Next Article in Journal
Catalytic Technologies for CO Hydrogenation for the Production of Light Hydrocarbons and Middle Distillates
Previous Article in Journal
Rigorous Model-Based Design and Experimental Verification of Enzyme-Catalyzed Carboligation under Enzyme Inactivation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrochemical CO2 Reduction to CO Catalyzed by 2D Nanostructures

1
Department of Energy Science & Engineering, DGIST, 333 Techno Jungang-daero, Hyeonpung-eup, Dalseong-gun, Daegu 42988, Korea
2
Chemical and Biological Engineering Department, Koç University, 34450 Sarıyer, Turkey
3
Linde + Robinson Laboratories, California Institute of Technology, Pasadena, CA 91125, USA
*
Author to whom correspondence should be addressed.
Submission received: 7 December 2019 / Revised: 3 January 2020 / Accepted: 5 January 2020 / Published: 9 January 2020
(This article belongs to the Section Electrocatalysis)

Abstract

:
Electrochemical CO2 reduction towards value-added chemical feedstocks has been extensively studied in recent years to resolve the energy and environmental problems. The practical application of electrochemical CO2 reduction technology requires a cost-effective, highly efficient, and robust catalyst. To date, vigorous research have been carried out to increase the proficiency of electrocatalysts. In recent years, two-dimensional (2D) graphene and transition metal chalcogenides (TMCs) have displayed excellent activity towards CO2 reduction. This review focuses on the recent progress of 2D graphene and TMCs for selective electrochemical CO2 reduction into CO.

Graphical Abstract

1. Introduction

The increasing utilization of fossil fuels leads to huge amounts of CO2 emissions which causes inevitable damage to the planet’s ecosystem [1,2,3,4,5]. One possible way to counter this is to capture and convert CO2 to industrially important organic compounds [6,7,8,9,10]. To date, several strategies have been employed for the reduction of CO2; for example, chemical [11], thermochemical [12,13], photocatalytic [14,15,16,17,18,19,20,21], electrocatalytic [22,23,24], biological [25], and inorganic transformation [26]. However, among those, a great deal of research has focused on electrochemical CO2 reduction technology aimed towards large scale applications due to its environmental compatibility and cost-effectiveness [27,28,29].
However, the low reactivity of CO2 requires high energy input for its activation [30,31,32,33,34]. CO2 can be reduced to various useful products by employing suitable cell potential in combination with appropriate catalysts (as shown in Table 1 (Equations (1)–(7)) [35,36]. For example, CO2 reduction to C1 products such as CO is thermodynamically more favorable because it requires only two electrons and two protons for the activation, and such C1 products are useful as building blocks for the production of higher hydrocarbons [37]. In contrast, long-chain hydrocarbons are industrially significant but CO2 reduction to higher hydrocarbons is a complex process, since C–C bond formation demands high energy input [38]. Hence, rather than CO2 reduction to higher hydrocarbons, the product selectivity and efficiency are crucial factors to be addressed.
However, developing a low cost, highly efficient, and stable catalyst remains an obstacle to be overcome. To date a number of catalysts including metals [39], alloys [40], metal oxides [41,42], metal complexes [43], carbides [44,45], and metal-organic frameworks [46,47,48] have been studied for electrocatalytic CO2 reduction. Of these, transition metal-based catalysts (e.g., Au, Pt, Ni, and Ag) displayed notable activity, but their scarcity in nature and high-cost limit them from large-scale employment [49]. Metal oxides, namely, TiO2, RuO2, and IrO2, have been widely investigated [50,51,52], but they also possess notable disadvantages—low efficiency, high cost, poor stability, and low current exchange density. Recently 2D materials, in particular, graphene derivatives and transition metal chalcogenides (TMCs), have emerged as active and inexpensive alternatives to traditional electrocatalysts [52]. They possess unique physicochemical properties and their large surface areas facilitate the redox reaction process to boost the catalytic activity [53,54]. Moreover, exposed active sites of 2D nanostructures are favorable for interfacial charge transfer [55]. Enhancement in mass transport and electron conductivity of 2D electrocatalysts can be achieved by co-catalyst doping [56]. For instance, doping of Nb into MoS2 increases the overall catalytic performance at the minimal potential of 50–150 mV [57]. Therefore, all these collective advantages of 2D nanostructures make them suitable for electrochemical CO2 reduction.
This short review will describe the recent progress and advances in emerging 2D materials in electrochemical CO2 reduction towards CO production. We will be discussing the fundamentals and pathways of electrochemical CO2 reduction. Various advanced synthesis methodologies of 2D catalysts and the effect of metal/and non-metal doping on catalytic activity will also be discussed along with the theoretical studies.

2. Fundamentals of Electrochemical CO2 Reduction

CO2 is a linear molecule and a large input of energy is needed to break the carbon and oxygen bonding (i.e., C=O) [58,59]. H2O reduction to H2 and ½O2 demands 237.2 kJ mol−1 of energy under standard conditions. Whereas, CO2 reduction to hydrocarbon fuels CH4 and CH3OH requires much higher free energy towards positive change, equivalent to 818.3 kJ mol−1 and 702 kJ mol−1, respectively [59]. Therefore, CO2 reduction is an uphill reaction and considered as a grand challenge [60]. The electrochemical cell used for CO2 reduction reaction consists of a cathode, an anode, and an ion-exchange membrane placed in-between them, as demonstrated in Figure 1. The process of electrochemical CO2 reduction begins with diffusion of CO2 molecule from the solution to cathode surface which is coated with a suitable catalyst. One of the postulated mechanisms for activating CO2 molecule is a single electron transfer to CO2, thereby generating a CO2•− radical anion [61]. A single-electron addition bends the linear molecule because of the repulsion between acquired electron over electrophilic carbon atom and free electron pairs on the oxygen atom. However, this mechanism has a setback, because for such a process to take place, a very high redox potential is required (i.e., −1.9 V versus NHE) [62]. But then, as seen in Table 1, which depicts the standard redox potential for CO2 reduction reactions, the possibility of single-electron transfer to a CO2 molecule becomes inconceivable [63]. In terms of thermodynamics, the edges of conduction band and valence band of a semiconductor must be above and below the standard reduction and oxidation potentials, respectively. Therefore, this makes it quite clear why a single electron transfer to the CO2 molecule is highly improbable. Hence, a better efficiency can be achieved with a multiple proton-coupled electron transfer (PCET) process. This process involves the transfer of protons with electrons, thereby bypassing the formation of high energy CO2 radical [64]. Even though the PCET phenomenon makes CO2 reduction feasible, the process is dependent upon the concentration of available electrons and protons. For example, as manifested from Equations (2)–(7), which show the main products obtained during CO2 reduction in the presence of H2O, by the PCET phenomenon (pH = 7) [65], generation of formic acid and carbon monoxide is more favorable than generating methane. The reason behind such a phenomenon is that the carbon monoxide requires two electrons and protons, whereas eight electrons and protons are necessary for producing methane. Another possible pathway involved in CO2 reduction to CO is decoupled electron-proton transfer, where the CO production can be carried out by the formation of a CO22 radical [55]. However, theoretical studies and quantum chemical calculations provide a better understanding of the mechanism and reaction pathway of electrochemical CO2 reduction. For example, the computational hydrogen electrode (CHE) model is one of the most convenient and widely employed methods for the PCET process [66]. Paterson et al. reported a CHE model for electrochemical CO2 reduction on the metallic copper electrode and it has been found that metal electrodes can easily generate hydrocarbon products upon CO2 reduction [67,68]. On the other hand, the natures of the catalysts used for catalytic reaction play important roles in the reaction pathway.
Another challenge in electrochemical CO2 reduction is competition with proton reduction, i.e., hydrogen evolution reaction (HER), because it is a two-electron and proton process, requiring far less energy [69]. This leads to poor selectivity to carbonaceous products. Last but not the least, CO2 has very low solubility in an aqueous medium (≈0.034 M), which negatively affects the overall CO2 reduction activity to chemical fuels. There are efforts to overcome this solubility problem by using gas diffusion layers to use the polymer electrolytes. In conclusion, both thermodynamic and kinetic barriers for the transformation to value-added fuels make it a challenging reaction. In the anode half of the electrochemical cell, oxygen evolution takes place. Lastly, the ion exchange membrane functions through avoiding oxidation of carbonaceous products by acting as a barrier to prevent the flow of electrons and allowing passage of protons only. A schematic of the electrochemical cell and reactions going on is shown in Figure 1, with the following key points:
  • CO2 adsorption on cathode surface;
  • PCET process;
  • Desorption of products formed from the cathode surface.

2.1. Parameters of Electrochemical CO2 Reduction

To examine the utility of electrochemical CO2 reaction there have been key parameters set that include current density (j), overpotential (η), faradaic efficiency (FE), turnover frequency (TOF), effect of electrolyte, and stability. Those parameters will be discussed in this section, which will assist the readers with understanding the concepts and applying them while evaluating their experiments pertaining to electrochemical CO2 reduction in the laboratory.

