Next Article in Journal
Theoretical and Experimental Studies on the Visible Light Activity of TiO2 Modified with Halide-Based Ionic Liquids
Next Article in Special Issue
Promoting Effect of Ti Species in MnOx-FeOx/Silicalite-1 for the Low-Temperature NH3-SCR Reaction
Previous Article in Journal
Purifying of Waste Tire Pyrolysis Oil Using an S-ZrO2/SBA-15-H2O2 Catalytic Oxidation Method
Previous Article in Special Issue
Zeolite Beta Doped with La, Fe, and Pd as a Hydrocarbon Trap
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Analysis of Research Status of CO2 Conversion Technology Based on Bibliometrics

1
School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
Beijing Key Laboratory of Resource-oriented Treatment of Industrial Pollutants, University of Science and Technology Beijing, Beijing 100083, China
3
Key Laboratory of Knowledge Automation for Industrial Processes, Ministry of Education, Beijing 100083, China
*
Author to whom correspondence should be addressed.
Submission received: 6 February 2020 / Revised: 21 March 2020 / Accepted: 26 March 2020 / Published: 30 March 2020
(This article belongs to the Special Issue Catalysis for the Removal of Gas-Phase Pollutants)

Abstract

:
The concentration of carbon dioxide in the air has risen sharply due to the use of fossil fuels, causing environmental problems such as the greenhouse effect, which seriously threatens humans’ living environment. Reducing carbon dioxide emissions while addressing energy shortages requires the conversion of CO2 into high added-value products. In this paper, the status of CO2 conversion research in the past ten years is analyzed using the bibliometric method; the influence of countries and institutions, journal article statistics and other aspects are statistically analyzed, and the research status of carbon dioxide catalytic conversion is briefly introduced. Finally, according to the analysis results and the existing problems of CO2 catalytic conversion research, the future development direction of CO2 catalytic conversion research is prospected.

1. Introduction

Due to the burning of fossil fuels, the large amount of CO2 in the atmosphere has exceeded the natural carbon cycle. The resulting series of global environmental climate problems seriously imperils human existence and ecological environment [1]. It is an important challenge, but also an opportunity, for humans to covert CO2 into value-added chemicals and fuels, not only for dealing with climate change, but also for reducing humans’ dependence on fossil energy. The survey has shown that the demand for oil, coal, natural gas and their derivatives will be maintained for a long time in the future, especially with the increasing population; the demand for carbon-based fuels, plastics and drugs will also increase [2]. In order to meet humans’ demand for carbon-based products, the concept of green development has been accepted more and more, so researchers have a strong interest in converting CO2 into high value-added products.
However, the conversion of CO2 into other high value-added products requires a high energy input, because of the highly symmetrical structure of CO2 and the highest oxidation state of carbon. So far, CO2 is mainly converted into C1 compounds (methane, methanol, formic acid, etc.), C2 compound (ethane, acetic acid, ethanol, formate, etc.), cyclic carbonate, dimethyl carbonate, carbonate, decanoate, salicylic acid and other chemical products through catalytic conversion [3], electrochemical conversion [4], enzymatic conversion [5], plasma conversion [6], and bioconversion [7].
Bibliometrics plays an important role in evaluating the research value of a certain field and predicting development directions. At present, bibliometric methods have been applied in many fields, such as balkanized research in ecological engineering [8], the study of mycobacterium chimera [9], drug repurposing [10], and so on. However, in the field of CO2 conversion, few scholars use bibliometric methods for analysis. To fill this gap, this article carries out data processing for relevant research articles on CO2 conversion through the bibliometric method, makes a reasonable analysis of the current research trend of CO2 conversion based on the processing results, and looks forward to the future research prospects.

2. Results and Discussion

2.1. Publication Outputs

Based on the number of articles published annually, the research hotspot changes in this field can be understood briefly and the future development trend of the field can be predicted. As shown in Figure 1, both the total number of articles and the number of research English articles have increased over time. The number of research articles in English increased from 51 in 2010 to 798 in 2019, which is enough to show that research on CO2 conversion is gaining popularity year by year. It is worth mentioning that the number of articles on CO2 conversion research in 2017 (622 articles) was a sharp increase from 2016 (441 articles), which may be because the CO2 content in the atmosphere reached a historical high at that time in 2016, and the formal entry into force of the Paris Climate Agreement has further accelerated the research process. Therefore, it can be predicted that the research intensity in this field will continue to increase in the next few years.

2.2. Journal Analysis

Impact factor (IF) is the most common standard for evaluating the influence of journals. In the same field, the higher the IF value of journals is, the stronger influence and greater reading value journals will have [11]. In addition, the h-index, proposed by Hirsch in 2005 to reflect the influence of the author, is widely accepted and gradually applied to evaluate the influence of countries, institutions and journals [12]. In this paper, the h-index of countries and journals are calculated, based on the filtered data for indicating the influence of countries and journals in the field of CO2 conversion research.
The 3670 screened articles were published in 481 journals, among which 1401 articles were separate subjects and the remaining 2269 were interdisciplinary. The articles are grouped based on the IF value that was published by the Journal Citation Report (JCR) in 2019, and there were 59 articles published in 0.1 ≤ IF2018 < 1 journals; 2886 articles published in 1 ≤ IF2018 < 10 journals; 690 articles in the 10 ≤ IF2018 journals. Besides, there are 35 journals not evaluated in JCR 2019.
Table 1 lists the top ten journals in the volume of publications. It can be seen that in the top ten journals, Applied Catalysis B: Environmental and Angewandte Chemie: International Edition has higher impact factors and h-index values, indicating that the two journals have higher impact in this field. The journal with the largest number of articles (214) is Journal of CO2 Utilization, but the journal has a lower h-index of only 33.

2.3. Country/Region Alanalysis

The analysis of country/regional publications provides a simple understanding of the contribution and influence of a country/region in the field of study [13]. The 3670 selected articles involved 74 countries/regions, of which 2750 (74.93%) were completed in separate countries, and 762 (27.71%) articles were completed jointly by two countries. The most collaborative articles were completed by six countries, and there was only one (0.13%) article like this.
Figure 2 shows the relationship between the output of the top ten countries and their h-index. Since 25.07% of the articles have an international cooperative relationship, the actual national participation of 3679 articles is 4784. As we can see, both China and the United States have higher publications and h-indexes than other countries. It is worth noting that, although China (1465, 30.62%) has a higher volume of publications than the United States (671, 14.03%), the h-index is not much different (79:77). That is to say, although China pays much attention to CO2 conversion research and plays a leading role in this field, the research depth needs to be improved. It is also worth mentioning that the two countries of Japan (204, 4.26%), and Spain (117, 2.45%) have a small amount of publications, but have a higher h-index than countries that publish in the same range.

2.4. Institution Analysis

Factors such as the volume of publications and cooperation between institutions in the field of analytical research can determine the influence of each institution. A total of 2074 institutions are involved in the selected data, but 1160 (55.93%) of the institutions appeared only once. Among the 3670 articles, 1077 (29.35%) were completed by a single institution, and the most collaborative article was one (0.02%) with a total of 10 cooperative institutions. Figure 3 is a network diagram of the cooperation between 75 institutions with a number of occurrences higher than 10. The different colors represent different institutions from different continents, and the most frequently occurring institution in Asia is the Chinese Academy of Sciences (283). The largest number of occurrences in the Americas was Penn State University (37); the most frequent occurrence in Europe was University of Antwerp (51).
It can be clearly seen from the diagram that the number of institutions from Asia is the largest, up to 52, accounting for 69.33% of 75 institutions. This means that Asia pays much more attention to CO2 conversion than other continents. At the same time, it can be clearly seen from Figure 3 that the Chinese Academy of Sciences plays a leading role in this field, and its partners mainly focus on Asian institutions; it also has direct cooperation with some institutions in the Americas and Europe.
Table 2 lists the top ten institutions with a large number of documents (TP > 30), using the volume of publications and h-index to indicate the influence of the top ten institutions in the field of CO2 conversion. It can be seen from the table that the Chinese Academy of Sciences is superior to other institutions, both in terms of the volume of publications and h-index, which indicates that the Chinese Academy of Sciences has a high degree of attention in this field and has a certain influence in this field. In addition, the top ten organizations in this issue come from three different countries, including six institutions from China, which further confirms the fact that China has high concern and influence in this field.

