imidazolate framework nanorods for highly efficient oxygen reduction reactions. J. Mater. Chem. A 2017, 5, 12322–12329. [Google Scholar] [CrossRef]
  70. Zhang, H.; Xu, J.; Jin, Y.; Tong, Y.; Lu, Q.; Gao, F. Quantum Effects Allow the Construction of Two-Dimensional Co3O4 -Embedded Nitrogen-Doped Porous Carbon Nanosheet Arrays from Bimetallic MOFs as Bifunctional Oxygen Electrocatalysts. Chem. Eur. J. 2018, 24, 14522–14530. [Google Scholar] [CrossRef]
  71. Zheng, L.; Yu, S.; Lu, X.; Fan, W.; Chi, B.; Ye, Y.; Shi, X.; Zeng, J.; Li, X.; Liao, S. Two-Dimensional Bimetallic Zn/Fe-Metal-Organic Framework (MOF)-Derived Porous Carbon Nanosheets with a High Density of Single/Paired Fe Atoms as High-Performance Oxygen Reduction Catalysts. ACS Appl. Mater. Interfaces 2020, 12, 13878–13887. [Google Scholar] [CrossRef]
  72. Chen, K.; Li, Y.; Wang, M.; Wang, Y.; Cheng, K.; Zhang, Q.; Kang, J.; Wang, Y. Functionalized Carbon Materials in Syngas Conversion. Small 2021, 17, 2007527. [Google Scholar] [CrossRef] [PubMed]
  73. Rahmati, M.; Huang, B.; Mortensen, M.K.; Keyvanloo, K.; Fletcher, T.H.; Woodfield, B.F.; Hecker, W.C.; Argyle, M.D. Effect of different alumina supports on performance of cobalt Fischer-Tropsch catalysts. J. Catal. 2018, 359, 92–100. [Google Scholar] [CrossRef]
  74. Campisi, S.; Chan-Thaw, C.E.; Villa, A. Understanding Heteroatom-Mediated Metal–Support Interactions in Functionalized Carbons: A Perspective Review. Appl. Sci. 2018, 8, 1159. [Google Scholar] [CrossRef]
  75. Kafrani, M.R.; Beni, A.A.; Nodeh, Z.P. Preparation of cobalt-ruthenium nanocatalysts supported on nitrogen-doped graphene aerogel and carbon nanotubes in Fischer-Tropsch synthesis. Arab. J. Chem. 2023, 16, 104914. [Google Scholar] [CrossRef]
  76. Wang, B.; Han, Y.; Chen, S.; Zhang, Y.; Li, J.; Hong, J. Construction of three-dimensional nitrogen-doped graphene aerogel (NGA) supported cobalt catalysts for Fischer-Tropsch synthesis. Catal. Today 2020, 355, 10–16. [Google Scholar] [CrossRef]
  77. Zhan, W.; Wang, Y.; Chen, J.; Chen, Z.; Li, Y. Boosting the Fischer-Tropsch synthesis performances of cobalt-based catalysts via geometric and electronic engineering: Construction of hollow structures. Appl. Catal. B Environ. 2022, 313, 121469. [Google Scholar] [CrossRef]
  78. Taghavi, S.; Tavasoli, A.; Asghari, A.; Signoretto, M. Loading and promoter effects on the performance of nitrogen functionalized graphene nanosheets supported cobalt Fischer-Tropsch synthesis catalysts. Int. J. Hydrog. Energy 2019, 44, 10604–10615. [Google Scholar] [CrossRef]
  79. Chernyak, S.; Stolbov, D.; Ivanov, A.; Klokov, S.; Egorova, T.; Maslakov, K.; Eliseev, O.; Maximov, V.; Savilov, S.; Lunin, V. Effect of type and localization of nitrogen in graphene nanoflake support on structure and catalytic performance of Co-based Fischer-Tropsch catalysts. Catal. Today 2020, 357, 193–202. [Google Scholar] [CrossRef]
  80. Dlamini, M.W.; Phaahlamohlaka, T.N.; Kumi, D.O.; Forbes, R.; Jewell, L.L.; Coville, N.J. Post doped nitrogen-decorated hollow carbon spheres as a support for Co Fischer-Tropsch catalysts. Catal. Today 2020, 342, 99–110. [Google Scholar] [CrossRef]