2.1.1. Current Density (j)

Current density is an important parameter for evaluating the activity of electrochemical CO2 reduction. Application of a constant cell voltage gives rise to electrolysis current. Therefore, current density (j) is described as the ratio between current (i) and the geometric surface area of working electrode (A), as shown in Equation (8).
j = i A

2.1.2. Faradaic Efficiency (FE)

Faradaic efficiency is yet another important guideline with which to evaluate the amount of product cross-over and electrochemical selectivity of CO2 reduction reaction [70]. It is given as the ratio of the charge from total product generated (αNF, whereas α is the number of electrons transferred, N is the amount of desired product’s moles, and F is the faraday constant 96,485 C mol1) to the total amount of charges passed during the electrolysis (Q), as shown in Equation (9) [55]. It also correlates with the selectivity of products during CO2 reduction.
E f a r a d a i c =   α N F Q

2.1.3. Overpotential (η)

The difference between the actual required potential and the equilibrium potential is expressed as overpotential. An electrocatalyst showcasing low overpotential is considered to show a better and efficient catalytic activity. Overpotential is calculated by the difference between the standard potential of the reaction of CO2 and the reduced product (Eeq), and actual electrode potential (E), as shown in Equation (10) [71].
η = E E e q

2.1.4. Turnover Frequency (TOF)

It is determined as the number of CO2 conversion products obtained over number of active sites per unit time. The catalyst with a high turnover frequency means it possess many innate active sites, which in turn, provide better activity. TOF is calculated by the ratio of the number of moles of the target product produced per unit time (Np) to the maximum number of moles of the active site of the catalyst (Nc), as shown in Equation (11) [71].
T O F =     N p N c

2.1.5. Effect of Electrolyte

The product selectivity can be altered/controlled by varying the electrolyte solution in electrochemical CO2 reduction reaction [72]. To date, various types of electrolytes have been used, such as non-aqueous (organic) solvents, aqueous and ionic liquids. Among them, aqueous electrolytes are cheap, environmentally friendly, and possess high sustainability. Moreover, in aqueous electrolytes, the concentration, pH value, and buffer capacity are of equal importance for eventually affecting the efficiency and product selectivity [73]. CO2 is highly soluble in organic electrolytes, but they are toxic, expensive and require large overpotentials compared to aqueous electrolytes. The ionic liquids are polar in nature which further helps to improve the conductivity and solubility of CO2.

2.1.6. Stability

Stability is one of the most important criteria in order to commercialize a catalyst. Therefore, even if a catalyst has high efficiency, it is equally as important to have long term stability.

3. 2D Electrocatalysts

The 2D materials possess excellent structural and morphological features, as shown in Figure 2. In the following sections, we will briefly describe the various 2D electrocatalysts for CO2 reduction into CO.

3.1. 2D Molybdenum (Mo) Transition Metal Dichalcogenides (TMDCs)

It is commonly known that the defects, edge sites, and dopants present with a catalyst greatly influence the activity of electrochemical CO2 reduction. In recent years, 2D MoS2 has shown great potential because of its unique physicochemical and structural properties. Asadi et al. reported bulk MoS2 with a layer-stacked structure for electrocatalytic CO2 reduction to produce CO [74]. STEM (Scanning transmission electron microscopy) investigation confirmed the presence of vertically aligned multi-layered sheets of MoS2 flakes. Remarkably, as-obtained metal terminated edges of MoS2 were considered reactive sites for electrocatalytic CO2 reduction. Excellent faradic efficiency of 98% was achieved with a high current density operated at a minimal overpotential of 54 mV. Therein, the presence of H+ from the water molecule (in electrolyte solution) and electron generated on the catalyst’s surface contribute to the reduction of CO2 to form CO. Density functional theory (DFT) studies revealed that the presence of MoS2 edge sites and metallic character of Mo atoms were collectively responsible for high catalytic activity and stability over 10 h. That theoretical study, significantly, provides a new idea for the experimental fabrication of a competent electrocatalyst. In another example, Nørskov et al. improved catalytic performance by breaking the linear scaling relationship among *CO, *CHO, and *COOH intermediates at the edge sites of MoS2 and MoSe2 [56]. Furthermore, in another study NÒrskov and coworkers demonstrated that transition metal-doped MoS2 significantly improved the catalytic activity [75]. Product selectivity mainly depends upon binding energies of edge sites present over the catalyst. For example, S edges of Ni-doped MoS2 and Mo edges of MoSe2 selectively produce CO. However, S edges of undoped MoS2 are selective towards H2 formation.
In 2016, Abbasi and coworkers reported Nb-doped vertically aligned MoS2 electrocatalyst for CO2 reduction into CO [57]. A small shift in Raman and XPS spectra confirmed Nb doping in MoS2 (Figure 3A,B). The CO2 reduction experiment was performed in an electrochemical cell by employing cyclic voltammetry at a −0.8 ± 1.0 V potential range. All tests were performed in an electrolyte solution composed of a CO2 saturated, 50:50 vol% mixture of 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIM-BF4) and deionized (DI) water. This optimized composition of electrolyte solution provides the maximum number of protons for CO2 reduction reaction. Here, 5% Nb-doped MoS2 nanosheets showed an improved current density compared to pristine MoS2 and Ta-doped MoS2 (Figure 3C). The catalyst displayed faradaic efficiencies oscillating from 12% to 82% towards CO formation in the potential range of −0.8 ± 0.16 V (Figure 3D). The formation rates for CO and H2 at 0.8 V were 2.23 × 10−9 mol min−1 and 1.98 × 10−9 mol min−1, respectively. As a result, Nb-MoS2 nanosheets exhibited an excellent catalytic activity in terms of turnover frequency (TOF) at an overpotential of 50–150 mV. In-depth DFT calculations were also conducted to understand the CO2 reduction pathway. The results showed that Mo edges of pure MoS2 preferably form an *COOH intermediate at lower overpotential. Moreover, Nb atoms help to reduce the binding strength among Mo and CO, thereby minimizing onset potential to produce CO via formation of *CO and *COOH intermediates. These findings revealed that Nb-doped MoS2 promotes faster desorption of CO compared to pristine MoS2. Furthermore, DFT calculations revealed that a low concentration of Nb-doping selectively reduced CO2 to CO. However, an increased Nb concentration over 5% showed a negative effect on CO2 reduction performance because of the enlarged work function of Mo1−xNbxS2 composite. Therefore, an appropriate amount of metal doping to a catalyst can significantly influence electronic properties that may further intensify the overall catalytic performance. MoS2 is a well-known catalyst for hydrogen evolution due to its unique structural characteristics where S atoms can make a strong bond with H+ in the electrolyte solution [76,77,78]. Therefore, a smaller number of S atoms are available for the CO2 reduction reaction. Addressing the issue, Yu and coworkers employed a structure-directing approach and fabricated 3D TiO2@MoS2 to make bonding between the unsaturated S atom and Ti metal, consequently making a large number of Mo atoms available for CO2 reduction reaction [76]. The porous nanosheets of TiO2 acted as a CO2 adsorbent due to high surface area and abundant active sites (Figure 4A). Bonding between Ti and S atoms improved the electronic and adsorption properties of MoS2 (Figure 4B showing Raman spectra of Ti-S band). Here, the electrolyte solution played an important role in product selectivity. The N2-saturated solution led to hydrogen evolution, whereas current density was increased towards CO formation in the CO2-saturated solution. Faradic efficiency of ≈2.8% towards CO formation was observed at the potential of −0.21 V, while at a more negative potential it reached up to 85% (Figure 4C). The catalyst showed excellent stability over 1000 cycles which revealed its superiority (Figure 4D). Interestingly, long-term stability over 14 h was observed at a steady potential of −0.70 V.