2.5. Research Topic Analysis

Among the filtered articles, there are 1037 (28.26%) articles without author keywords, and nearly 21.50% (223/1037) of the articles were published in 2019. Therefore, it is not scientifically rigorous to analyze the research trends in the field of CO2 conversion research based on author keywords. At the same time, since different authors have individualized differences in the expression of the same noun, keywords with different expressions of the same meaning need to be combined.
As shown in Figure 4, the keywords after screening and classification can be divided into five categories: CO2 conversion mode, reaction process, product/reactant, theoretical research and other CO2 related keywords. It can be seen from the figure that during the decade of 2010-2019, the conversion mode of CO2 roughly divides into catalytic conversion, direct electrochemical reduction conversion, biotransformation, enzymatic conversion, plasma conversion, etc.

2.6. Catalytic Conversion

For CO2 conversion, the most-studied technology is catalytic conversion technology, and this conclusion can be clearly seen from Figure 4. In this method, technologies such as photocatalysis, electrocatalysis, plasma catalysis, and conventional thermal catalysis are widely studied. The mechanism, catalyst systems and prospect of these methods will be introduced briefly in this article.

2.6.1. Photocatalytic

The photocatalytic reduction of CO2 is achieved by simulating photosynthesis of plants. The essence is that semiconductor photocatalysts generate photo-generated electron-hole pairs under light conditions, and promote the oxidation-reduction reaction between CO2 and H2O to generate hydrocarbons. Figure 5 is a schematic diagram of hydrocarbons produced by the photocatalytic reduction of CO2 [14]. In the process of photocatalytic reaction, when the energy of light radiation is higher than its own forbidden band width, the electrons will be excited to transition from the valence band to the conduction band, thereby forming holes having oxidizing ability and electrons having the reducing ability required for catalytic reaction. Subsequently, the electrons react with H2O to reduce CO2 to hydrocarbons such as methanol, methane, formaldehyde, and formic acid; while photo-generated holes with strong oxidation ability oxidize H2O to release O2. Catalysts are essential for photocatalytic processes. Here are three photocatalyst materials used commonly: semiconductor materials, graphene-based nanomaterials and MOFs-based heterogeneous catalysts.
Photocatalyst systems have been widely studied in three systems: semiconductor materials, graphene composites and metal-organic framework (MOF). Semiconductor materials include many types of inorganic binary compounds, such as TiO2, ZnO, CdS, SiC, etc. Among them, TiO2, as a typical semiconductor material, has received more research, due to its high stability and low cost [15,16]. However, the photo-response range of TiO2 is narrow, the utilization rate of light is extremely low, and photo-generated electron-holes are easily generated. In order to improve the conversion and selectivity of the photocatalytic process, a variety of TiO2 surface modification methods have been developed, including doping, metal deposition, alkali modification, heterojunction structure, and the loading of carbon-based materials [17,18,19,20].
Graphene or reduced graphite oxide (RGO) has attracted wide attention, due to its unique electronic properties, large theoretical specific surface area and stable chemical properties [21]. Hsu et al. [22] used graphene oxide to efficiently convert CO2 to methanol. The modified Hummer method was used to synthesize GO photocatalyst, which improved the catalytic activity of the catalyst. The modified graphene oxide converted CO2 to methanol under visible light irradiation, and the conversion rate was 0.172 μmol · gcatalyst−1·h−1, which was six times higher than that of pure TiO2. Tan et al. [23] successfully synthesized a new graphene oxide material, GO–OTiO2, with the highest photoactivity when the GO loading was 5%. After 6 h of reaction, the yield of CH4 was 1.718 μmol·gcatalyst−1·h−1, which has significantly improved light stability, and maintains 95.8% reactivity, even after six hours of light exposure. Liu et al. [24] prepared TiO2-RGO as a photocatalyst, and the yields of CH4 and CH3OH can reach 2.10 μmol·gcatalyst−1·h−1 and 2.20 μmol·gcatalyst−1·h−1, respectively. The rapid transfer from TiO2 to graphene inhibits photo-generated electron-hole recombination. Takayama et al. [25] prepared a graphene CuGaS2/RGO/TiO2 composite material. Since the RGO bridging layer provides an electron channel between CuGaS2 and TiO2, it has a high efficiency. The yield of CO is 0.15 μmol·gcatalyst−1·h−1.
MOF is a three-dimensional crystalline microporous material composed of metal oxygen clusters and organic connecting molecules. It has a large surface area and porosity, and has high flexibility in adjusting structure and composition. It is applied in the fields of adsorption, separation and gas storage [26,27]. Generally, the MOF is a structure formed by an organic linker, and the thermal stability of the material is weak, so it must be modified to achieve photochemical activity [28]. Wang et al. [29] reported doping Zr6O4(OH)4(para-biphenyldicarboxylic acid)6 frameworks (UIO-67MOF) with ligands containing Ir, Re and Ru, in which Re-MOF showed the activity of CO2 photocatalytic reduction. Subsequently, Fu et al. [30] prepared Ti and amine-functionalized Ti8O8(OH)4(benzene-1,4-dicarboxylate-NH2)6 frameworks (NH2-MIL-125 (Ti)). CO2 was reduced to HCOO under visible light irradiation, while MIL-125 (Ti) (MOF without amine functional group) showed only UV activity. In addition, amine functional groups increase the absorption of carbon dioxide, which in turn facilitates the conversion process. Sun et al. [31] prepared amine-functionalized UIO-66 (Zr) MOF. Compared with NH2-MIL-125 (Ti), the yield of HCOO in NH2-UIO-66 (Zr) increased by 1.6 times. Choi et al. [32] synthesized ReI (CO)3(BPYDC)Cl (BPYDC= 2,2′-bipyridine-5,5′-dicarboxylate) and combined it with UIO-67MOF. By systematically controlling its density in the framework, it was found that Re3-MOF had the highest photocatalytic activity. Based on the structure and activity of Re-MOFs, coating the surface of Ag nanocubes (Ag / Re3-MOF) with Re3-MOF to limit the photoactive Re center to the surface of Ag nanocubes can increase the CO2 conversion rate by a factor of seven, and has long-term stability under visible light for up to 48 hours.
For the photocatalytic reduction of CO2, the problems of photo-generated electron-hole recombination and low solar energy utilization still exist. Therefore, the overall conversion and selectivity of photocatalytic reduction of CO2 need to be improved. At present, a single catalyst system cannot solve these problems well. How to efficiently combine different types of catalysts and work together for the catalytic reduction of CO2 is still a future development direction in the field of photocatalysis.