  81. Cheng, Q.; Zhao, N.; Lyu, S.; Tian, Y.; Gao, F.; Dong, L.; Jiang, Z.; Zhang, J.; Tsubaki, N.; Li, X. Tuning interaction between cobalt catalysts and nitrogen dopants in carbon nanospheres to promote Fischer-Tropsch synthesis. Appl. Catal. B Environ. 2019, 248, 73–83. [Google Scholar]
  82. Guo, S.; Niu, C.; Ma, Z.; Wang, J.; Hou, B.; Jia, L.; Li, D. A novel and facile strategy to decorate Al2O3 as an effective support for Co-based catalyst in Fischer-Tropsch synthesis. Fuel 2021, 289, 119780. [Google Scholar] [CrossRef]
  83. Li, S.; Yao, N.; Zhao, F.; Li, X. Nitrogen-doped carbon species: A promising nonmetallic promoter for the Co/SiO2 Fischer–Tropsch synthesis catalyst. Catal. Sci. Technol. 2016, 6, 2188–2194. [Google Scholar]
  84. Yang, Y.; Jia, L.; Hou, B.; Li, D.; Wang, J.; Sun, Y. The Correlation of Interfacial Interaction and Catalytic Performance of N-Doped Mesoporous Carbon Supported Cobalt Nanoparticles for Fischer–Tropsch Synthesis. J. Phys. Chem. C 2014, 118, 268–277. [Google Scholar]
  85. Fu, T.; Liu, R.; Lv, J.; Li, Z. Influence of acid treatment on N-doped multi-walled carbon nanotube supports for Fischer–Tropsch performance on cobalt catalyst. Fuel Process. Technol. 2014, 122, 49–57. [Google Scholar] [CrossRef]
  86. Fu, T.; Li, Z. Highly dispersed cobalt on N-doped carbon nanotubes with improved Fischer–Tropsch synthesis activity. Catal. Commun. 2014, 47, 54–57. [Google Scholar] [CrossRef]
  87. Yang, Y.; Jia, L.; Hou, B.; Li, D.; Wang, J.; Sun, Y. The Effect of Nitrogen on the Autoreduction of Cobalt Nanoparticles Supported on Nitrogen-Doped Ordered Mesoporous Carbon for the Fischer-Tropsch Synthesis. ChemCatChem 2014, 6, 319–327. [Google Scholar] [CrossRef]
  88. Chai, J.; Jiang, J.; Gong, Y.; Wu, P.; Wang, A.; Zhang, X.; Wang, T.; Meng, X.; Lin, Q.; Lv, Y.; et al. Recent Mechanistic Understanding of Fischer-Tropsch Synthesis on Fe-Carbide. Catalysts 2023, 13, 1052. [Google Scholar]
  89. Liu, G.; Chen, Q.; Oyunkhand, E.; Ding, S.; Yamane, N.; Yang, G.; Yoneyama, Y.; Tsubaki, N. Nitrogen-rich mesoporous carbon supported iron catalyst with superior activity for Fischer-Tropsch synthesis. Carbon 2018, 130, 304–314. [Google Scholar] [CrossRef]
  90. Tian, Z.; Wang, C.; Si, Z.; Wen, C.; Xu, Y.; Lv, W.; Chen, L.; Zhang, X.; Ma, L. Enhancement of Light Olefins Selectivity Over N-Doped Fischer-Tropsch Synthesis Catalyst Supported on Activated Carbon Pretreated with KMnO4. Catalysts 2019, 9, 505. [Google Scholar] [CrossRef]
  91. Lee, J.H.; Lee, H.-K.; Chun, D.H.; Choi, H.; Rhim, G.B.; Youn, M.H.; Jeong, H.; Kang, S.W.; Yang, J.-I.; Jung, H.; et al. Phase-controlled synthesis of thermally stable nitrogen-doped carbon supported iron catalysts for highly efficient Fischer-Tropsch synthesis. Nano Res. 2019, 12, 2568–2575. [Google Scholar] [CrossRef]
  92. Tang, L.; Dong, X.-L.; Xu, W.; He, L.; Lu, A.-H. Iron-based catalysts encapsulated by nitrogen-doped graphitic carbon for selective synthesis of liquid fuels through the Fischer-Tropsch process. Chin. J. Catal. 2018, 39, 1971–1979. [Google Scholar] [CrossRef]
  93. Hwang, J.; Ejsmont, A.; Freund, R.; Goscianska, J.; Schmidt, B.V.K.J.; Wuttke, S. Controlling the morphology of metal–organic frameworks and porous carbon materials: Metal oxides as primary architecture-directing agents. Chem. Soc. Rev. 2020, 49, 3348–3422. [Google Scholar] [CrossRef]