3.2. 2D Graphene

Graphene has been playing a significant role in many advanced applications because of its high chemical stability, superior electronic and thermal properties, and high specific surface area [79]. Typically, pristine graphene is inactive towards electrocatalysis because of its ineffectiveness towards CO2 activation. This allows a greater extent of freedom for surface modifications to tune the catalytic properties. The graphene templates, graphene oxide (GO), and reduced graphene oxide (RGO) have been largely employed for catalytic applications [80,81,82,83]. The metal and non-metal doping and composite formation with other semiconductors have been exercised to improve the catalytic performance of graphene. To date, metals like Cu, Ni, Au, and Co have been successfully used as dopants/co-catalyst [84]. Organometallic complexes and metal-organic frameworks (MOFs) have also been tried with graphene for CO2 reduction [85,86]. Heteroatom doping (such as by nitrogen, boron, and sulfur) to graphene can effectively improve the CO2 adsorption [87]. These heteroatoms act as reaction centers for CO2 reduction and improve the catalytic activity [88,89]. For example, pyridinic N in the N-doped graphene possesses a lower free energy threshold to produce the COOH intermediate and selectively produce CO [90,91,92,93]. Additionally, the electrocatalyst comprising both metal and non-metal is significant for heterogeneous catalysis aiming towards product selectivity [94]. Su et al. synthesized Ni-nitrogenated-graphene (Ni-N-Gr) sheets for generation of CO from CO2 [95]. The CO2 reduction capability was examined by using cyclic voltammetry in the buffer solution and products were detected by GC-MS technique. At an optimized condition, Ni-N-Gr displayed the highest faradaic efficiency over 90% for CO production at −0.7 to −0.9 V. The Ni-N-Gr showed excellent catalytic performance with much higher activity than pristine Ni and N-Gr samples. To explore the significance of Ni in Ni-N-Gr composite, Cu-N-Gr was also tested as a control sample and results showed positive feedback towards the presence of Ni in composite. Further, the turnover frequencies of CO per electrochemically active Ni atom was up to 2700 h−1 at −0.7 V and 4600 h−1 at −0.8 V versus RHE. In another similar study, Ni2+ on nitrogen-doped graphene (Ni2+@NG) was investigated for CO formation through electrocatalysis [88]. Previously, such types of catalysts were synthesized by utilizing the conservative pyrolysis technique (800–1000 °C) and acid leaching process, which are not so ecofriendly. Therefore, in this work the synthesis was carried out by employing a simple solution-chemistry approach which is convenient for large scale applicability. In this method, first N-doped graphene with an ultrathin nanosheet structure was synthesized by pyrolyzing a mixture of g-C3N4 and glucose. Then, immobilization of Ni2+ in N-doped graphene was carried out by reacting Ni(NO3)2 with N-doped graphene in the aqueous solution. The presence of Ni2+ ions and Ni valence state on N-doped graphene were confirmed by XPS and XANES spectra, respectively. The electrocatalyst produced CO with the faradaic efficiency of 92% at −0.68 V versus RHE. In contrast, the efficiency of pristine N-doped graphene was below 60% at all potential capacities. Therefore, Ni2+ sites significantly facilitated the kinetics of CO2 reduction. Furthermore, to understand the effects of other transition metals ion, metals such as Co2+ and Fe3+ were likewise restrained on N-doped graphene. All the catalysts showed almost 4–5 times higher activity than pristine N-doped graphene. The composite with Ni2+ showed the highest faradaic efficiency of 92% compared to that of Fe3+ (85%) and Co2+ (65%); see Figure 5A. The formation of CO occurred via the *COOH (* is surface sites) intermediate (Figure 5B). The quadruple-pyridinic configuration of N-doped graphene exhibited the maximum free energy barrier which protonates *COOH to form *CO.
In certain composites, stabilization of the single metal atom within the 2D nanostructure is crucial to prevent aggregation. To address the issue, Jeong et al. reported single-atom Ni activated on RGO sheets for CO formation which showed a faradaic efficiency of 97% at 0.8 V versus RHE [96]. The linkage between Ni atoms and GO was accomplished by using the tris(2-benzimidazolylmethyl)amine (NTB) ligand (Figure 6A). Firstly, Ni(NTB) complex was prepared. The NTB ligand makes a strong bond with GO via π–π interaction and Ni ions via ligation by forming Ni(NTB)-GO complex. A uniform distribution of Ni atoms on 2D-RGO sheets was confirmed by TEM analysis. XANES analysis was used to confirm that nickel was in Ni2+ state rather than Ni0. In an inert condition, the catalysts displayed moderate activity towards the H2 at −0.71 V versus RHE. Significantly the current density was improved in CO2-saturated electrolyte, reaching the onset potential of −5 mA cm−2 at −0.60 V versus RHE. The observed current density was almost double at −1.1 V versus RHE, which implies its superior activity compared to that of control samples (Figure 6B). Ni-N-RGO showed the maximum faradaic efficiency (97%) for CO formation at −0.8 V versus RHE (Figure 6C). The reactivity of Ni-N4 active sites towards the intermediate formed in the reaction are responsible for such high catalytic performance. DFT study revealed that the proton interactions with active sites of Ni-N-RGO are unfavorable towards HER. Therefore, selectivity was observed towards CO2 reduction rather than H2 formation.
Jiang and coworkers also demonstrated that graphene shells with embedded Ni atoms were active electrocatalysts for CO2 reduction [97]. From DFT calculations it was observed that Ni atoms in the graphene layer can significantly lower the activation barrier of CO2 molecule, which further weakens the binding with CO to form a facile product by lowering the proton reduction reaction. In this way, Ni-graphene electrocatalyst showed an excellent catalytic activity towards CO formation with a faradaic efficiency of 93.2% at current of 20 mA/mg. The turnover frequency (TOF) of 8 s−1 was observed for the active Ni center and a 454,000 cumulative turnover number (TON) was recorded below −0.7 V overpotential during 20 h of continuous electrolysis.
Jiang and coworkers reported more results with single Ni atoms emended in a graphene sheet [98]. Resulting catalysts endow CO selectivity over 90% of in aqueous solution, and the highest faradaic efficiency of CO (95%) was obtained at an overpotential of 550 mV. Specifically, Ni single atomic sites present in graphene were responsible for high selectivity of CO with 90% of faradaic efficiency over the stability of 20 h. Further, CO formation current, more than 200 mA (50 mA cm−2) under a faradaic efficiency of 97%, was established by using a 4 cm2 anion membrane electrode setup. Other transition metals, including Co, Fe, and Mn, were also dispersed on a graphene layer and it was observed that the reaction pathway changes with change in transition metals, specifically hydrogen was produced in the presence of Co. DFT studies revealed that for the Ni single atoms, compared to Co atoms, the weaker binding of CO and higher barrier for hydrogen evolution reaction accounted for the superior CO2 reduction towards selective formation of CO.
The stability of the oxidation state among metals is very crucial for catalytic reactions. Copper is proven as one of the best and inexpensive catalysts for CO2 reduction reaction. Karapinar et al. reported highly active, single-site Cu-N-C for electroreduction of CO2 [99]. They found that during the catalytic process, isolated copper sites converted into metallic copper, which eventually act as active sites for the CO2 reduction. Apart from nickel and copper, the iron-based catalysts are also considered as the superior and most active materials due to their chemical stability towards the CO2 reduction under negative potential. The key parameter of this material is that Fe3+ can easily coordinate with N heteroatom supported on carbon and maintains its 3+ oxidation state throughout the electrocatalytic reaction [100]. Also, such Fe3+ sites can adsorb a greater number of CO2 for electrocatalysis. Zhang and co-workers reported a hybrid composite of Fe/N-Graphene [101]. The electrocatalyst was fabricated by annealing the mixture of GO with iron chloride (FeCl3) at 700–800 °C in an inert atmosphere, as shown in Figure 7A. The oxygen functional groups in GO adsorbed metal cations through the electrostatic interaction. Whereas, Fe atoms in N-graphene matrices created Fe-N4 sites, which further acted as CO2 activators. N dopant acted as an immobilizer among Fe atoms and the graphene surface through metal–nitrogen interaction. As a result, catalysts undergo selective CO2 reduction towards CO formation with a faradaic efficiency of 80% at low potential of −0.60 V versus RHE. The catalyst was operated over 10 h at −0.60 V versus RHE to examine its long-term stability. The mechanism of CO2 to CO conversion was studied by theoretical simulations. The formation of CO took place by sharing of two protons and two electrons via COOH* intermediate, as revealed in Figure 7. Here, substitution of nitrogen to graphene lowered the energy barrier of COOH* intermediate and facilitated the CO* formation.
Zhu and coworkers reported the superior catalytic activity of 94.2%, the faradaic efficiency for CO formation by employing 2D N-doped carbon nanorod arrays/ultrathin carbon nanosheet catalysts [102]. In the synthesis, aniline polymerization was carried out in the presence of carbon nanosheets and acid treatment, and resulting catalysts were obtained by pyrolysis. It was observed that the catalyst was proficient of retaining its activity for 30 h at a minimal onset overpotential of 97 mV. In the Tafel slope, a low value of 65 mV dec−1 specified a fast pre-equilibration process of transfer of one electron followed by a rate-determining step. DFT calculation suggests that the N pyridinic sites and 2D structure were responsible for lowering the free energy barriers of intermediates (*COOH and *CO) and boosting CO desorption to form CO. Therefore, these results showed an excellent catalytic activity among other reported doped carbon materials.
Apart from single metal atom insertion, the transition metal sulfides have also been tried with graphene. For example, silver sulfide (Ag2S) possesses excellent physical, chemical, and thermal properties that are suitable for catalytic applications. Along those lines, Zeng et al. reported Ag2S/N-S doped rGO hybrid for CO2 reduction [103]. A hydrothermally synthesized, Ag2S/N-S-doped rGO electrocatalyst was employed for CO2 reduction toward CO. Here, CO, H2, and CH4 were obtained as products, while selectivity for CO was observed in the wide potential range of −1.08 to −0.58 V versus RHE. The catalyst manifested a current density of 470 µAcm−2 at −0.759 V versus RHE. An excellent catalytic activity owing to a faradaic efficiency of 87%, and long-term stability over 40 h, were achieved by the catalyst. The faradaic efficiency of this catalyst was enormously higher than pristine N-S-rGO (3.2%) and Ag2S nanowires (41.7%) [104]. Significantly, an increase in current density was observed, as the potential was swept cathodically which specified the increase in CO2 reduction kinetics in the system. In the CO formation mechanism, Ag ions bind with CO2 molecule and generate a COOH intermediate by combining one proton and one electron. Afterwards, the intermediate (COOH) further reacts with a second proton and electron pair to form CO. The high surface area of rGO maximized the reactive sites for CO2 molecule and N-S doping eventually enhanced the electrical conductivity of rGO.
Pitch derived 2D N-doped porous carbon nanosheets, which are similar to graphene, were also found to be a superior catalyst for selective CO2 reduction with high stability [105]. For synthesis, the authors used a NH3 etching strategy using NaCl template. NH3 etching increases the catalyst’s surface area by creating nanopores, and further, helping to insert nitrogen atoms into a carbon framework. Resulting catalysts showed high catalytic performance with 84% of faradaic efficiency at an overpotential of −0.49 V and a negligible onset overpotential of −0.9 V for CO production. Such excellent catalytic behavior was due to presence of nanoporous 2D architecture and the high content of pyridinic N active sites which favored fast charge transfer, and high CO2 adsorption. DFT studies confirmed that the pyridinic N sites favorably bind with COOH* intermediates and further facilitate catalytic reduction.