2.6.2. Electrocatalysis

Electrocatalytic reduction is a technology that converts CO2 into carbon monoxide (CO), hydrocarbons, alcohols, esters, carboxylic acids and other new substances, through the cathode and anode in aqueous or non-aqueous solvents [33]. As shown in Table 3, according to the number of electrons transferred, the electrochemical reduction of CO2 can usually be divided into 2, 4, 6, or 8 electron reactions [34], and due to the small difference in thermodynamic potential energy between several possible reactions, the products obtained are usually mixtures of carbon compounds, such as CO, formic acid (HCOOH), methanol (CH3OH), formaldehyde (HCHO), methane (CH4), Ethylene (C2H4), etc. [35]
The CO2 reduction reaction (heterogeneous) is roughly divided into three steps: (1) CO2 molecules are adsorbed on the catalyst surface; (2) electron transfer/proton migration; (3) the target product is desorbed from the catalyst surface and diffused into the electrolyte. Therefore, it is generally believed that the mechanism of electrocatalytic reduction of CO2 in aqueous solution is:
CO2+e→ ·CO2ads
CO2ads+H2O→· HCO2ads+OH
HCO2ads+e→ HCO2
CO2+BH+e→HHCO2+B
In the formula, BH is a proton donor.
As shown in equations (1)–(4), CO2 is first adsorbed on the surface of the cathode catalyst to form an adsorbed free radical anion (·CO2ads). Under different applied voltages, ·CO2ads is protonated and subsequently, under the action of the catalyst, different reduction products are generated through a series of reduction processes (as shown in Figure 6). If the catalyst’s adsorption capacity for CO2 is weak, CO2ads will be desorbed and the final product is HCOOH; if the catalyst’s adsorption capacity is strong, the final product is more complex and CO, CH3OH and hydrocarbons may be generated [36,37].
The CO2 electroreduction reaction occurs in the gas phase, aqueous solution or non-aqueous solution. Electrocatalyst, electrode material, reaction medium, molecular weight, buffer strength, pH value, CO2 concentration, reaction temperature and pressure will affect the selectivity of the reaction product [38]. At present, CO2 electrocatalytic reduction catalysts can be divided into nano-metal catalysts, metal-organic frameworks (MOFs), covalent organic frameworks (COFs) and carbon nanomaterials. Nanometal catalysts with high specific surface area and active sites have attracted much attention, including nanowires, nanoparticles, nanotubes, and nanoporous membranes [39,40,41,42,43]. In order to improve the reactivity and selectivity of electroreduction, some scholars have studied alloy catalysts to improve the bonding strength of intermediates on the catalyst surface to improve the reaction kinetics, such as Au-Cu, Ag-Sn, Bi-Cu; the synergy of electronic and geometric effects of bimetals can make the product’s Faraday efficiency up to 96.4% [44]. Hollingsworth et al. used super-strong alkaline ionic liquid as an electrolyte and solvent for CO2 electroreduction, and found that a formic acid product with a Faraday efficiency of 93% was obtained on an Ag electrode, and the reaction potential was only 0.17V [45].
Metal-organic frameworks (MOFs) have attracted increasing attention from researchers, due to their high porosity, large specific surface area, adjustable pore size, and variable functional groups [46]. In addition to hybrid molecular material systems [47,48,49], some Cu and Zn-based MOFs can also be directly used as electrocatalysts for CO2 reduction. Hinogami et al. [50] used copper sulfate MOF, with the characteristics of electron conductivity, proton conductivity and uniformly dispersed reaction sites, as a catalyst for CO2 reduction. Interestingly, copper borate MOF produces only HCOOH products at 1.2V (vs. SHE) potential in aqueous electrolyte, while Cu metal electrodes produce a series of products. In addition, the yield of HCOOH is 13 times that of Cu metal electrodes. Kang et al. [51] deposited Zn-1,3,5-trimellitic acid metal organic framework (Zn-BTCMOF) on carbon paper, as an electrocatalyst for CO2 reduction in an ionic liquid-based electrolyte system. The morphology of Zn-MOF has a significant effect on the performance. The flaky Zn-MOF shows the highest catalytic activity, due to its large electroactive surface area. By comparing different ionic liquids, it was found that the interaction between fluorine-containing imidazolium ionic liquids and CO2 was the strongest.
Covalent organic frameworks (COFs) are compounds composed of light elements (H, C, N, O, B, etc.) connected by strong covalent bonds. They belong to a new type of nanoporous polymer composites. The characteristics of high porosity and orderly controllable structure can improve the reactivity and selectivity by changing the chemical structure, and show excellent performance in the adsorption and fixation of CO2 [52,53]. Lin et al. [54] used COFs to achieve the electrocatalytic reduction of CO2. Cobalt porphyrin units (TAP) and 1,4-benzaldehyde (BDA) were used as precursor units to successfully prepare two-dimensional COFs materials, which had a 90% conversion rate of CO2 to CO at 55V potential. Song et al. [55] used cylindrical mesoporous nitrogen-doped carbon nanotubes to directly convert CO2 to ethanol. At a potential of −0.56V, the Faraday efficiency of ethanol is 77%, and the formation of CO is completely suppressed. The selectivity of ethanol in the CO2 reduction product was as high as 100%.
Carbon nanomaterials have the advantages of high electrical conductivity, good stability, and low price. They are the most promising catalysts for replacing precious metals for electrochemical reduction. For example, Kumar et al. [56] successfully prepared carbon nanofibers through carbonized polyaniline. This carbon nanomaterial can achieve carbon dioxide reduction under an overpotential of 0.17V, and under the same conditions, it can also obtain a higher current density than Ag nanoparticles. In addition, research has found that heterogeneous atoms, such as nitrogen, boron, sulfur, and phosphorus can improve the catalytic activity and stability of carbon nanomaterials. Among them, nitrogen-doped nanocarbon materials have demonstrated catalytic performance comparable to that of Pt/C electrodes [57,58].
Although electrocatalytic reduction of CO2 has the advantages of a controllable reaction process, an electrolyte that can be recycled, and a compact reaction system, it also has problems such as relatively slow reaction kinetics and low energy utilization. How to reduce overpotential and improve product selectivity are the main problems facing the electrocatalytic reduction of CO2. Choosing a more suitable electrode and catalyst may be a breakthrough to solve these problems.

2.6.3. Plasma Catalysis

Plasma is considered to be a fourth material form which is different from solid-liquid-gas. When the temperature of other particles such as ions and gas molecules in the plasma is much lower than the electron temperature, it is called non-thermal plasma (NTP). NTP occurs in the presence of high-energy electrons. It is caused by the inelastic collision between high-energy electrons and CO2 gas molecules under the action of an electric field. The process is shown in Figure 7. The C=O bond dissociation energy of CO2 molecules is 5.5eV, and the electron energy in low temperature plasma is mostly between 1–20eV. Therefore, high energy electrons in low temperature plasma can dissociate CO2 molecules and convert them into CO and O2. Similarly, plasma can also perform other thermochemically difficult chemical reactions under mild conditions, providing new technical support for traditional chemical reactions [59].
Low-temperature plasma is a promising CO2 conversion technology, because it can activate and transform reactive molecules at room temperature, and pressure and reactions are rapid. However, the plasma alone participating in the reaction has low selectivity for the target product, and the conversion rate and energy efficiency need to be improved. The most effective improvement is to add a catalyst in the plasma [60]. The addition of Cu and Mn-supported catalysts to the dielectric barrier discharge plasma can increase the CO2 conversion rate from 6.7% to 36%, and increase the energy efficiency of the product CO by 116% [61]. Adding a catalyst can also improve the conversion rate and energy efficiency of the DRM reactants. In the process of CO2–H2 system conversion using microwave discharge plasma, the addition of Ni / TiO2 catalyst can increase the CO2 conversion rate from 14% to 28% [62].
Cryogenic plasma has many advantages not available in other new technologies, including sustainable electrical energy operation at room temperature, great flexibility for materials to be processed, efficient energy storage, low investment and operating costs, and strong applicability. However, the low-temperature plasma technology has a trade-off problem between CO2 conversion rate and energy efficiency, that is, it is difficult to achieve high CO2 conversion rate and high energy efficiency at the same time. Studies [63] have shown that this problem may be overcome by improving the plasma system, such as combining plasma discharge with a catalyst, but this requires further research. In addition, by modeling the plasma process, coupling it with other technologies, and further exploring the synergistic effects of plasma-bound catalysts, we can better understand the mechanism of the plasma conversion of CO2.