  94. Chaikittisilp, W.; Hu, M.; Wang, H.; Huang, H.-S.; Fujita, T.; Wu, K.C.-W.; Chen, L.-C.; Yamauchi, Y.; Ariga, K. Nanoporous carbons through direct carbonization of a zeolitic imidazolate framework for supercapacitor electrodes. Chem. Commun. 2012, 48, 7259–7261. [Google Scholar] [CrossRef] [PubMed]
  95. Jiang, Y.; Liu, H.; Tan, X.; Guo, L.; Zhang, J.; Liu, S.; Guo, Y.; Zhang, J.; Wang, H.; Chu, W. Monoclinic ZIF-8 Nanosheet-Derived 2D Carbon Nanosheets as Sulfur Immobilizer for High-Performance Lithium Sulfur Batteries. ACS Appl. Mater. Interfaces 2017, 9, 25239–25249. [Google Scholar] [CrossRef] [PubMed]
  96. Zhao, Q.; Han, X.; Liang, H.; Wang, Y.; Lv, J.; Wang, M.-Y.; Huang, S.; Ma, X. Activating nitrogen-doped carbon nanosheets by KOH treatment to promote the Fischer-Tropsch synthesis performance. Chem. Eng. J. 2023, 455, 140810. [Google Scholar] [CrossRef]
  97. Park, B.J.; Jang, S.; Lee, J.H.; Chun, D.H.; Park, J.C.; Park, H.S. Hyperactive iron carbide@N-doped reduced graphene oxide/carbon nanotube hybrid architecture for rapid CO hydrogenation. J. Mater. Chem. A 2018, 6, 11134–11139. [Google Scholar] [CrossRef]
  98. Liu, R.; Liu, R.; Ma, X.; Davis, B.H.; Li, Z. Efficient diesel production over the iron-based Fischer–Tropsch catalyst supported on CNTs treated by urea/NaOH. Fuel 2018, 211, 827–836. [Google Scholar] [CrossRef]
  99. Lama, S.M.; Weber, J.L.; Heil, T.; Hofmann, J.P.; Yan, R.; de Jong, K.P.; Oschatz, M. Tandem promotion of iron catalysts by sodium-sulfur and nitrogen-doped carbon layers on carbon nanotube supports for the Fischer-Tropsch to olefins synthesis. Appl. Catal. A Gen. 2018, 568, 213–220. [Google Scholar] [CrossRef]
  100. Wu, J.; Qin, L.; Wang, C.; Lv, B.; Wang, L.; Chen, J.; Xu, Y. Ultrathin N-rich boron nitride nanosheets supported iron catalyst for Fischer–Tropsch synthesis. RSC Adv. 2016, 6, 38356–38364. [Google Scholar] [CrossRef]
  101. An, B.; Cheng, K.; Wang, C.; Wang, Y.; Lin, W. Pyrolysis of Metal–Organic Frameworks to Fe3O4@Fe5C2 Core–Shell Nanoparticles for Fischer–Tropsch Synthesis. ACS Catal. 2016, 6, 3610–3618. [Google Scholar] [CrossRef]
  102. Li, Z.; Liu, R.; Xu, Y.; Ma, X. Enhanced Fischer–Tropsch synthesis performance of iron-based catalysts supported on nitric acid treated N-doped CNTs. Appl. Surf. Sci. 2015, 347, 643–650. [Google Scholar] [CrossRef]
  103. Chen, X.; Deng, D.; Pan, X.; Hu, Y.; Bao, X. N-doped graphene as an electron donor of iron catalysts for CO hydrogenation to light olefins. Chem. Commun. 2015, 51, 217–220. [Google Scholar] [CrossRef] [PubMed]
  104. Xiong, H.; Moyo, M.; Motchelaho, M.A.; Tetana, Z.N.; Dube, S.M.; Jewell, L.L.; Coville, N.J. Fischer–Tropsch synthesis: Iron catalysts supported on N-doped carbon spheres prepared by chemical vapor deposition and hydrothermal approaches. J. Catal. 2014, 311, 80–87. [Google Scholar] [CrossRef]
  105. Xiong, H.; Motchelaho, M.A.; Moyo, M.; Jewell, L.L.; Coville, N.J. Fischer–Tropsch synthesis: Iron-based catalysts supported on nitrogen-doped carbon nanotubes synthesized by post-doping. Appl. Catal. A Gen. 2014, 482, 377–386. [Google Scholar] [CrossRef]
  106. Xiong, H.; Moyo, M.; Rayner, M.K.; Jewell, L.L.; Billing, D.G.; Coville, N.J. Autoreduction and Catalytic Performance of a Cobalt Fischer-Tropsch Synthesis Catalyst Supported on Nitrogen-Doped Carbon Spheres. ChemCatChem 2010, 2, 514–518. [Google Scholar] [CrossRef]