3.3. 2D Tungsten TMDCs

Tungsten is an emerging candidate for sustainable energy applications due to its low price, non-toxicity, and large atomic size compared with other TMDCs [106,107,108,109,110,111]. Due to its exciting electronic properties, tungsten dichalcogenides act as a catalytic site for electrochemical CO2 reduction reactions. Also, tungsten dichalcogenides have the anisotropic properties of its basal and edge plane; therefore, most of the catalytic reactions occur on the edge sites [112]. These limited catalytic sites can be increased by transition metal doping [113]. In this regard, Asadi et al. fabricated the WSe2 nanoflakes (NFs) by a chemical vapor transport growth and exfoliation technique [114]. They found that the CO2 reduction activity of TMDCs in 50% EMIM-BF4 with water was much higher compared to other TMDCs and Ag NPs. According to work functions estimation, the trend of monolayer TMDCs showed activity in the order WSe2 < MoSe2 < WS2 < MoS2 (Figure 8A). WSe2 NFs had the lowest work function of those comprised of monolayered TMDCs. Also, electrochemical impedance spectroscopy (EIS) outcomes were matched with the work function trend (Figure 8B).
Charge transfer resistance (Rct) is an important kinetic factor for the electrochemical rate constant. Charge transmission resistance of WSe2 NFs (≈180 Ω) was lower than MoS2 and Ag nanoparticles (Figure 8B) [114]. Therefore, WSe2 possesses a fast electron transfer property which eventually affects the CO2 reduction reaction. Experiments were performed in an electrolyte solution of 50% EMIM-BF4, 50% water. It was found that WSe2 NFs displayed a 18.95 mA cm−2 current density for CO at a minimal potential of 54 mV (Figure 8C,D). Here, EMIM+ ion in the ionic liquid acted as a transporter of CO2 to the catalyst surface. Also, the CO intermediate was more stable on the metallic edges of WSe2 NFs, suggesting the possibility of CO formation after CO2 reduction. Therefore, such systems owing to high catalytic performance operating at low overpotentials have the disruptive potential to solve the current obstacles of CO2 reduction technology.

3.4. 2D InSe

Among various 2D materials, InSe is also considered as an emerging electrocatalyst. Although InSe is not a transition metal chalcogenide, we intentionally added this section to our study. Recently, the theoretical study about InSe was carried out by Jiang et al., which demonstrated its potential for CO2 reduction [115]. The single metal (e.g., Rh, Mn, Cu, Zn, and Cr) was embedded in a 2D InSe substrate and catalytic reactivity towards CO, HCOOH, H2, and CH4 were studied. They showed that InSe systems with Rh, Mn, and Cu metals obeyed the linear scaling relationships among adsorption energies of intermediates, such as *CO and *COOH. However, the relationship broke down for system containing Zn and Cr metals. Such an interesting feature of 2D InSe offers a reliable method for the experimental usage, and in turn, the product selectivity. The Zn@2D InSe system effectively produces CO with limiting the potential of −0.78 V because it is situated on the scaling line among *COOH and *CO. Its lowered desorption energy (0.14 eV) executes selectivity towards CO. Therefore, such theoretical studies enrich the understanding of 2D systems for electrocatalytic applications.
The summary of various 2D nanomaterials for electrochemical CO2 reduction is displayed in Table 2.

4. Conclusions

In the field of electrocatalytic CO2 reduction, recent studies were mainly concentrated on the design of various types of efficient catalysts. A variety of tailored 2D nanostructures were fabricated by researchers with the rational design of experimental procedures. In particular, metal and non-metal doped 2D templates could be highly efficient and stable catalysts. In this study, we have briefly reviewed the recent developments in selective electrochemical CO2 reduction to CO using various 2D nanostructures. The foremost goal of this study was to introduce the fundamentals and different synthetic strategies of 2D materials. The high surface area of 2D nanostructure encourages CO2 adsorption, and the presence of edge sites/dopants helps to generate reactive sites for CO2 reduction. Besides, single atom doped 2D structures hold great potential for electrocatalytic CO2 reduction, since they have been shown to have high faradaic efficiency along with CO selectivity. While highlighting the different fabrication methods, we have also paid considerable attention to the theoretical studies because theoretical simulations are of equal importance to understand the reactivity, product selectivity, and mechanism of CO2 reduction pathway. Although several noteworthy successes in CO2 reduction have been achieved by 2D based nanostructures, further efforts are required for large scale applications. Additionally, more focus should be given toward improving upon the green, scalable, and sustainable electrocatalyst fabrications, which is becoming a crucial factor for realistic industrial applications. Apart from that, CO2 transformation into higher hydrocarbon products such as ethane, ethanol, or ethylene is also of similar importance, and has a great challenge ahead due to the low efficiency of the existing catalysts. Copper-based catalysts have been proven to be superior materials for generating higher hydrocarbons, including C2 and C3 products. However, we believe that the interesting surface, electronic, and structural properties of 2D materials could also be significant in higher hydrocarbon production in the near future.