2.6.4. Thermal Catalysis

Converting CO2 into fuel or other value-added products is the ideal way to reduce the continuous emission of carbon dioxide into the atmosphere, and it may be more advantageous to convert to alternative fuels for industrial production or transportation [64]. At present, the fuel products converted by CO2 catalysis include methane, methanol, formic acid, dimethyl ether, and syngas, etc. [65,66,67,68] The transformation of CO2 to methane is an essential process for the fuel market. Methane is the main component of natural gas, which is easier to transport, handle, and store [69]. By 2030, the CO2 methanation market is expected to reach 4 to 65 billion cubic meters per year [70]. CO2 methanation is the most widely studied, and it is mostly conventional thermal catalytic conversion [71]. Therefore, this article mainly introduces the thermal catalytic conversion of CO2 to methane.
CO2 methanation refers to the strongly exothermic reaction process of H2 and CO2 to form CH4 and H2O under the action of 150–500 °C, 0.1–10 MPa, and a metal catalyst [72]. Table 4 lists the main reactions and side reactions performed during the CO2 methanation reaction. The main reactions in this gaseous catalytic process include reverse water gas shift, CO methanation, reverse dry reforming, and CO2 methanation. Due to the fact that only the reverse water gas shift reaction absorbs heat, and the other three reactions are highly exothermic, leading to a strong exotherm in the CO2 methanation process, in order to keep the CO2 methanation within the ideal temperature range, effective heat dissipation is essential. In addition to temperature problems, side reactions can also cause problems, such as scaling on the catalyst surface, plugging pores, and carbon deposition. In general, the selectivity of the CO2 methanation reaction depends on the catalyst used and its chemical equilibrium [73].
Through investigation of catalyst performance [74,75], the CO2 methanation activity of different metals is Ru > Rh > Ni > Fe > Co > Os > Pt > Ir > Mo > Pd > Ag > Au; selectivity is Pd > Pt > Ir > Ni > Rh > Co > Fe > Ru > Mo > Ag > Au. Among them, Ru not only exhibits higher low-temperature catalytic activity but also maintains stability during long-term use. Abe et al. [76] prepared a highly dispersed Ru nanoparticle Ru/TiO2 catalyst supported by TiO2 by ion sputtering, and the CH4 yield reached 100% at 160℃. The carbon dioxide turnover frequency (TOF) of Ru-based catalysts is closely related to the dispersion of active components and the type of support. The dispersion of active components depends on the interaction between Ru and the support. Under the same reaction conditions, the TOF order of Ru catalysts with different supports is Ru/Al2O3 > Ru/MgAl2O4 > Ru/MgO > Ru/C, and the activity of 15% (mass fraction) Ru/Al2O3 catalyst is 10 times higher than that of the Ni-based catalyst [77]. However, the price of metal Ru is high and cannot meet the needs of industrialized large-scale production [78].
In addition to precious metals, transition metals have gradually become an important component of the development of CO2 methanation catalysts, because of their low cost and availability. Fe-based catalysts were studied earlier and industrialized earlier. Still, the operating conditions are generally high temperature and pressure, which will easily cause the catalyst to deactivate carbon deposits and gradually be replaced by other active metals [79]. Co-based catalysts have relatively low methanation activity and strict requirements on the reaction environment. However, Pt has a better effect on the performance of Co-based catalysts. Co–Pt catalysts obtained by physical mixing have a carbon dioxide hydrogenation activity that is three times higher than single-metal Co catalysts at low temperatures. This is mainly because H2 and CO2 are activated on the surface of Pt particles and Co particles, respectively. The dissociated H is transferred to the surface of the Co particles through the SiO2 support, followed by a carbon dioxide hydrogenation reaction to generate CH4. As the distance between Pt and Co nanoparticles increases, the apparent activation energy decreases, obviously [80].
Compared with other transition metals, Ni-based catalysts have higher methanation catalytic activity, and the reaction conditions are relatively easy to control, which has been widely studied [81]. However, Ni-based catalysts have poor thermal conductivity and are prone to sintering and deactivation during the reaction. Therefore, research has mostly focused on improving the sintering resistance of the catalyst. For example, using MgO as a carrier and using the formed MgNiO2 structural stability to prevent sintering of Ni species [82]; loading the Ni species into carbon nanotubes, using the tube wall to restrict the growth and agglomeration of Ni species, and properly adding Ca ions, will further improve the dispersibility of Ni on carbon nanotubes [83]; strong thermal conductivity, SiC, as a carrier reduces heat accumulation and prevents the high-temperature sintering of Ni species [84]. At the same time, in order to overcome the shortcomings of single active component catalysts in the reaction process, there are also studies of dual active metal or even multiple active metal catalysts for CO2 methanation, such as the Ni-Ru/Al2O3 catalyst prepared by Zhen et al. [85] for the conversion of CO2 and CH4 selectivity, which reached 82.7% and 100%, respectively. Although there is much research on Ni-based catalysts, and industrial applications are most likely to be achieved in CO2 methanation, the current stability is poor. Therefore, it is still necessary to continue the research and development of catalysts to improve catalyst stability while ensuring catalyst activity.
Thermal catalysis has evolved from the addition of catalysts in traditional thermochemical methods and has been put into industrial applications. However, reaction conditions at high temperatures and pressures have limited its development. Therefore, typical thermal catalytic technologies such as CO2 methanation focus on the development of low-temperature catalysts on the one hand, and the other should pay more attention to the energy replacement and selection and development of high-stability and selective catalysts that meet high-temperature and high-pressure reaction conditions.

3. Materials and Methods

Data sources and processing ideas are shown in Figure 8. Based on the Web of Science SCI-Expanded database, a total of 4740 articles were found on the keywords "conversion of CO2" or "conversion of carbon dioxide" or "carbon dioxide conversion" or "CO2 conversion" or "utilization of CO2" or "utilization of carbon dioxide" or "carbon dioxide utilization" or “CO2 utilization”and "2010–2019" published. However, the original intention of Web of Science is not to provide accurate data for bibliometrics, so when searching in the form of "topic", all articles with keywords will appear even the keywords in KeyWords Plus only. Since KeyWords Plus is generated according to the title of the article citation, these articles are not closely related to the topic of the retrieval purpose [86], so these articles should be removed. This paper refers to the "front page" method proposed by Ho et al. [87], to eliminate the articles whose keywords only appear in the KeyWords Plus.
In addition, some articles, such as those on the effect of carbonic anhydrase in human body [88,89,90] and on carbon isotope conversion [91,92], may contain key words in the abstract, but they may not be related to the subject of this study, so they need to be excluded. At the same time, articles that contain retrieval errors such as “...CO2. Conversion... ” [93],”...carbon dioxide. Conversion... ” [94] should also be removed.
After the above screening, there are 4552 articles, including 4504 (98.94%) articles in English and 3704 (81.37%) articles in research. For the convenience of analysis, the 3670 (80.63%) English research articles are selected to analyze the basic processing of data in Microsoft Excel 2016. Then, the processed data is imported into Bibexcel for cooperative network and clustering processing, and finally, Gephi is used for data visualization processing.

4. Summary and Outlook

In the past decade, research on the conversion of CO2 into high value-added products has increased year by year, which is attributed to the increasing environmental problems, such as the greenhouse effect caused by CO2, and on the other hand, concerns about the shortage of energy resources. Therefore, China, the United States and other major countries have played a leading role in the utilization of CO2 resources, fully reflecting the responsibility and responsibility of the major powers. It must be pointed out that although China has a wide range of research in the field of CO2 conversion, the overall research level, especially the research on some key theoretical foundations, is weak, so the influence in this field needs to be improved. It is worth noting that, although China’s overall level of influence in this field is not outstanding, the Chinese Academy of Sciences is leading the way in many international research institutions, with its strong research team and rich research resources.
By screening and analyzing the keywords, we found that catalytic conversion technology is the most widely studied CO2 conversion method. Among them, photocatalysis, electrocatalysis, plasma catalysis, and thermal catalysis are the most frequently used methods to reduce CO2. However, all methods have problems such as high overpotential, low conversion efficiency, and poor product selectivity. Therefore, choosing the type of catalyst has become the top priority, and future research should focus on finding catalysts with multi-electron transport capabilities, improving the selectivity of the product; how to design reaction systems in different media is also the direction of future efforts.