  107. 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] [PubMed]
  108. Gao, R.; Wang, L.; Zhang, L.; Zhang, C.; Jun, K.-W.; Kim, S.K.; Park, H.-G.; Zhao, T.; Gao, Y.; Zhu, Y.; et al. Upcycling of CO2 into sustainable hydrocarbon fuels via the integration of Fe-based Fischer-Tropsch synthesis and olefin oligomerization: A comparative case study. Fuel 2022, 325, 124855. [Google Scholar] [CrossRef]
  109. Arslan, M.T.; Tian, G.; Ali, B.; Zhang, C.; Xiong, H.; Li, Z.; Luo, L.; Chen, X.; Wei, F. Highly Selective Conversion of CO2 or CO into Precursors for Kerosene-Based Aviation Fuel via an Aldol–Aromatic Mechanism. ACS Catal. 2022, 12, 2023–2033. [Google Scholar] [CrossRef]
  110. Li, Y.; Cai, X.; Chen, S.; Zhang, H.; Zhang, K.H.L.; Hong, J.; Chen, B.; Kuo, D.-H.; Wang, W. Highly Dispersed Metal Carbide on ZIF-Derived Pyridinic-N-Doped Carbon for CO2 Enrichment and Selective Hydrogenation. ChemSusChem 2018, 11, 1040–1047. [Google Scholar] [CrossRef]
  111. Arpini, B.H.; Braga, A.H.; Borges, L.R.; Vidinha, P.; Gonçalves, R.V.; Szanyi, J.; Rossi, L.M. Tuning CO2 Hydrogenation Selectivity by N-Doped Carbon Coating over Nickel Nanoparticles Supported on SiO2. ACS Sustain. Chem. Eng. 2022, 10, 2331–2342. [Google Scholar] [CrossRef]
  112. Li, Y.; Zhao, Z.; Lu, W.; Zhu, H.; Sun, F.; Mei, B.; Jiang, Z.; Lyu, Y.; Chen, X.; Guo, L.; et al. Single-atom Co-N-C catalysts for high-efficiency reverse water-gas shift reaction. Appl. Catal. B Environ. 2023, 324, 122298. [Google Scholar] [CrossRef]
  113. Zhang, J.; Yang, B.; Luo, K.H. Unveiling the mechanism of controllable CO2 hydrogenation by group VIB metal single atom anchored on N-doped graphite: A density functional theory study. Int. J. Hydrog. Energy 2022, 47, 40972–40985. [Google Scholar] [CrossRef]
  114. Jiang, Y.; Sung, Y.; Choi, C.; Joo Bang, G.; Hong, S.; Tan, X.; Wu, T.-S.; Soo, Y.-L.; Xiong, P.; Li, M.M.-J.; et al. Single-Atom Molybdenum-N3 Sites for Selective Hydrogenation of CO2 to CO. Angew. Chem. 2022, 134, e202203836. [Google Scholar] [CrossRef]
  115. Hasan, M.F.; Rossi, L.M.; Debecker, D.P.; Leonard, K.C.; Li, Z.; Makhubela, B.C. Can CO2 and Renewable Carbon Be Primary Resources for Sustainable Fuels and Chemicals? ACS Sustain. Chem. Eng. 2021, 9, 12427–12430. [Google Scholar] [CrossRef]
  116. Hidalgo, D.; Martín-Marroquín, J. Power-to-methane, coupling CO2 capture with fuel production: An overview. Renew. Sustain. Energy Rev. 2020, 132, 110057. [Google Scholar] [CrossRef]
  117. Di Stasi, C.; Renda, S.; Greco, G.; González, B.; Palma, V.; Manyà, J.J. Wheat-Straw-Derived Activated Biochar as a Renewable Support of Ni-CeO2 Catalysts for CO2 Methanation. Sustainability 2021, 13, 8939. [Google Scholar] [CrossRef]
  118. Roldán, L.; Marco, Y.; García-Bordejé, E. Origin of the Excellent Performance of Ru on Nitrogen-Doped Carbon Nanofibers for CO2 Hydrogenation to CH4. ChemSusChem 2017, 10, 1139–1144. [Google Scholar] [CrossRef]
  119. Wu, J.; Wen, C.; Zou, X.; Jimenez, J.; Sun, J.; Xia, Y.; Rodrigues, M.-T.F.; Vinod, S.; Zhong, J.; Chopra, N.; et al. Carbon Dioxide Hydrogenation over a Metal-Free Carbon-Based Catalyst. ACS Catal. 2017, 7, 4497–4503. [Google Scholar] [CrossRef]