Author Contributions

S.-I.I. conceptualized and edited the manuscript; C.B.H., H.K., J.L., S.S., and C.E. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thankfully acknowledge the support of the Ministry of Science and ICT (2017R1E1A1A01074890 and 2017M2A2A6A01070912). This research was also supported by the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science and ICT (2015M1A2A2074670) and by Flux Photon Corporation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, H.; Li, P.; Liu, J.; Chen, Z.; Liu, L.; Dontsova, D.; Yan, R.; Fan, T.; Zhang, D.; Ye, J. Biomimetic polymeric semiconductor based hybrid nanosystems for artificial photosynthesis towards solar fuels generation via CO2 reduction. Nano Energy 2016, 25, 128–135. [Google Scholar] [CrossRef]
  2. Sorcar, S.; Thompson, J.; Hwang, Y.; Park, Y.H.; Majima, T.; Grimes, C.A.; Durrant, J.R.; In, S.-I. High-rate solar-light photoconversion of CO2 to fuel: Controllable transformation from C1 to C2 products. Energy Environ. Sci. 2018, 11, 3183–3193. [Google Scholar] [CrossRef]
  3. Marszewski, M.; Cao, S.; Yu, J.; Jaroniec, M. Semiconductor-based photocatalytic CO2 conversion. Mater. Horizons 2015, 2, 261–278. [Google Scholar] [CrossRef]
  4. Bushuyev, O.S.; De Luna, P.; Dinh, C.T.; Tao, L.; Saur, G.; van de Lagemaat, J.; Kelley, S.O.; Sargent, E.H. What should we make with CO2 and how can we make it? Joule 2018, 2, 825–832. [Google Scholar] [CrossRef] [Green Version]
  5. Razzaq, A.; In, S.-I. TiO2 Based Nanostructures for Photocatalytic CO2 Conversion to Valuable Chemicals. Micromachines 2019, 10, 326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Thomas, J.M.; Harris, K.D.M. Some of tomorrow’s catalysts for processing renewable and non-renewable feedstocks, diminishing anthropogenic carbon dioxide and increasing the production of energy. Energy Environ. Sci. 2016, 9, 687–708. [Google Scholar] [CrossRef]
  7. Hiragond, C.; Ali, S.; Sorcar, S.; In, S.-I. Hierarchical Nanostructured Photocatalysts for CO2 Photoreduction. Catalysts 2019, 9, 370. [Google Scholar] [CrossRef] [Green Version]
  8. Ali, S.; Razzaq, A.; In, S.-I. Development of graphene based photocatalysts for CO2 reduction to C1 chemicals: A brief overview. Catal. Today 2019, 335, 39–54. [Google Scholar] [CrossRef]
  9. 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]
  10. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and perspectives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic processes. Energy Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef] [Green Version]
  11. Taheri Najafabadi, A. CO2 chemical conversion to useful products: An engineering insight to the latest advances toward sustainability. Int. J. Energy Res. 2013, 37, 485–499. [Google Scholar] [CrossRef]
  12. Lorentzou, S.; Karagiannakis, G.; Pagkoura, C.; Zygogianni, A.; Konstandopoulos, A.G. Thermochemical CO2 and CO2/H2O splitting over NiFe2O4 for solar fuels synthesis. Energy Procedia 2014, 49, 1999–2008. [Google Scholar] [CrossRef] [Green Version]
  13. Chueh, W.C.; Falter, C.; Abbott, M.; Scipio, D.; Furler, P.; Haile, S.M.; Steinfeld, A. High-flux solar-driven thermochemical dissociation of CO2 and H2O using nonstoichiometric ceria. Science 2010, 330, 1797–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sorcar, S.; Hwang, Y.; Grimes, C.A.; In, S.-I. Highly enhanced and stable activity of defect-induced titania nanoparticles for solar light-driven CO2 reduction into CH4. Mater. Today 2017, 20, 507–515. [Google Scholar] [CrossRef]
  15. Sorcar, S.; Hwang, Y.; Lee, J.; Kim, H.; Grimes, K.M.; Grimes, C.A.; Jung, J.-W.; Cho, C.-H.; Majima, T.; Hoffmann, M.R.; et al. CO2, water, and sunlight to hydrocarbon fuels: A sustained sunlight to fuel (Joule-to-Joule) photoconversion efficiency of 1%. Energy Environ. Sci. 2019, 12, 2685–2696. [Google Scholar] [CrossRef] [Green Version]
  16. Razzaq, A.; Sinhamahapatra, A.; Kang, T.-H.; Grimes, C.A.; Yu, J.-S.; In, S.-I. Efficient solar light photoreduction of CO2 to hydrocarbon fuels via magnesiothermally reduced TiO2 photocatalyst. Appl. Catal. B Environ. 2017, 215, 28–35. [Google Scholar] [CrossRef]
  17. Zubair, M.; Kim, H.; Razzaq, A.; Grimes, C.A.; In, S.-I. Solar spectrum photocatalytic conversion of CO2 to CH4 utilizing TiO2 nanotube arrays embedded with graphene quantum dots. J. CO2 Util. 2018, 26, 70–79. [Google Scholar] [CrossRef]
  18. Ali, S.; Flores, M.C.; Razzaq, A.; Sorcar, S.; Hiragond, C.B.; Kim, H.R.; Park, Y.H.; Hwang, Y.; Kim, H.S.; Kim, H.; et al. Gas Phase Photocatalytic CO2 Reduction,“A Brief Overview for Benchmarking”. Catalysts 2019, 9, 727. [Google Scholar] [CrossRef] [Green Version]
  19. Kim, H.R.; Razzaq, A.; Grimes, C.A.; In, S.-I. Heterojunction pnp Cu2O/S-TiO2/CuO: Synthesis and application to photocatalytic conversion of CO2 to methane. J. CO2 Util. 2017, 20, 91–96. [Google Scholar] [CrossRef]
  20. Kim, K.; Razzaq, A.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid mesoporous Cu2 ZnSnS4 (CZTS)–TiO2 photocatalyst for efficient photocatalytic conversion of CO2 into CH4 under solar irradiation. RSC Adv. 2016, 6, 38964–38971. [Google Scholar] [CrossRef]
  21. Park, S.-M.; Razzaq, A.; Park, Y.H.; Sorcar, S.; Park, Y.; Grimes, C.A.; In, S.-I. Hybrid CuxO–TiO2 Heterostructured Composites for Photocatalytic CO2 Reduction into Methane Using Solar Irradiation: Sunlight into Fuel. ACS Omega 2016, 1, 868–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ganesh, I. Electrochemical conversion of carbon dioxide into renewable fuel chemicals–The role of nanomaterials and the commercialization. Renew. Sustain. Energy Rev. 2016, 59, 1269–1297. [Google Scholar] [CrossRef]
  23. Benson, E.E.; Kubiak, C.P.; Sathrum, A.J.; Smieja, J.M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89–99. [Google Scholar] [CrossRef] [PubMed]
  24. Bhugun, I.; Lexa, D.; Savéant, J.-M. Catalysis of the electrochemical reduction of carbon dioxide by iron (0) porphyrins: Synergystic effect of weak Brönsted acids. J. Am. Chem. Soc. 1996, 118, 1769–1776. [Google Scholar] [CrossRef]
  25. 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]
  26. Olajire, A.A. A review of mineral carbonation technology in sequestration of CO2. J. Pet. Sci. Eng. 2013, 109, 364–392. [Google Scholar] [CrossRef]
  27. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294. [Google Scholar] [CrossRef]
  28. Yu, F.; Wei, P.; Yang, Y.; Chen, Y.; Guo, L.; Peng, Z. Material design at nano and atomic scale for electrocatalytic CO2 reduction. Nano Mater. Sci. 2019, 1, 60–69. [Google Scholar] [CrossRef]
  29. Khezri, B.; Fisher, A.C.; Pumera, M. CO2 reduction: The quest for electrocatalytic materials. J. Mater. Chem. A 2017, 5, 8230–8246. [Google Scholar] [CrossRef]
  30. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and perspective of electrocatalytic CO2 reduction for renewable carbonaceous fuels and chemicals. Adv. Sci. 2018, 5(5), 1700275. [Google Scholar] [CrossRef]
  31. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Commun. 2017, 8, 825–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  33. Kuhl, K.P.; Hatsukade, T.; Cave, E.R.; Abram, D.N.; Kibsgaard, J.; Jaramillo, T.F. Electrocatalytic conversion of carbon dioxide to methane and methanol on transition metal surfaces. J. Am. Chem. Soc. 2014, 136, 14107–14113. [Google Scholar] [CrossRef] [PubMed]
  34. Rogers, C.; Perkins, W.S.; Veber, G.; Williams, T.E.; Cloke, R.R.; Fischer, F.R. Synergistic enhancement of electrocatalytic CO2 reduction with gold nanoparticles embedded in functional graphene nanoribbon composite electrodes. J. Am. Chem. Soc. 2017, 139, 4052–4061. [Google Scholar] [CrossRef] [Green Version]
  35. Hernández, S.; Farkhondehfal, M.A.; Sastre, F.; Makkee, M.; Saracco, G.; Russo, N. Syngas production from electrochemical reduction of CO2: Current status and prospective implementation. Green Chem. 2017, 19, 2326–2346. [Google Scholar] [CrossRef] [Green Version]
  36. Zhu, W.; Michalsky, R.; Metin, O.; Lv, H.; Guo, S.; Wright, C.J.; Sun, X.; Peterson, A.A.; Sun, S. Monodisperse Au nanoparticles for selective electrocatalytic reduction of CO2 to CO. J. Am. Chem. Soc. 2013, 135, 16833–16836. [Google Scholar] [CrossRef]
  37. Zhou, W.; Cheng, K.; Kang, J.; Zhou, C.; Subramanian, V.; Zhang, Q.; Wang, Y. New horizon in C1 chemistry: Breaking the selectivity limitation in transformation of syngas and hydrogenation of CO2 into hydrocarbon chemicals and fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. [Google Scholar] [CrossRef]