Author Contributions

Y.X. provided funding. Q.W. and W.S. revised the manuscripts. X.W. and H.Z. provided software related help. Z.M. collected, analyzed and wrote the original draft. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2017YFC0210300), the National Science Foundation for Young Scientists of China (Grant No. 201707007), the National Science Foundation for Young Scientists of China (No.51774038), Beijing Science and Technology Project (No.Z191100009119008).

Acknowledgments

The authors would like to thank the editor and anonymous reviewers for their constructive feedback and helpful suggestions. In addition, we would like to acknowledge Yuhshan Ho, who taught us all the analytical skills required to conduct this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alain, G.; Miklos, C.; John-Paul, J.; Prakash, G.K.S.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products-closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar]
  2. Snoeckx, R.; Bogaerts, A. Plasma technology-a novel solution for CO2 conversion. Chem. Soc. Rev. 2017, 46, 5805–5863. [Google Scholar] [CrossRef] [Green Version]
  3. Olajire, A.A. Valorization of greenhouse carbon dioxide emissions into value-added products by catalytic processes. J. CO2 Util. 2013, 3–4, 74–92. [Google Scholar] [CrossRef]
  4. Ogura, K. Electrochemical reduction of carbon dioxide to ethylene: Mechanistic approach. J. CO2 Util. 2013, 1, 43–49. [Google Scholar] [CrossRef]
  5. Shi, J.; Jiang, Y.; Jiang, Z.; Wang, X.; Wang, X.; Zhang, S.; Han, P.; Yang, C. Enzymatic conversion of carbon dioxide. Chem. Soc. Rev. 2015, 44, 5981–6000. [Google Scholar] [CrossRef]
  6. Bogaerts, A.; Berthelot, A.; Heijkers, S.; Kolev, S.; Snoeckx, R.; Sun, S.; Trenchev, G.; Laer, K.V.; Wang, W. CO2 conversion by plasma technology: Insights from modeling the plasma chemistry and plasma reactor design. Plasma Sources Sci. Technol. 2017, 26, 063001. [Google Scholar] [CrossRef]
  7. Zabranska, J.; Pokorna, D. Bioconversion of carbon dioxide to methane using hydrogen and hydrogenotrophic methanogens. Biotechnol. Adv. 2017, 36, 707–720. [Google Scholar] [CrossRef]
  8. Blouin, M.; Sery, N.; Cluzeau, D.; Brun, J.J.; Bédécarrats, A. Balkanized Research in Ecological Engineering Revealed by a Bibliometric Analysis of Earthworms and Ecosystem Services. Environ. Manag. 2013, 52, 309–320. [Google Scholar] [CrossRef] [Green Version]
  9. Ortiz-Martínez, Y.; Galindo-Regino, C.; González-Hurtado, M.R.; Vanegas-Pastrana, J.J.; Valdes-Villegas, F. State of the art on Mycobacterium chimaera research: A bibliometric analysis. J. Hosp. Infect. 2017, e159–e160. [Google Scholar] [CrossRef]
  10. Baker, N.C.; Ekins, S.; Williams, A.J.; Tropsha, A. A bibliometric review of drug repurposing. Drug Discov. Today 2018, 23, 661–672. [Google Scholar] [CrossRef]
  11. Eugene, G. The history and meaning of the journal impact factor. JAMA 2006, 295, 90–93. [Google Scholar]
  12. Hirsch, J.E. An index to quantify an individual’s scientific research output. Proc. Natl. Acad. Sci. USA 2005, 102, 16569–16572. [Google Scholar] [CrossRef] [Green Version]
  13. López-Illescas, C.; de Moya Anegón, F.; Moed, H.F. Comparing bibliometric country-by-country rankings derived from the Web of Science and Scopus: The effect of poorly cited journals in oncology. J. Inf. Sci. 2009, 35, 244–256. [Google Scholar] [CrossRef]
  14. Xie, S.; Zhang, Q.; Liu, G.; Wang, Y. Photocatalytic and photoelectrocatalytic reduction of CO2 using heterogeneous catalysts with controlled nanostructures. Chem. Commun. 2016, 52, 35–59. [Google Scholar] [CrossRef]
  15. Wang, T.; Meng, X.; Li, P.; Ouyang, S.; Chang, K.; Liu, G.; Mei, Z.; Ye, J. Photoreduction of CO2 over the well-crystallized ordered mesoporous TiO2 with the confined space effect. Nano Energy 2014, 9, 50–60. [Google Scholar] [CrossRef]
  16. Xu, Q.; Yu, J.; Zhang, J.; Zhang, J.; Liu, G. Cubic anatase TiO2 nanocrystals with enhanced photocatalytic CO2 reduction activity. Chem. Commun. 2015, 51, 7950–7953. [Google Scholar] [CrossRef]
  17. Low, J.; Bei, C.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2016, 392, 658–686. [Google Scholar] [CrossRef]
  18. Liu, G.; Hoivik, N.; Wang, K.; Jakobsen, H. Engineering TiO2 nanomaterials for CO2 conversion/solar fuels. Sol. Energy Mat. Sol. C. 2012, 105, 53–68. [Google Scholar] [CrossRef]
  19. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  20. Ong, W.; Tan, L.; Chai, S.; Yong, S.; Mohamed, A.R. Highly reactive {001} facets of TiO2-based composites: Synthesis, formation mechanism and characterization. Nanoscale 2014, 6, 1946–2008. [Google Scholar] [CrossRef]
  21. Chae, H.K.; Siberio-Perez, D.Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A.J.; O’Keeffe, M.; Yaghi, O.M. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427, 523–527. [Google Scholar] [CrossRef] [PubMed]
  22. Hsu, H.; Shown, I.; Wei, H.; Chang, Y.; Du, H.; Lin, Y.; Tseng, C.; Wang, C.; Chen, L.; Lin, Y. Graphene oxide as a promising photocatalyst for CO2 to methanol conversion. Nanoscale 2013, 5, 262–268. [Google Scholar] [CrossRef] [PubMed]
  23. Tan, L.; Ong, W.; Chai, S.; Goh, B.T.; Mohamed, A.R. Visible-light-active oxygen-rich TiO2 decorated 2D graphene oxide with enhanced photocatalytic activity toward carbon dioxide reduction. Appl. Catal. B Environ. 2015, 179, 160–170. [Google Scholar] [CrossRef]
  24. Liu, J.; Niu, Y.; He, X.; Qi, J.; Li, X. Photocatalytic reduction of CO2 using TiO2-graphene nanocomposites. J. Nanomater. 2016, 6012896. [Google Scholar]
  25. Takayama, T.; Nakanishi, H.; Matsui, M.; Iwase, A.; Kudo, A. Photocatalytic CO2 reduction using water as an electron donor over Ag-loaded metal oxide photocatalysts consisting of several polyhedra of Ti4+, Zr4+, and Ta5+. J. Photochem. Photobiol. A Chem. 2018, 358, 416–421. [Google Scholar] [CrossRef]
  26. Zhao, Z.; Ma, X.; Kasik, A.; Li, Z.; Lin, Y.S. Gas separation properties of metal organic framework (MOF-5) membranes. Ind. Eng. Chem. Res. 2013, 52, 1102–1108. [Google Scholar] [CrossRef]
  27. Li, Y.; Yang, R.T. Gas adsorption and storage in metal− organic framework MOF-177. Langmuir 2007, 23, 12937–12944. [Google Scholar] [CrossRef] [PubMed]
  28. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.; Long, J.R. Carbon dioxide capture in metal–organic frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, C.; Xie, Z.; DeKrafft, K.E.; Lin, W. Doping metal–organic frameworks for water oxidation, carbon dioxide reduction, and organic photocatalysis. J. Am. Chem. Soc. 2011, 133, 13445–13454. [Google Scholar] [CrossRef]