  120. Wang, W.; Duong-Viet, C.; Ba, H.; Baaziz, W.; Tuci, G.; Caporali, S.; Nguyen-Dinh, L.; Ersen, O.; Giambastiani, G.; Pham-Huu, C. Nickel Nanoparticles Decorated Nitrogen-Doped Carbon Nanotubes (Ni/N-CNT); a Robust Catalyst for the Efficient and Selective CO2 Methanation. ACS Appl. Energy Mater. 2019, 2, 1111–1120. [Google Scholar] [CrossRef]
  121. Gonçalves, L.P.; Meledina, M.; Meledin, A.; Petrovykh, D.Y.; Sousa, J.P.; Soares, O.S.G.; Kolen’Ko, Y.V.; Pereira, M.F.R. Understanding the importance of N-doping for CNT-supported Ni catalysts for CO2 methanation. Carbon 2022, 195, 35–43. [Google Scholar] [CrossRef]
  122. Gödde, J.; Merko, M.; Xia, W.; Muhler, M. Nickel nanoparticles supported on nitrogen–doped carbon nanotubes are a highly active, selective and stable CO2 methanation catalyst. J. Energy Chem. 2021, 54, 323–331. [Google Scholar] [CrossRef]
  123. Weber, D.; Rui, N.; Zhang, F.; Zhang, H.; Vovchok, D.; Wildy, M.; Arizapana, K.; Saporita, A.; Zhang, J.Z.; Senanayake, S.D. Carbon Nanosphere-Encapsulated Fe Core–Shell Structures for Catalytic CO2 Hydrogenation. ACS Appl. Nano Mater. 2022, 5, 11605–11616. [Google Scholar] [CrossRef]
  124. Liu, J.; Sun, Y.; Jiang, X.; Zhang, A.; Song, C.; Guo, X. Pyrolyzing ZIF-8 to N-doped porous carbon facilitated by iron and potassium for CO2 hydrogenation to value-added hydrocarbons. J. CO2 Util. 2018, 25, 120–127. [Google Scholar] [CrossRef]
  125. Zhang, P.; Han, F.; Yan, J.; Qiao, X.; Guan, Q.; Li, W. N-doped ordered mesoporous carbon (N-OMC) confined Fe3O4-FeCx heterojunction for efficient conversion of CO2 to light olefins. Appl. Catal. B Environ. 2021, 299, 120639. [Google Scholar] [CrossRef]
  126. Peng, L.; Jurca, B.; Primo, A.; Gordillo, A.; Parvulescu, V.I.; García, H. High C2-C4 selectivity in CO2 hydrogenation by particle size control of Co-Fe alloy nanoparticles wrapped on N-doped graphitic carbon. iScience 2022, 25, 104252. [Google Scholar] [CrossRef]
  127. LLimleamthong, P.; Chuchuan, A.; Thepphankulngarm, N.; Kongkachuichay, P. Structured Porous Carbon-Based Catalysts: Cu–ZnO/CMK-3 and Cu–CeO2/CMK-3 for Direct CO2 Conversion to Methanol. Top. Catal. 2023. [Google Scholar] [CrossRef]
  128. Sun, Y.; Chen, L.; Bao, Y.; Wang, G.; Zhang, Y.; Fu, M.; Wu, J.; Ye, D. Roles of nitrogen species on nitrogen-doped CNTs supported Cu-ZrO2 system for carbon dioxide hydrogenation to methanol. Catal. Today 2018, 307, 212–223. [Google Scholar] [CrossRef]
  129. Deerattrakul, V.; Yigit, N.; Rupprechter, G.; Kongkachuichay, P. The roles of nitrogen species on graphene aerogel supported Cu-Zn as efficient catalysts for CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2019, 580, 46–52. [Google Scholar] [CrossRef]
  130. Ma, Q.; Geng, M.; Zhang, J.; Zhang, X.; Zhao, T.-S. Enhanced Catalytic Performance for CO2 Hydrogenation to Methanol over N-doped Graphene Incorporated Cu-ZnO-Al2O3 Catalysts. ChemistrySelect 2019, 4, 78–83. [Google Scholar] [CrossRef]
  131. Donphai, W.; Thepphankulngarm, N.; Chaisuwan, T.; Tanangteerapong, D.; Rood, S.C.; Kongkachuichay, P. Catalytic performance of copper and ruthenium loaded on N-doped modified PBZ-derived carbons for CO2 hydrogenation. Chem. Eng. Sci. 2023, 274, 118693. [Google Scholar] [CrossRef]