  38. Ju, W.; Bagger, A.; Hao, G.-P.; Varela, A.S.; Sinev, I.; Bon, V.; Cuenya, B.R.; Kaskel, S.; Rossmeisl, J.; Strasser, P. Understanding activity and selectivity of metal-nitrogen-doped carbon catalysts for electrochemical reduction of CO2. Nat. Commun. 2017, 8, 944. [Google Scholar] [CrossRef]
  39. Dai, L.; Qin, Q.; Wang, P.; Zhao, X.; Hu, C.; Liu, P.; Qin, R.; Chen, M.; Ou, D.; Xu, C. Ultrastable atomic copper nanosheets for selective electrochemical reduction of carbon dioxide. Sci. Adv. 2017, 3, e1701069. [Google Scholar] [CrossRef] [Green Version]
  40. Rasul, S.; Pugnant, A.; Xiang, H.; Fontmorin, J.-M.; Eileen, H.Y. Low cost and efficient alloy electrocatalysts for CO2 reduction to formate. J. CO2 Util. 2019, 32, 1–10. [Google Scholar] [CrossRef]
  41. Qiu, J.-P.; Tong, Y.-W.; Zhao, D.-M.; He, Z.-Q.; Chen, J.-M.; Song, S. Electrochemical reduction of CO2 to methanol at TiO2 nanotube electrodes. Acta Phys. Chim. Sin. 2017, 33, 1411–1420. [Google Scholar]
  42. Gao, D.; Zhang, Y.; Zhou, Z.; Cai, F.; Zhao, X.; Huang, W.; Li, Y.; Zhu, J.; Liu, P.; Yang, F. Enhancing CO2 electroreduction with the metal–oxide interface. J. Am. Chem. Soc. 2017, 139, 5652–5655. [Google Scholar] [CrossRef] [PubMed]
  43. Lin, S.; Diercks, C.S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.M. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Li, N.; Chen, X.; Ong, W.-J.; MacFarlane, D.R.; Zhao, X.; Cheetham, A.K.; Sun, C. Understanding of electrochemical mechanisms for CO2 capture and conversion into hydrocarbon fuels in transition-metal carbides (MXenes). ACS Nano 2017, 11, 10825–10833. [Google Scholar] [CrossRef]
  45. Back, S.; Jung, Y. TiC-and TiN-supported single-atom catalysts for dramatic improvements in CO2 electrochemical reduction to CH4. ACS Energy Lett. 2017, 2, 969–975. [Google Scholar] [CrossRef]
  46. Wang, Y.; Hou, P.; Wang, Z.; Kang, P. Zinc Imidazolate Metal–Organic Frameworks (ZIF-8) for Electrochemical Reduction of CO2 to CO. ChemPhysChem 2017, 18, 3142–3147. [Google Scholar] [CrossRef] [Green Version]
  47. Dou, S.; Song, J.; Xi, S.; Du, Y.; Wang, J.; Huang, Z.; Xu, Z.J.; Wang, X. Boosting Electrochemical CO2 Reduction on Metal–Organic Frameworks via Ligand Doping. Angew. Chemie 2019, 131, 4081–4085. [Google Scholar] [CrossRef]
  48. Shao, P.; Yi, L.; Chen, S.; Zhou, T.; Zhang, J. Metal-organic frameworks for electrochemical reduction of carbon dioxide: The role of metal centers. J. Energy Chem. 2019, 40, 156–170. [Google Scholar] [CrossRef] [Green Version]
  49. Hao, J.; Shi, W. Transition metal (Mo, Fe, Co, and Ni)-based catalysts for electrochemical CO2 reduction. Chin. J. Catal. 2018, 39, 1157–1166. [Google Scholar] [CrossRef]
  50. Tayyebi, E.; Hussain, J.; Abghoui, Y.; Skúlason, E. Trends of electrochemical CO2 reduction reaction on transition metal oxide catalysts. J. Phys. Chem. C 2018, 122, 10078–10087. [Google Scholar] [CrossRef]
  51. Qu, J.; Zhang, X.; Wang, Y.; Xie, C. Electrochemical reduction of CO2 on RuO2/TiO2 nanotubes composite modified Pt electrode. Electrochim. Acta 2005, 50, 3576–3580. [Google Scholar] [CrossRef]
  52. 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]
  53. Zhang, H.; Cheng, H.-M.; Ye, P. 2D nanomaterials: Beyond graphene and transition metal dichalcogenides. Chem. Soc. Rev. 2018, 47, 6009–6012. [Google Scholar] [CrossRef] [PubMed]
  54. Cai, X.; Luo, Y.; Liu, B.; Cheng, H.-M. Preparation of 2D material dispersions and their applications. Chem. Soc. Rev. 2018, 47, 6224–6266. [Google Scholar] [CrossRef]
  55. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and challenges of electrochemical CO2 reduction using two-dimensional materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef] [Green Version]
  56. Hong, X.; Chan, K.; Tsai, C.; Nørskov, J.K. How doped MoS2 breaks transition-metal scaling relations for CO2 electrochemical reduction. ACS Catal. 2016, 6, 4428–4437. [Google Scholar] [CrossRef]
  57. Abbasi, P.; Asadi, M.; Liu, C.; Sharifi-Asl, S.; Sayahpour, B.; Behranginia, A.; Zapol, P.; Shahbazian-Yassar, R.; Curtiss, L.A.; Salehi-Khojin, A. Tailoring the edge structure of molybdenum disulfide toward electrocatalytic reduction of carbon dioxide. ACS Nano 2016, 11, 453–460. [Google Scholar] [CrossRef]
  58. Indrakanti, V.P.; Kubicki, J.D.; Schobert, H.H. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook. Energy Environ. Sci. 2009, 2, 745–758. [Google Scholar] [CrossRef]
  59. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chemie Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  60. Wu, J.; Huang, Y.; Ye, W.; Li, Y. CO2 reduction: From the electrochemical to photochemical approach. Adv. Sci. 2017, 4, 1700194. [Google Scholar] [CrossRef]
  61. Yang, Y.; Lee, J.-W. Toward ideal carbon dioxide functionalization. Chem. Sci. 2019, 10, 3905–3926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. 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] [PubMed]
  63. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  64. Hong, J.; Zhang, W.; Ren, J.; Xu, R. Photocatalytic reduction of CO2: A brief review on product analysis and systematic methods. Anal. Methods 2013, 5, 1086–1097. [Google Scholar] [CrossRef]
  65. Mori, K.; Yamashita, H.; Anpo, M. Photocatalytic reduction of CO2 with H2O on various titanium oxide photocatalysts. Rsc Adv. 2012, 2, 3165–3172. [Google Scholar] [CrossRef]
  66. Liu, L.; Liu, C. Origin of the overpotentials for HCOO and CO formation in the electroreduction of CO2 on Cu (211): The reductive desorption processes decide. Phys. Chem. Chem. Phys. 2018, 20, 5756–5765. [Google Scholar] [CrossRef]
  67. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  68. Tian, Z.; Priest, C.; Chen, L. Recent Progress in the Theoretical Investigation of Electrocatalytic Reduction of CO2. Adv. Theory Simul. 2018, 1, 1800004. [Google Scholar] [CrossRef]
  69. Elgrishi, N.; Chambers, M.B.; Fontecave, M. Turning it off! Disfavouring hydrogen evolution to enhance selectivity for CO production during homogeneous CO2 reduction by cobalt–terpyridine complexes. Chem. Sci. 2015, 6, 2522–2531. [Google Scholar] [CrossRef] [Green Version]
  70. Rongé, J.; Bosserez, T.; Martel, D.; Nervi, C.; Boarino, L.; Taulelle, F.; Decher, G.; Bordiga, S.; Martens, J.A. Monolithic cells for solar fuels. Chem. Soc. Rev. 2014, 43, 7963–7981. [Google Scholar] [CrossRef] [Green Version]
  71. Costentin, C.; Passard, G.; Robert, M.; Savéant, J.-M. Ultraefficient homogeneous catalyst for the CO2-to-CO electrochemical conversion. Proc. Natl. Acad. Sci. USA 2014, 111, 14990–14994. [Google Scholar] [CrossRef] [Green Version]
  72. Albo, J.; Alvarez-Guerra, M.; Castaño, P.; Irabien, A. Towards the electrochemical conversion of carbon dioxide into methanol. Green Chem. 2015, 17, 2304–2324. [Google Scholar] [CrossRef]
  73. König, M.; Vaes, J.; Klemm, E.; Pant, D. Solvents and Supporting Electrolytes in the Electrocatalytic Reduction of CO2. iScience 2019, 19, 135. [Google Scholar] [CrossRef] [PubMed]
  74. Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B.A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R. Robust carbon dioxide reduction on molybdenum disulphide edges. Nat. Commun. 2014, 5, 4470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Chan, K.; Tsai, C.; Hansen, H.A.; Nørskov, J.K. Molybdenum sulfides and selenides as possible electrocatalysts for CO2 reduction. ChemCatChem 2014, 6, 1899–1905. [Google Scholar] [CrossRef]
  76. Yu, L.; Xie, Y.; Zhou, J.; Li, Y.; Yu, Y.; Ren, Z. Robust and selective electrochemical reduction of CO2: The case of integrated 3DTiO2@ MoS2 architectures and Ti–S bonding effects. J. Mater. Chem. A 2018, 6, 4706–4713. [Google Scholar] [CrossRef]
  77. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jørgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Nørskov, J.K. Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef]
  78. Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic effect of MoS2 and graphene as cocatalysts for enhanced photocatalytic H2 production activity of TiO2 nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575–6578. [Google Scholar] [CrossRef]
  79. Ma, T.; Fan, Q.; Li, X.; Qiu, J.; Wu, T.; Sun, Z. Graphene-based materials for electrochemical CO2 reduction. J. CO2 Util. 2019, 30, 168–182. [Google Scholar] [CrossRef]