  30. Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium metal–organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem. Int. Ed. 2012, 51, 3364–3367. [Google Scholar] [CrossRef]
  31. Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Studies on Photocatalytic CO2 Reduction over NH2-Uio-66 (Zr) and Its Derivatives: Towards a Better Understanding of Photocatalysis on Metal–Organic Frameworks. Chem. A Eur. J. 2013, 19, 14279–14285. [Google Scholar] [CrossRef] [PubMed]
  32. Choi, K.M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C.A.; Barmanbek, J.T.D.; Alshammari, A.S.; Yang, P.; Yaghi, O.M. Plasmon-enhanced photocatalytic CO2 conversion within metal–organic frameworks under visible light. J. Am. Chem. Soc. 2017, 139, 356–362. [Google Scholar] [CrossRef] [PubMed]
  33. O’Toole, T.R.; Margerum, L.D.; Westmoreland, T.D.; Vining, W.J.; Murray, R.W.; Meyer, T.J. Electrocatalytic reduction of CO2 at a chemically modified electrode. J. Chem. Soc. Chem. Commun. 1985, 1416–1417. [Google Scholar] [CrossRef]
  34. Lim, R.J.; Xie, M.; Sk, M.A.; Lee, J.; Fisher, A.; Wang, X.; Lim, K.H. A review on the electrochemical reduction of CO2 in fuel cells, metal electrodes and molecular catalysts. Catal. Today 2014, 233, 169–180. [Google Scholar] [CrossRef]
  35. Karamad, M.; Hansen, H.A.; Rossmeisl, J.; Nørskov, J.K. Mechanistic pathway in the electrochemical reduction of CO2 on RuO2. ACS Catal. 2015, 5, 4075–4081. [Google Scholar] [CrossRef]
  36. Chaplin, R.; Wragg, A.A. Effects of process conditions and electrode material on reaction pathways for carbon dioxide electroreduction with particular reference to formate formation. J. Appl. Electrochem. 2003, 33, 1107–1123. [Google Scholar] [CrossRef]
  37. Javier, A.; Chmielowiec, B.; Sanabria-Chinchilla, J.; Kim, Y.; Baricuatro, J.H.; Soriaga, M.P. A DEMS study of the reduction of CO2, CO, and HCHO pre-adsorbed on Cu electrodes: Empirical inferences on the CO2 RR mechanism. Electrocatal. Us 2015, 6, 127–131. [Google Scholar] [CrossRef]
  38. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef]
  39. Lu, Q.; Rosen, J.; Jiao, F. Nanostructured metallic electrocatalysts for carbon dioxide reduction. ChemCatChem 2015, 7, 38–47. [Google Scholar] [CrossRef]
  40. Lu, Q.; Rosen, J.; Zhou, Y.; Hutchings, G.S.; Kimmel, Y.C.; Chen, J.G.; Jiao, F. A selective and efficient electrocatalyst for carbon dioxide reduction. Nat. Commun. 2014, 5, 3342. [Google Scholar] [CrossRef]
  41. Rosen, J.; Hutchings, G.S.; Lu, Q.; Rivera, S.; Zhou, Y.; Vlachos, D.G.; Jiao, F. Mechanistic insights into the electrochemical reduction of CO2 to CO on nanostructured Ag surfaces. ACS Catal. 2015, 5, 4293–4299. [Google Scholar] [CrossRef]
  42. Back, S.; Yeom, M.S.; Jung, Y. Active sites of Au and Ag nanoparticle catalysts for CO2 electroreduction to CO. ACS Catal. 2015, 5, 5089–5096. [Google Scholar] [CrossRef] [Green Version]
  43. Baturina, O.A.; Lu, Q.; Padilla, M.A.; Xin, L.; Li, W.; Serov, A.; Artyushkova, K.; Atanassov, P.; Xu, F.; Epshteyn, A. CO2 electroreduction to hydrocarbons on carbon-supported Cu nanoparticles. ACS Catal. 2014, 4, 3682–3695. [Google Scholar] [CrossRef]
  44. Zhong, H.; Qiu, Y.; Zhang, T.; Li, X.; Zhang, H.; Chen, X. Bismuth nanodendrites as a high performance electrocatalyst for selective conversion of CO2 to formate. J. Mater. Chem. A 2016, 4, 13746–13753. [Google Scholar] [CrossRef]
  45. Hollingsworth, N.; Taylor, S.R.; Galante, M.T.; Jacquemin, J.; Longo, C.; Holt, K.B.; de Leeuw, N.H.; Hardacre, C. Reduction of carbon dioxide to formate at low overpotential using a superbase ionic liquid. Angew. Chem. Int. Ed. 2015, 54, 14164–14168. [Google Scholar] [CrossRef] [Green Version]
  46. Zhang, H.; Li, J.; Tan, Q.; Lu, L.; Wang, Z.; Wu, G. Metal–organic frameworks and their derived materials as electrocatalysts and photocatalysts for CO2 reduction: Progress, challenges, and perspectives. Chem. A Eur. J. 2018, 24, 18137–18157. [Google Scholar] [CrossRef]
  47. Shen, J.; Kortlever, R.; Kas, R.; Birdja, Y.Y.; Diaz-Morales, O.; Kwon, Y.; Ledezma-Yanez, I.; Schouten, K.J.P.; Mul, G.; Koper, M.T. Electrocatalytic reduction of carbon dioxide to carbon monoxide and methane at an immobilized cobalt protoporphyrin. Nat. Commun. 2015, 6, 8177. [Google Scholar] [CrossRef]
  48. Hod, I.; Farha, O.K.; Hupp, J.T. Electrocatalysis: Powered by porphyrin packing. Nat. Mater. 2015, 14, 1192–1193. [Google Scholar] [CrossRef]
  49. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal–organic frameworks for electrocatalytic reduction of carbon dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hinogami, R.; Yotsuhashi, S.; Deguchi, M.; Zenitani, Y.; Hashiba, H.; Yamada, Y. Electrochemical reduction of carbon dioxide using a copper rubeanate metal organic framework. ECS Electrochem. Lett. 2012, 1, H17–H19. [Google Scholar] [CrossRef]
  51. Kang, X.; Zhu, Q.; Sun, X.; Hu, J.; Zhang, J.; Liu, Z.; Han, B. Highly efficient electrochemical reduction of CO2 to CH4 in an ionic liquid using a metal-organic framework cathode. Chem. Sci. 2016, 7, 266–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Zhi, Y.; Shao, P.; Feng, X.; Xia, H.; Zhang, Y.; Shi, Z.; Mu, Y.; Liu, X. Covalent organic frameworks: Efficient, metal-free, heterogeneous organocatalysts for chemical fixation of CO2 under mild conditions. J. Mater. Chem. A 2018, 6, 374–382. [Google Scholar] [CrossRef]
  53. Buyukcakir, O.; Je, S.H.; Talapaneni, S.N.; Kim, D.; Coskun, A. Charged covalent triazine frameworks for CO2 capture and conversion. ACS Appl. Mater. Interface 2017, 9, 7209–7216. [Google Scholar] [CrossRef] [PubMed]
  54. Lin, S.; Diercks, C.S.; Zhang, Y.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.M. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Song, Y.; Chen, W.; Zhao, C.; Li, S.; Wei, W.; Sun, Y. Metal-free nitrogen-doped mesoporous carbon for electroreduction of CO2 to ethanol. Angew. Chem. Int. Ed. 2017, 56, 10840–10844. [Google Scholar] [CrossRef] [PubMed]
  56. Kumar, B.; Asadi, M.; Pisasale, D.; Sinha-Ray, S.; Rosen, B.A.; Haasch, R.; Abiade, J.; Yarin, A.L.; Salehi-Khojin, A. Renewable and metal-free carbon nanofibre catalysts for carbon dioxide reduction. Nat. Commun. 2013, 4, 2819. [Google Scholar] [CrossRef]
  57. Zhao, L.; He, R.; Rim, K.T.; Schiros, T.; Kim, K.S.; Zhou, H.; Gutiérrez, C.; Chockalingam, S.P.; Arguello, C.J.; Pálová, L. Visualizing individual nitrogen dopants in monolayer graphene. Science 2011, 333, 999–1003. [Google Scholar] [CrossRef] [Green Version]