  132. Moret, S.; Dyson, P.J.; Laurenczy, G. Direct synthesis of formic acid from carbon dioxide by hydrogenation in acidic media. Nat. Commun. 2014, 5, 4017. [Google Scholar] [CrossRef]
  133. Yang, G.; Kuwahara, Y.; Mori, K.; Louis, C.; Yamashita, H. PdAg alloy nanoparticles encapsulated in N-doped microporous hollow carbon spheres for hydrogenation of CO2 to formate. Appl. Catal. B Environ. 2021, 283, 119628. [Google Scholar] [CrossRef]
  134. Ahn, S.; Park, K.; Lee, K.R.; Haider, A.; Van Nguyen, C.; Jin, H.; Yoo, S.J.; Yoon, S.; Jung, K.-D. Atomically dispersed Ru(III) on N-doped mesoporous carbon hollow spheres as catalysts for CO2 hydrogenation to formate. Chem. Eng. J. 2022, 442, 136185. [Google Scholar] [CrossRef]
  135. Jaleel, A.; Haider, A.; Van Nguyen, C.; Lee, K.R.; Choung, S.; Han, J.W.; Baek, S.-H.; Shin, C.-H.; Jung, K.-D. Structural effect of Nitrogen/Carbon on the stability of anchored Ru catalysts for CO2 hydrogenation to formate. Chem. Eng. J. 2022, 433, 133571. [Google Scholar] [CrossRef]
  136. Zhang, Y.; Mo, Y.; Cao, Z. Rational Design of Main Group Metal-Embedded Nitrogen-Doped Carbon Materials as Frustrated Lewis Pair Catalysts for CO2 Hydrogenation to Formic Acid. ACS Appl. Mater. Interfaces 2022, 14, 1002–1014. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of three fresh (left) and spent (right) Co/NGA catalysts [76]. Copyright (2020) Catalysis Today.
Figure 1. XRD patterns of three fresh (left) and spent (right) Co/NGA catalysts [76]. Copyright (2020) Catalysis Today.
Atmosphere 14 01510 g001
Figure 2. Fabrication strategy of H-Co@NCNHP [77]. Copyright (2022) Applied Catalysis B: Environmental.
Figure 2. Fabrication strategy of H-Co@NCNHP [77]. Copyright (2022) Applied Catalysis B: Environmental.
Atmosphere 14 01510 g002
Figure 3. CO-TPD/MS profiles of the reduced Co/NCS-T catalysts [81]. Copyright (2019) Applied Catalysis B: Environmental.
Figure 3. CO-TPD/MS profiles of the reduced Co/NCS-T catalysts [81]. Copyright (2019) Applied Catalysis B: Environmental.
Atmosphere 14 01510 g003
Figure 4. Synthetic scheme of carbon encapsulated iron-based nanoparticles supported on nitrogen-doped porous carbon nanostructures [91]. Copyright (2019) Nano Research.
Figure 4. Synthetic scheme of carbon encapsulated iron-based nanoparticles supported on nitrogen-doped porous carbon nanostructures [91]. Copyright (2019) Nano Research.
Atmosphere 14 01510 g004
Figure 5. TEM images of 40Fe/N1(a), TEM-mapping of 40Fe/N1(b) [89]. Copyright (2018) Carbon.
Figure 5. TEM images of 40Fe/N1(a), TEM-mapping of 40Fe/N1(b) [89]. Copyright (2018) Carbon.
Atmosphere 14 01510 g005
Figure 6. Schematic illustration of the synthesis process for preparing Fe/CNS catalyst [55]. Copyright (2019) Carbon.
Figure 6. Schematic illustration of the synthesis process for preparing Fe/CNS catalyst [55]. Copyright (2019) Carbon.
Atmosphere 14 01510 g006
Figure 7. CO conversions on TOS (a); TEM images (bd) of the used Fe/CNS(1000), Fe/CP, and Fe/AC [55]. Copyright (2019) Carbon.
Figure 7. CO conversions on TOS (a); TEM images (bd) of the used Fe/CNS(1000), Fe/CP, and Fe/AC [55]. Copyright (2019) Carbon.
Atmosphere 14 01510 g007
Figure 8. XPS wide-survey spectra (a) and high-resolution N1s spectra of the CNTs-110 (b), CNTs-190 (c), and CNTs-220 (d) [98]. Copyright (2018) Fuel.