  80. Hossain, M.N.; Wen, J.; Chen, A. Unique copper and reduced graphene oxide nanocomposite toward the efficient electrochemical reduction of carbon dioxide. Sci. Rep. 2017, 7, 3184. [Google Scholar] [CrossRef] [Green Version]
  81. Zhang, Z.; Ahmad, F.; Zhao, W.; Yan, W.; Zhang, W.; Huang, H.; Ma, C.; Zeng, J. Enhanced Electrocatalytic Reduction of CO2 via Chemical Coupling between Indium Oxide and Reduced Graphene Oxide. Nano Lett. 2019, 19, 4029–4034. [Google Scholar] [CrossRef] [PubMed]
  82. Yuan, J.; Zhi, W.-Y.; Liu, L.; Yang, M.-P.; Wang, H.; Lu, J.-X. Electrochemical reduction of CO2 at metal-free N-functionalized graphene oxide electrodes. Electrochim. Acta 2018, 282, 694–701. [Google Scholar] [CrossRef]
  83. Zhao, Y.; Wang, C.; Liu, Y.; MacFarlane, D.R.; Wallace, G.G. Engineering surface amine modifiers of ultrasmall gold nanoparticles supported on reduced graphene oxide for improved electrochemical CO2 reduction. Adv. Energy Mater. 2018, 8, 1801400. [Google Scholar] [CrossRef] [Green Version]
  84. Liu, X.; Zhu, L.; Wang, H.; He, G.; Bian, Z. Catalysis performance comparison for electrochemical reduction of CO2 on Pd–Cu/graphene catalyst. RSC Adv. 2016, 6, 38380–38387. [Google Scholar] [CrossRef]
  85. 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.M. Electrocatalytic reduction of carbon dioxide to carbon monoxide and methane at an immobilized cobalt protoporphyrin. Nat. Commun. 2015, 6, 8177. [Google Scholar] [CrossRef] [PubMed]
  86. 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] [Green Version]
  87. Liu, T.; Ali, S.; Lian, Z.; Li, B.; Su, D.S. CO2 electoreduction reaction on heteroatom-doped carbon cathode materials. J. Mater. Chem. A 2017, 5, 21596–21603. [Google Scholar] [CrossRef]
  88. Bi, W.; Li, X.; You, R.; Chen, M.; Yuan, R.; Huang, W.; Wu, X.; Chu, W.; Wu, C.; Xie, Y. Surface immobilization of transition metal ions on nitrogen-doped graphene realizing high-efficient and selective CO2 reduction. Adv. Mater. 2018, 30, 1706617. [Google Scholar] [CrossRef]
  89. Cui, H.; Guo, Y.; Guo, L.; Wang, L.; Zhou, Z.; Peng, Z. Heteroatom-doped carbon materials and their composites as electrocatalysts for CO2 reduction. J. Mater. Chem. A 2018, 6, 18782–18793. [Google Scholar] [CrossRef]
  90. Li, W.; Seredych, M.; Rodríguez-Castellón, E.; Bandosz, T.J. Metal-free Nanoporous Carbon as a Catalyst for Electrochemical Reduction of CO2 to CO and CH4. ChemSusChem 2016, 9, 606–616. [Google Scholar] [CrossRef]
  91. Liu, S.; Yang, H.; Huang, X.; Liu, L.; Cai, W.; Gao, J.; Li, X.; Zhang, T.; Huang, Y.; Liu, B. Identifying Active Sites of Nitrogen-Doped Carbon Materials for the CO2 Reduction Reaction. Adv. Funct. Mater. 2018, 28, 1800499. [Google Scholar] [CrossRef]
  92. Zou, X.; Liu, M.; Wu, J.; Ajayan, P.M.; Li, J.; Liu, B.; Yakobson, B.I. How nitrogen-doped graphene quantum dots catalyze electroreduction of CO2 to hydrocarbons and oxygenates. ACS Catal. 2017, 7, 6245–6250. [Google Scholar] [CrossRef]
  93. Wang, H.; Jia, J.; Song, P.; Wang, Q.; Li, D.; Min, S.; Qian, C.; Wang, L.; Li, Y.F.; Ma, C. Efficient Electrocatalytic Reduction of CO2 by Nitrogen-Doped Nanoporous Carbon/Carbon Nanotube Membranes: A Step Towards the Electrochemical CO2 Refinery. Angew. Chemie Int. Ed. 2017, 56, 7847–7852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Varela, A.S.; Ranjbar Sahraie, N.; Steinberg, J.; Ju, W.; Oh, H.; Strasser, P. Metal-doped nitrogenated carbon as an efficient catalyst for direct CO2 electroreduction to CO and hydrocarbons. Angew. Chemie Int. Ed. 2015, 54, 10758–10762. [Google Scholar] [CrossRef] [PubMed]
  95. Su, P.; Iwase, K.; Nakanishi, S.; Hashimoto, K.; Kamiya, K. Nickel-Nitrogen-Modified Graphene: An Efficient Electrocatalyst for the Reduction of Carbon Dioxide to Carbon Monoxide. Small 2016, 12, 6083–6089. [Google Scholar] [CrossRef] [PubMed]
  96. Jeong, H.; Balamurugan, M.; Choutipalli, V.S.K.; Jo, J.; Baik, H.; Subramanian, V.; Kim, M.; Sim, U.; Nam, K.T. Tris (2-benzimidazolylmethyl) amine-Directed Synthesis of Single-Atom Nickel Catalysts for Electrochemical CO Production from CO2. Chem. Eur. J. 2018, 24, 18444–18454. [Google Scholar] [CrossRef]
  97. Jiang, K.; Siahrostami, S.; Akey, A.J.; Li, Y.; Lu, Z.; Lattimer, J.; Hu, Y.; Stokes, C.; Gangishetty, M.; Chen, G. Transition-metal single atoms in a graphene shell as active centers for highly efficient artificial photosynthesis. Chem 2017, 3, 950–960. [Google Scholar] [CrossRef] [Green Version]
  98. Jiang, K.; Siahrostami, S.; Zheng, T.; Hu, Y.; Hwang, S.; Stavitski, E.; Peng, Y.; Dynes, J.; Gangisetty, M.; Su, D. Isolated Ni single atoms in graphene nanosheets for high-performance CO2 reduction. Energy Environ. Sci. 2018, 11, 893–903. [Google Scholar] [CrossRef]
  99. Karapinar, D.; Huan, N.T.; Ranjbar, N.S.; Li, J.; Wakerley, D.; Touati, N.; Zanna, S.; Taverna, D.; Galvão, L.H.T.; Zitolo, A. Electroreduction of CO2 on Single-Site Copper-Nitrogen-Doped Carbon Material: Selective Formation of Ethanol and Reversible Restructuration of the Metal Sites. Angew. Chem. Int. Ed. Engl. 2019, 58, 15098–15103. [Google Scholar] [CrossRef]
  100. Gu, J.; Hsu, C.-S.; Bai, L.; Chen, H.M.; Hu, X. Atomically dispersed Fe3+ sites catalyze efficient CO2 electroreduction to CO. Science 2019, 364, 1091–1094. [Google Scholar] [CrossRef]
  101. Zhang, C.; Yang, S.; Wu, J.; Liu, M.; Yazdi, S.; Ren, M.; Sha, J.; Zhong, J.; Nie, K.; Jalilov, A.S. Electrochemical CO2 Reduction with Atomic Iron-Dispersed on Nitrogen-Doped Graphene. Adv. Energy Mater. 2018, 8, 1703487. [Google Scholar] [CrossRef]
  102. Zhu, Y.; Lv, K.; Wang, X.; Yang, H.; Zheng Xiao, G. 1D/2D nitrogen-doped carbon nanorod arrays/ultrathin carbon nanosheets: Outstanding catalysts for highly efficient electroreduction CO2 to CO. J. Mater. Chem. A 2019, 7, 14895–14903. [Google Scholar] [CrossRef]
  103. Zeng, L.; Shi, J.; Luo, J.; Chen, H. Silver sulfide anchored on reduced graphene oxide as a high-performance catalyst for CO2 electroreduction. J. Power Sources 2018, 398, 83–90. [Google Scholar] [CrossRef]
  104. Liu, S.; Tao, H.; Liu, Q.; Xu, Z.; Liu, Q.; Luo, J.-L. Rational design of silver sulfide nanowires for efficient CO2 electroreduction in ionic liquid. ACS Catal. 2018, 8, 1469–1475. [Google Scholar] [CrossRef]
  105. Li, H.; Xiao, N.; Hao, M.; Song, X.; Wang, Y.; Ji, Y.; Liu, C.; Li, C.; Guo, Z.; Zhang, F. Efficient CO2 electroreduction over pyridinic-N active sites highly exposed on wrinkled porous carbon nanosheets. Chem. Eng. J. 2018, 351, 613–621. [Google Scholar] [CrossRef]
  106. Share, K.; Lewis, J.; Oakes, L.; Carter, R.E.; Cohn, A.P.; Pint, C.L. Tungsten diselenide (WSe2) as a high capacity, low overpotential conversion electrode for sodium ion batteries. RSC Adv. 2015, 5, 101262–101267. [Google Scholar] [CrossRef]
  107. Bissett, M.A.; Worrall, S.D.; Kinloch, I.A.; Dryfe, R.A.W. Comparison of two-dimensional transition metal dichalcogenides for electrochemical supercapacitors. Electrochim. Acta 2016, 201, 30–37. [Google Scholar] [CrossRef]
  108. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699. [Google Scholar] [CrossRef]
  109. Eftekhari, A. Tungsten dichalcogenides (WS2, WSe2, and WTe2): Materials chemistry and applications. J. Mater. Chem. A 2017, 5, 18299–18325. [Google Scholar] [CrossRef]
  110. Zheng, Z.; Zhang, T.; Yao, J.; Zhang, Y.; Xu, J.; Yang, G. Flexible, transparent and ultra-broadband photodetector based on large-area WSe2 film for wearable devices. Nanotechnology 2016, 27, 225501. [Google Scholar] [CrossRef]
  111. Pumera, M.; Loo, A.H. Layered transition-metal dichalcogenides (MoS2 and WS2) for sensing and biosensing. TrAC Trends Anal. Chem. 2014, 61, 49–53. [Google Scholar] [CrossRef]
  112. Chia, X.; Pumera, M. Layered transition metal dichalcogenide electrochemistry: Journey across the periodic table. Chem. Soc. Rev. 2018, 47, 5602–5613. [Google Scholar] [CrossRef] [PubMed]
  113. Tedstone, A.A.; Lewis, D.J.; O’Brien, P. Synthesis, properties, and applications of transition metal-doped layered transition metal dichalcogenides. Chem. Mater. 2016, 28, 1965–1974. [Google Scholar] [CrossRef]