  58. Cui, X.; Yang, S.; Yan, X.; Leng, J.; Shuang, S.; Ajayan, P.M.; Zhang, Z. Pyridinic-Nitrogen-Dominated Graphene Aerogels with Fe–N–C Coordination for Highly Efficient Oxygen Reduction Reaction. Adv. Funct. Mater. 2016, 26, 5708–5717. [Google Scholar] [CrossRef]
  59. Qin, Y.; Niu, G.; Wang, X.; Luo, D.; Duan, Y. Status of CO2 conversion using microwave plasma. J. CO2 Util. 2018, 28, 283–291. [Google Scholar] [CrossRef]
  60. Chen, G.; Wang, L.; Godfroid, T.; Snyders, R. Progress in Plasma-Assisted Catalysis for Carbon Dioxide Reduction. In Plasma Chemistry and Gas Conversion; IntechOpen: London, UK, 2018. [Google Scholar]
  61. Zeng, Y.; Tu, X. Plasma-catalytic CO2 hydrogenation at low temperatures. IEEE Trans. Plasma Sci. 2015, 44, 405–411. [Google Scholar] [CrossRef] [Green Version]
  62. Chen, G.; Britun, N.; Godfroid, T.; Georgieva, V.; Snyders, R.; Delplancke-Ogletree, M. An overview of CO2 conversion in a microwave discharge: The role of plasma-catalysis. J. Phys. D Appl. Phys. 2017, 50, 84001. [Google Scholar] [CrossRef]
  63. Ashford, B.; Tu, X. Non-thermal plasma technology for the conversion of CO2. Curr. Opin. Green Sustain. Chem. 2017, 3, 45–49. [Google Scholar] [CrossRef] [Green Version]
  64. Hu, B.; Guild, C.; Suib, S.L. Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products. J. Co2 Util. 2013, 1, 18–27. [Google Scholar] [CrossRef]
  65. Leonzio, G. State of art and perspectives about the production of methanol, dimethyl ether and syngas by carbon dioxide hydrogenation. J. Co2 Util. 2018, 27, 326–354. [Google Scholar] [CrossRef]
  66. Álvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the greener production of formates/formic acid, methanol, and DME by heterogeneously catalyzed CO2 hydrogenation processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef] [PubMed]
  67. Catizzone, E.; Bonura, G.; Migliori, M.; Frusteri, F.; Giordano, G. CO2 recycling to dimethyl ether: State-of-the-art and perspectives. Molecules 2018, 23, 31. [Google Scholar] [CrossRef] [Green Version]
  68. Stangeland, K.; Li, H.; Yu, Z. Thermodynamic analysis of chemical and phase equilibria in CO2 hydrogenation to methanol, dimethyl ether, and higher alcohols. Ind. Eng. Chem. Res. 2018, 57, 4081–4094. [Google Scholar] [CrossRef]
  69. Shima, A.; Sakurai, M.; Sone, Y.; Ohnishi, M.; Abe, T. Development of a CO2 reduction catalyst for the Sabatier reaction. In Proceedings of the 42nd International Conference on Environmental Systems, San Diego, CA, USA, 15–19 July 2012; p. 3552. [Google Scholar]
  70. Global, C.O. Initiative Global Roadmap for Implementing CO2 Utilization; University of Michigan: Ann Arbor, MI, USA, 2016. [Google Scholar]
  71. Galadima, A.; Muraza, O. Catalytic thermal conversion of CO2 into fuels: Perspective and challenges. Renew. Sustain. Energy Rev. 2019, 115, 109333. [Google Scholar] [CrossRef]
  72. Chang, F.; Kuo, M.; Tsay, M.; Hsieh, M. Hydrogenation of CO2 over nickel catalysts on rice husk ash-alumina prepared by incipient wetness impregnation. Appl. Catal. A Gen. 2003, 247, 309–320. [Google Scholar] [CrossRef]
  73. Ghaib, K.; Nitz, K.; Ben Fares, F.Z. Chemical methanation of CO2: A review. ChemBioEng Rev. 2016, 3, 266–275. [Google Scholar] [CrossRef]
  74. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation—From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  75. Vannice, M.A. The catalytic synthesis of hydrocarbons from carbon monoxide and hydrogen. Catal. Rev. Sci. Eng. 1976, 14, 153–191. [Google Scholar] [CrossRef]
  76. Abe, T.; Tanizawa, M.; Watanabe, K.; Taguchi, A. CO2 methanation property of Ru nanoparticle-loaded TiO2 prepared by a polygonal barrel-sputtering method. Energy Environ. Sci. 2009, 2, 315–321. [Google Scholar] [CrossRef]
  77. Kowalczyk, Z.; Stołecki, K.; Rarog-Pilecka, W.; Miśkiewicz, E.; Wilczkowska, E.; Karpiński, Z. Supported ruthenium catalysts for selective methanation of carbon oxides at very low COx/H2 ratios. Appl. Catal. A Gen. 2008, 342, 35–39. [Google Scholar] [CrossRef]
  78. Tada, S.; Ochieng, O.J.; Kikuchi, R.; Haneda, T.; Kameyama, H. Promotion of CO2 methanation activity and CH4 selectivity at low temperatures over Ru/CeO2/Al2O3 catalysts. Int. J. Hydrogen Energy 2014, 39, 10090–10100. [Google Scholar] [CrossRef]
  79. Gogate, M.R.; Davis, R.J. Comparative study of CO and CO2 hydrogenation over supported Rh–Fe catalysts. Catal. Commun. 2010, 11, 901–906. [Google Scholar] [CrossRef]
  80. Beaumont, S.K.; Alayoglu, S.; Specht, C.; Kruse, N.; Somorjai, G.A. A nanoscale demonstration of hydrogen atom spillover and surface diffusion across silica using the kinetics of CO2 methanation catalyzed on spatially separate Pt and Co nanoparticles. Nano Lett. 2014, 14, 4792–4796. [Google Scholar] [CrossRef] [Green Version]
  81. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported catalysts for CO2 methanation: A review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  82. Takezawa, N.; Terunuma, H.; Shimokawabe, M.; Kobayashib, H. Methanation of carbon dioxide: Preparation of Ni/MgO catalysts and their performance. Appl. Catal. 1986, 23, 291–298. [Google Scholar] [CrossRef]
  83. Hu, X.F.; Yang, W.; Wang, N.; Luo, S.Z.; Chu, W. Catalytic Properties of Ni/CNTs and Ca-Promoted Ni/CNTs for Methanation Reaction of Carbon Dioxide. Advanced Materials Research 2014, 924, 217–226. [Google Scholar] [CrossRef]
  84. Le, T.A.; Kang, J.K.; Park, E.D. CO and CO2 methanation over Ni/SiC and Ni/SiO2 catalysts. Top. Catal. 2018, 61, 1537–1544. [Google Scholar] [CrossRef]
  85. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni-Ru/γ-Al2O3 via modulating impregnation sequence and controlling surface active species. RSC Adv. 2014, 4, 16472–16479. [Google Scholar] [CrossRef]
  86. Garfield, E. KeyWords Plus-ISI’s breakthrough retrieval method. 1. Expanding your searching power on current-contents on diskette. Curr. Contents 1990, 32, 5–9. [Google Scholar]
  87. Fu, H.; Wang, M.; Ho, Y. The most frequently cited adsorption research articles in the Science Citation Index (Expanded). J. Colloid Interface Sci. 2012, 379, 148–156. [Google Scholar] [CrossRef]
  88. Avvaru, B.S.; Kim, C.U.; Sippel, K.H.; Gruner, S.M.; Agbandjemckenna, M.; Silverman, D.N.; Mckenna, R. A short, strong hydrogen bond in the active site of human carbonic anhydrase II. Biochem. Us 2010, 49, 249–251. [Google Scholar] [CrossRef] [Green Version]