Figure 8. XPS wide-survey spectra (a) and high-resolution N1s spectra of the CNTs-110 (b), CNTs-190 (c), and CNTs-220 (d) [98]. Copyright (2018) Fuel.
Atmosphere 14 01510 g008
Figure 9. CO2 conversion of M/ZIF-8-C [110]. Copyright (2018) ChemSusChem.
Figure 9. CO2 conversion of M/ZIF-8-C [110]. Copyright (2018) ChemSusChem.
Atmosphere 14 01510 g009
Figure 10. Reaction coordinate of CO2 hydrogenation [110]. Copyright (2018) ChemSusChem.
Figure 10. Reaction coordinate of CO2 hydrogenation [110]. Copyright (2018) ChemSusChem.
Atmosphere 14 01510 g010
Figure 11. HAADF-STEM image of the 5% Co-N-C catalyst [112]. Copyright (2022) Angewandte Chemie.
Figure 11. HAADF-STEM image of the 5% Co-N-C catalyst [112]. Copyright (2022) Angewandte Chemie.
Atmosphere 14 01510 g011
Figure 12. N types and contents of catalysts [57]. Copyright (2020) Catalysts.
Figure 12. N types and contents of catalysts [57]. Copyright (2020) Catalysts.
Atmosphere 14 01510 g012
Figure 13. (ad) HR-TEM images of 0.8Fe@N-OMC; (e,f) the HAADF-STEM image and element mapping of 0.8Fe@N-OMC [125]. Copyright (2021) Applied Catalysis B: Environmental.
Figure 13. (ad) HR-TEM images of 0.8Fe@N-OMC; (e,f) the HAADF-STEM image and element mapping of 0.8Fe@N-OMC [125]. Copyright (2021) Applied Catalysis B: Environmental.
Atmosphere 14 01510 g013
Figure 14. Binding energies of RuCl3 on N-coordinated graphitic carbon structures [136]. Copyright (2022) Chemical Engineering Journal.
Figure 14. Binding energies of RuCl3 on N-coordinated graphitic carbon structures [136]. Copyright (2022) Chemical Engineering Journal.
Atmosphere 14 01510 g014
Table 1. Performance of various cobalt-based catalysts in CO hydrogenation.
Table 1. Performance of various cobalt-based catalysts in CO hydrogenation.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:CO
Ratio
CO Conversion
(%)
C5+ SelectivityReference
Co-Ru/NGA2528010,000 b175.2517.2[75]
2.7%Co/NGA102502000 b29.088.2[76]
H-Co@NCNHP2025028,400 b216.490.4[77]
Co/GHFox202406000 *231.930.0[79]
10Co/N-HCSs90010220N.A.234.075.8[80]
Co/NCS-50020220N.A.2N.A.74.4[81]
Co/Al2O3-15CN20250N.A.271.377.8[82]
Co/SiO2-CN2025027,000 b234.476.3[83]
10Co/N-MC202301000 a2N.A.73.0[84]
Co/NCNTs-2202306750 b251.283.4[85]
Co/A-NCNTs202306750 b274.380.0[86]
15Co/NMC-2202401000 a2N.A.61.2[87]
a GHSV, h−1; b WHSV, mL/g/h; * NmL/g/h.
Table 2. Performance of various iron-based catalysts in CO hydrogenation.
Table 2. Performance of various iron-based catalysts in CO hydrogenation.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:Co
Ratio
CO Conversion
(%)
C5+
Selectivity
Reference
Fe/CNS(1000)103409000 b145.939.2[55]
15%Fe/AG(12h)-W(10)203205 #297.240.0[29]
40Fe/N1102605 #174.767.0[89]
FeN-AC2032015,000 a159.718.1[90]
Fe5C2@C/NPC1534042 *196.433.5[91]
FeC-80020300N.A.247.052.0[92]
Fe/CNS-KOH(700, 0.6)103409000 b187.055.6[96]
Fe5C2@Ns-rGO/CNT15340210 *180.530.0[97]
Fe/CNTs-NaU202704500 b160.090.8[98]
FeBN2202701500 a247.048.0[100]
Fe-MIL-88B-NH2/C2030036,000 a181.850.7[101]
Fe/NCNTs-10202704500 b145.076.0[102]
Fe/NG−16.45340600 a121.119.8[103]
Fe/NCSver82752700 aN.A.50.051.6[104]
Fe/N-CNT-h82752400 a270.160.9[105]
2.3 wt% Co/N-CSs82301200 a2N.A.47.1[106]
a GHSV, h−1; b WHSV, mL/g/h; * NmL/g/h; # g·h/mol.
Table 3. Performance of various catalysts in CO2 hydrogenation to CO.