  114. Asadi, M.; Kim, K.; Liu, C.; Addepalli, A.V.; Abbasi, P.; Yasaei, P.; Phillips, P.; Behranginia, A.; Cerrato, J.M.; Haasch, R. Nanostructured transition metal dichalcogenide electrocatalysts for CO2 reduction in ionic liquid. Science 2016, 353, 467–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Zhao, C.-X.; Zhang, G.-X.; Gao, W.; Jiang, Q. Single metal atoms regulated flexibly by a 2D InSe substrate for CO2 reduction electrocatalysts. J. Mater. Chem. A 2019, 7, 8210–8217. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram showing the illustration of CO2 reduction to chemical fuels using an electrochemical cell.
Figure 1. Schematic diagram showing the illustration of CO2 reduction to chemical fuels using an electrochemical cell.
Catalysts 10 00098 g001
Figure 2. Schematics showing the properties of 2D materials.
Figure 2. Schematics showing the properties of 2D materials.
Catalysts 10 00098 g002
Figure 3. (A) Raman spectra and (B) XPS analyses of pristine MoS2 and MoS2: Nb; (C) current densities of different samples with respect to doping percentage, and (D) faradaic efficiencies (FE) of CO and H2 formation at different applied potentials. Reprinted with permission from [57].
Figure 3. (A) Raman spectra and (B) XPS analyses of pristine MoS2 and MoS2: Nb; (C) current densities of different samples with respect to doping percentage, and (D) faradaic efficiencies (FE) of CO and H2 formation at different applied potentials. Reprinted with permission from [57].
Catalysts 10 00098 g003
Figure 4. (A) TEM image showing edges inset; (B) Raman spectra; (C) Faradic efficiencies at applied potentials for CO and H2 formation; and (D) Linear sweep voltammetry (LSV) curves showing current density at initial and after 1000 CV scans, for optimized 3D TiO2@MoS2. Reprinted with permission from [76].
Figure 4. (A) TEM image showing edges inset; (B) Raman spectra; (C) Faradic efficiencies at applied potentials for CO and H2 formation; and (D) Linear sweep voltammetry (LSV) curves showing current density at initial and after 1000 CV scans, for optimized 3D TiO2@MoS2. Reprinted with permission from [76].
Catalysts 10 00098 g004
Figure 5. (A) The CO2 performance for various transition metal ions restrained on N-graphene and (B) a free energy diagram showing the formation of CO via electrocatalytic CO2 reduction. Reprinted with permission from [88].
Figure 5. (A) The CO2 performance for various transition metal ions restrained on N-graphene and (B) a free energy diagram showing the formation of CO via electrocatalytic CO2 reduction. Reprinted with permission from [88].
Catalysts 10 00098 g005
Figure 6. (A) Formation of Ni-N-RGO via chemical transformation of NTB adsorbed over RGO sheets; (B) LSV curves of different samples; and (C) faradaic efficiency of CO formation and current density versus applied potentials, reprinted with permission from [96].
Figure 6. (A) Formation of Ni-N-RGO via chemical transformation of NTB adsorbed over RGO sheets; (B) LSV curves of different samples; and (C) faradaic efficiency of CO formation and current density versus applied potentials, reprinted with permission from [96].
Catalysts 10 00098 g006
Figure 7. (A) The synthesis method of the Fe/NG catalyst; (B) free energy diagram showing mechanism of CO2 reduction to CO; and (C) the structure for Fe-N4 moieties on layered graphene showing nitrogen-substitution. Reprinted with permission from [97].
Figure 7. (A) The synthesis method of the Fe/NG catalyst; (B) free energy diagram showing mechanism of CO2 reduction to CO; and (C) the structure for Fe-N4 moieties on layered graphene showing nitrogen-substitution. Reprinted with permission from [97].
Catalysts 10 00098 g007
Figure 8. (A) Work functions of transition metal dichalcogenides (TMDCs) and Ag NPs; (B) CO2 reduction electrochemical impedance spectroscopy (EIS); (C) Cyclic voltammetry curves and (D) faradaic efficiency curves by applied potential. Reprinted with permission from [114].
Figure 8. (A) Work functions of transition metal dichalcogenides (TMDCs) and Ag NPs; (B) CO2 reduction electrochemical impedance spectroscopy (EIS); (C) Cyclic voltammetry curves and (D) faradaic efficiency curves by applied potential. Reprinted with permission from [114].
Catalysts 10 00098 g008
Table 1. The standard redox potentials for the main products of CO2 reduction in aqueous solution.
Table 1. The standard redox potentials for the main products of CO2 reduction in aqueous solution.
Half−Electrochemical Thermodynamic ReactionsProductE° RedoxEquation
CO2 + e → CO2CO2 anion radical−1.90 V(1)
CO2 + 2H+ + 2e → HCOOHFormic acid−0.61 V(2)
CO2 + 2H+ + 2e → CO + H2OCarbon monoxide−0.53 V(3)
CO2 + 4H+ + 4e → HCHO + H2OFormaldehyde−0.48 V(4)
CO2 + 6H+ + 6e → CH3OH + H2OMethanol−0.38 V(5)
CO2 + 8H+ + 8e → CH4 + 2H2OMethane−0.24 V(6)
2CO2 + 12H+ + 12e → C2H4 + 4H2OEthylene−0.41 V(7)
Table 2. The summary of various 2D nanomaterials for electrochemical CO2 reduction. Details included reactor type, example of catalysts, current density, onset potential, over potential, and the main product of CO2 reduction with faradaic efficiency and stability.
Table 2. The summary of various 2D nanomaterials for electrochemical CO2 reduction. Details included reactor type, example of catalysts, current density, onset potential, over potential, and the main product of CO2 reduction with faradaic efficiency and stability.
NoCatalystCurrent Density (versus Applied Potential)Onset Potential (V) or Overpotential (versus RHE)CO Faradaic Efficacy
(FE% versus RHE)
Stability
(versus RHE)
Reactor (Type, Volume), ElectrolyteRef, Year
1MoS265 mA cm−2−0.764 V 98% @ −0.764 V10 h @ −0.764 VTwo-compartment three-electrode electrochemical cell in Ionic liquid, 96:4 mol% water:EMI-BF4[74], 2014
2Metal doped MoS2Theoretical study[56], 2016
3MoS2 and MoSe2Theoretical study[75], 2014
4Nb-MoS2237 mA cm−231 mV @ −0.8 V 82% @ −0.8 V10 h @−0.8 V Two-compartment three-electrode electrochemical cell, Ionic liquid, CO2 saturated50:50 vol% EMIM-BF3:water[57], 2016
53D TiO2@MoS268 mA cm−2100 mV @ −0.7 V 85% @ −0.70 V14 h @ −0.70 VThree-electrode system, CO2-saturated 0.1 M Potassium bicarbonate [76], 2018
6Nickel-nitrogen-modified graphene (Ni-N-Gr)≈1.45 mA cm−2−0.5 V90% @ −0.7 to −0.9 V5 h @ −0.65 V CO2-saturated 0.1 M Potassium bicarbonate [95], 2016
7Ni2+ on N-doped graphene (Ni2+@NG)10.2 mA cm−2≈−0.40 V 92% @ −0.68 V20 h (≈2.85 → ≈−2.44 mA cm2) @ −0.58 V Three-electrode system, CO2-saturated 0.5 M Potassium bicarbonate [88], 2019
8Atomic iron dispersed on N-doped graphene (Fe/NG)2.6 mA cm−2− 0.28 V 80% @ −0.60 V10 h @ −0.60 V CO2-saturated 0.1 M Potassium bicarbonate solution [101], 2018
9Nickel- and nitrogen-doped reduced graphene oxide (Ni-N-rGO)5 mA cm−2−0.60 V 97%@ −0.8V-Two-compartment electrochemical cell (Nafion 117 membrane) CO2-saturated 0.5 M Potassium bicarbonate[96], 2018
10Nickel-N-doped graphene shell (Ni-N-GS)4 mA cm−20.7 V 93.2% @ 0.7 V20 h @ 0.7 VA homemade Teflon H-cell, CO2-saturated 0.5 M Potassium bicarbonate[97], 2017
11Ni single atoms in graphene nanosheets (Ni-G)50 mA cm−2480 mV 90% @ −0.87 V20 h @ 0.64 VA homemade Teflon H-cell, CO2-saturated 0.5 M Potassium bicarbonate[98], 2018
121D/2D N-doped carbon nanorod arrays/ultrathin carbon nanosheets3.78 mA cm−2 97 mV, @ −0.45 V 94.2% @ −0.45 V30 h, @ −0.45 VA homemade three-electrode system, CO2-saturated 0.5 M Potassium bicarbonate [102], 2019
13Ag2S/N-S-doped rGO70 μA cm−2−0.34 V87.4% @ −0.76V40 h @ −0.759 VDouble-chamber electrochemical cell with a Nafion 117 membrane 0.1 M Potassium bicarbonate[103], 2018
142D N-doped porous carbon nanosheets 1.15 mA cm−2 −0.19 V 84% @ −0.49 V8 h @ −0.60 VAr or CO2-saturated 0.1 M Potassium bicarbonate[105], 2018
15WSe2 nanoflakes (NFs)18.95 mA cm−254 mV 24% @ 54 mV27 h @ −0.364 VThree-electrode two- compartment cell ionic liquid set-up (50% EMIM-water)[114], 2016
16Metal@2D InSeTheoretical study[115], 2019

Share and Cite

MDPI and ACS Style

Hiragond, C.B.; Kim, H.; Lee, J.; Sorcar, S.; Erkey, C.; In, S.-I. Electrochemical CO2 Reduction to CO Catalyzed by 2D Nanostructures. Catalysts 2020, 10, 98. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010098

AMA Style

Hiragond CB, Kim H, Lee J, Sorcar S, Erkey C, In S-I. Electrochemical CO2 Reduction to CO Catalyzed by 2D Nanostructures. Catalysts. 2020; 10(1):98. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010098

Chicago/Turabian Style

Hiragond, Chaitanya B., Hwapyong Kim, Junho Lee, Saurav Sorcar, Can Erkey, and Su-Il In. 2020. "Electrochemical CO2 Reduction to CO Catalyzed by 2D Nanostructures" Catalysts 10, no. 1: 98. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010098

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