  89. Dayne, W.; Chae Un, K.; Chingkuang, T.; Robbins, A.H.; Gruner, S.M.; Silverman, D.N.; Robert, M.K. Structural and kinetic effects on changes in the CO(2) binding pocket of human carbonic anhydrase II. Biochem. Us 2012, 51, 9156–9163. [Google Scholar]
  90. Adeva-Andany, M.M.; Fernández-Fernández, C.; Sánchez-Bello, R.; Donapetry-García, C.; Martínez-Rodríguez, J. The role of carbonic anhydrase in the pathogenesis of vascular calcification in humans. Atherosclerosis 2015, 241, 183–191. [Google Scholar] [CrossRef]
  91. Nordeman, P.; Friis, S.D.; Andersen, T.L.; Audrain, H.; Larhed, M.; Skrydstrup, T.; Antoni, G. Chemical Conversion of (CO2)-C-11 to (CO)-C-11 via Silacarboxylic Acids: Applications in Palladium-Mediated Carbonylations. J. Label. Compd. Radiopharm. 2015, 58, S383. [Google Scholar]
  92. Patrik, N.; Friis, S.D.; Andersen, T.L.; Hélène, A.; Mats, L.; Troels, S.; Gunnar, A. Rapid and Efficient Conversion of (11) CO2 to (11) CO through Silacarboxylic Acids: Applications in Pd-Mediated Carbonylations. Chemistry 2016, 21, 17601–17604. [Google Scholar]
  93. Rivas, B.D.; López-Fonseca, R.; Jiménez-González, C.; Gutiérrez-Ortiz, J.I. Highly active behaviour of nanocrystalline Co3O4 from oxalate nanorods in the oxidation of chlorinated short chain alkanes. Chem. Eng. J. 2012, 184, 184–192. [Google Scholar] [CrossRef]
  94. Elmore, C.S. The synthesis and use of [14C] carbon monoxide in Pd-catalyzed carbonylation reactions. J. Label. Compd. Radiopharm. 2011, 54, 59–64. [Google Scholar] [CrossRef]
Figure 1. Annual number of publications from 2010 to 2019.
Figure 1. Annual number of publications from 2010 to 2019.
Catalysts 10 00370 g001
Figure 2. The performance of the top 10 representative countries (TP > 103).
Figure 2. The performance of the top 10 representative countries (TP > 103).
Catalysts 10 00370 g002
Figure 3. The collaboration network of 75 institutions (TP > 10).
Figure 3. The collaboration network of 75 institutions (TP > 10).
Catalysts 10 00370 g003
Figure 4. Classification map of the top 108 keywords (Frequency > 10).
Figure 4. Classification map of the top 108 keywords (Frequency > 10).
Catalysts 10 00370 g004
Figure 5. Schematic illustration of the basic mechanism of photocatalytic reduction of CO2 with H2O on a semiconductor photocatalyst. Adapted and reprinted with permissions from [14], Copyright (2016) Royal Society of Chemistry (RSC).
Figure 5. Schematic illustration of the basic mechanism of photocatalytic reduction of CO2 with H2O on a semiconductor photocatalyst. Adapted and reprinted with permissions from [14], Copyright (2016) Royal Society of Chemistry (RSC).
Catalysts 10 00370 g005
Figure 6. Electrocatalytic reduction of CO2 in aqueous solution.
Figure 6. Electrocatalytic reduction of CO2 in aqueous solution.
Catalysts 10 00370 g006
Figure 7. Basic process of CO2 decomposition in low-temperature plasma (M = CO2, CO or O2). Adapted and reprinted with permissions from [59], Copyright 2018, Elsevier.
Figure 7. Basic process of CO2 decomposition in low-temperature plasma (M = CO2, CO or O2). Adapted and reprinted with permissions from [59], Copyright 2018, Elsevier.
Catalysts 10 00370 g007
Figure 8. The general flow diagram for systematic bibliometric analyses.
Figure 8. The general flow diagram for systematic bibliometric analyses.
Catalysts 10 00370 g008
Table 1. The performance of the 10 most popular journals.
Table 1. The performance of the 10 most popular journals.
JournalTPIFh-IndexPercentage
Journal of CO2 Utilization2145.189335.83%
Applied Catalysis B: Environmental14814.2292054.03%
ACS Catalysis11212.2211393.05%
International Journal of Hydrogen Energy1034.0841872.81%
Journal of Materials Chemistry A8510.7331522.32%
Chemsuschem837.8041302.26%
Angewandte Chemie: International Edition7912.2574822.15%
Catalysis Science & Technology725.726861.96%
Chemical Engineering Journal628.3551721.69%
RSC Advances613.0491131.66%
Note: TP = the number of publications, IF = impact factor (published by the JCR in 2019).
Table 2. Top ten institutions with a large number of documents (TP > 39).
Table 2. Top ten institutions with a large number of documents (TP > 39).
InstitutionTPh-IndexCountryPercentage
Chinese Academy of Sciences28335China3.91%
University of Chinese Academy of Sciences11818China1.63%
Tianjin University7129China0.98%
Dalian University of Technology6916China0.95%
Korea Advanced Institute of Science and Technology5817South Korea0.80%
Nankai University5517China0.76%
Korea University5313South Korea0.73%
University of Antwerp5120Belgium0.70%
University of Science and Technology of China4517China0.62%
Korea Institute of Science and Technology3915South Korea0.54%
Table 3. CO2 electrocatalytic reduction reaction and corresponding electrode potential.
Table 3. CO2 electrocatalytic reduction reaction and corresponding electrode potential.
Electroreduction Reaction EquationElectrode Potential/V(vs.NHE)
CO2 + 2H+ + 2e→ CO + H2O−0.52
CO2 + 2H++ 2e→ HCOOH−0.61
CO2 + 4H++ 4e→ HCHO + H2O−0.51
CO2 + 6H++ 6e→ CH3OH + H2O−0.38
CO2 + 8H+ + 8e→ CH4 + 2H2O−0.24
2CO2 + 12H+ + 12e→ C2H2 + 4H2O−0.34
Table 4. Main reactions and side reactions that occur during CO2 methanation.
Table 4. Main reactions and side reactions that occur during CO2 methanation.
Reaction Equation Δ H R 298 K /kJ·mol-1Response Type
CO2 + H2 ⇌CO + H2O41reverse water gas shift
CO + 3H2 ⇌CH4 + H2O−206CO methanation
2CO + 2H2 ⇌CH4 + CO2−247reverse dry reforming
CO2 + 4H2 ⇌CH4 + 2H2O−165CO2 methanation
2CO ⇌C + CO2−172Boudouard reaction
CO + H2 ⇌C + H2O−131CO reduction
CO2 + 2H2 ⇌C + 2H2O−90CO2 reduction
CH4 ⇌ C + 2H2−75Methane pyrolysis

Share and Cite

MDPI and ACS Style

Xing, Y.; Ma, Z.; Su, W.; Wang, Q.; Wang, X.; Zhang, H. Analysis of Research Status of CO2 Conversion Technology Based on Bibliometrics. Catalysts 2020, 10, 370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10040370

AMA Style

Xing Y, Ma Z, Su W, Wang Q, Wang X, Zhang H. Analysis of Research Status of CO2 Conversion Technology Based on Bibliometrics. Catalysts. 2020; 10(4):370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10040370

Chicago/Turabian Style

Xing, Yi, Zhiliang Ma, Wei Su, Qunhui Wang, Xiaona Wang, and Hui Zhang. 2020. "Analysis of Research Status of CO2 Conversion Technology Based on Bibliometrics" Catalysts 10, no. 4: 370. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10040370

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