Table 3. Performance of various catalysts in CO2 hydrogenation to CO.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:CO
Ratio
CO2 Conversion
(%)
CO
Selectivity
Reference
Ni/ZIF-8-C142015,000 b443.842.7[82]
Fe/ZIF-8-C142015,000 b443.842.6[82]
5%Co-N-C15006000 b452.498.3[84]
Mo/NC15008000 b346.3100[86]
b WHSV, mL/g/h.
Table 4. Performance of various catalysts in CO2 hydrogenation to CH4.
Table 4. Performance of various catalysts in CO2 hydrogenation to CH4.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:Co
Ratio
CO2 Conversion
(%)
CH4
Selectivity
Reference
BCCe30Ni20140013,200 a465.095.0[27]
NGQDs/Al2O3140018,000 b461.855.2[117]
Ni/N-CNTs1360120,000 b475.094.0[118]
60,000 b81.098.0[118]
Ru/N-ABC-600103806000 b493.899.7[65]
Co/C(L)130012,000 b444.084.0[56]
Co/C(67)130012,000 b433.056.0[56]
Ni/CNT-N140060,000 b481.299.2[119]
40Ni/NCNT134050,000 b451.495.8[120]
a GHSV, h−1; b WHSV, mL/g/h.
Table 5. Performance of various catalysts in CO2 hydrogenation to C2+.
Table 5. Performance of various catalysts in CO2 hydrogenation to C2+.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:CO
Ratio
CO2 Conversion
(%)
C2+
Selectivity
Reference
FeZn-NC303207200 b329.363.5 [124]
FeCo/NC-600203206240 b337.049.4 [58]
0.8Fe@N-OMC303204800 b354.582.9 [127]
Co-Fe@(N)403006000 a458.044.0 [128]
a GHSV, h−1; b WHSV, mL/g/h.
Table 6. Performance of various catalysts in CO2 hydrogenation to CH3OH.
Table 6. Performance of various catalysts in CO2 hydrogenation to CH3OH.
CatalystPressure
(bar)
Temperature
(°C)
GHSVH2:CO
Ratio
CO2 Conversion
(%)
STY of
Methanol (mg g−1cat h−1)
Reference
CZ/NCMK-3U152502444 a343.0512.0[127]
CC/NCMK-3U152502444 a330.0367.0[127]
CZ/CNTs-N302603600 b311.5102.2[128]
15%CuZn/NrGOae-H152502444 a320.5329.4[129]
15%CuZn/NrGOae-A152502444 a316.4264.4[129]
15%CuZn/NrGOae-U152502444 a324.2405.5[129]
10NG-CZA3020010 #38.2N.A.[130]
Cu-Ru-Ce/Am-P123152102400 a327.0530[131]
Cu-Ru-Zr/Am-P123152102400 a337.0642[131]
Cu-Ru-Ce/U-P123152102400 a326.0504[131]
Cu-Ru-Zr/U-P123152102400 a336.0639[131]
a GHSV, h−1; b WHSV, mL/g/h; # g·h/mol.
Table 7. Performance of various catalysts in CO2 hydrogenation to CHOOH.
Table 7. Performance of various catalysts in CO2 hydrogenation to CHOOH.
CatalystPressure
(bar)
Temperature
(°C)
Reaction
Time(h)
H2:CO
Ratio
Formate Conc
(M)
TONReference
[email protected]2010021N.A.640[135]
Ru/N-MCHS-90080120210.9507550[136]
2Ru/NC75080120211.37418,212[136]
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

Deng, C.; Xu, L.; Hu, K.; Chen, X.; Gao, R.; Zhang, L.; Wang, L.; Zhang, C. Research Advances on Nitrogen-Doped Carbon Materials in COx Hydrogenation. Atmosphere 2023, 14, 1510. https://doi.org/10.3390/atmos14101510

AMA Style

Deng C, Xu L, Hu K, Chen X, Gao R, Zhang L, Wang L, Zhang C. Research Advances on Nitrogen-Doped Carbon Materials in COx Hydrogenation. Atmosphere. 2023; 14(10):1510. https://doi.org/10.3390/atmos14101510

Chicago/Turabian Style

Deng, Chao, Lujing Xu, Kehao Hu, Xixi Chen, Ruxing Gao, Leiyu Zhang, Lei Wang, and Chundong Zhang. 2023. "Research Advances on Nitrogen-Doped Carbon Materials in COx Hydrogenation" Atmosphere 14, no. 10: 1510. https://doi.org/10.3390/atmos14101510

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