Efficient and Selective Electroreduction of CO2 by Single-Atom

Sep 21, 2018 - Our results show that the bonding of single metal atoms with Pc can be large enough for the ... the reduction overpotential of other TM...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of South Dakota

Article

Efficient Selective Electroreduction of CO2 by Single Atom Catalyst TM-Pc Monolayers Jinhang Liu, Li-Ming Yang, and Eric Ganz ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b03945 • Publication Date (Web): 21 Sep 2018 Downloaded from http://pubs.acs.org on September 24, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Efficient Selective Electroreduction of CO2 by Single Atom Catalyst TM-Pc Monolayers Jin-Hang Liu,1 Li-Ming Yang*,1 and Eric Ganz2

1

Hubei Key Laboratory of Bioinorganic Chemistry and Materia Medica; Key Laboratory of Material Chemistry

for Energy Conversion and Storage, Ministry of Education; Hubei Key Laboratory of Materials Chemistry and Service Failure; School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan 430074, China. 2Department of Physics, University of Minnesota, 115 Union St. SE, Minneapolis, Minnesota 55455, USA. (email: [email protected], [email protected])

Abstract: The reduction of CO2 directly to useful chemical products (such as methane, methanol, formic acid, carbon monoxide, etc.) could provide potential contributions to the reduction of the greenhouse effect that if one could find efficient, low–cost, and stable electrocatalysts. However, this remains a huge challenge. In this work, the catalytic performance of Transition Metal–Phthalocyanine (TM–Pc) monolayers as single-atom catalysts for the reduction of CO2 was investigated by spin–polarized density functional theory (DFT) calculations. Our results show that the bonding of single metal atoms with Pc can be large enough for the individual atoms to be uniformly dispersed and stable in a modified 2D– Pc monolayer. Considering the competing hydrogen evolution reaction, TM–Pc has a good hydrogen evolution inhibition. The main CRR reduction products of Sc–Pc, Ti–Pc, V–Pc and Fe–Pc monolayers are CH4. For Cr–Pc, Mn–Pc and Co–Pc HCOOH is dominant, while for Co-Pc HCOH is predicted. Except for the Sc–Pc, Ti–Pc and V–Pc monolayers (with overpotential too large, exceeding 1 V), the reduction overpotential of other catalysts are in the range of 0.017 ~ 0.819V. Mn–Pc has the lowest overpotential (0.017 V) and V–Pc has the highest overpotential (0.819 V). These were all lower than the well–studied overpotentials of the Cu(211) surface which has the best catalytic performance (except Fe–Pc). Therefore, our work may open up new avenues for the development of highly efficient catalytic materials for CO2 reduction. Keywords: CO2 reduction reaction, Electrocatalysis, Two-dimensional materials, Transition metal–phthalocyanine monolayers, Density functional theory calculations. ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 2 of 23

Page 3 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Introduction The utilization of CO2 has aroused the widespread interest of researchers around the globe. Converting CO2 into useful chemicals or fuels (such as CO, HCOOH, CH3OH, CH4, etc) can provide alternative energy stocks.[1-7] There is also the potential to reduce the amount of CO2 in the atmosphere, which could play a future role in combating global warming.[8, 9] Among the CO2 reduction methods, electrochemical reduction is the most direct and controllable method. This can operate under ambient and mild conditions. Therefore, electrochemical reduction has attracted increasing interest from both theoretical and experimental researchers.[10-12] Transition metal nanostructures and their alloys have dominated the previous work on catalytic materials for electrocatalytic CO2 reduction reaction (CRR).[13] In this group, copper have demonstrated the best catalytic performance.[14] However, the overpotential for copper can be in excess of 1V[15, 16] and more than one product is produced. Therefore, it remains a challenge to reduce the overpotential of the catalytic reaction and simultaneously improve the stability and selectivity of the catalyst. Single atom catalysis (SAC)[17] is a new frontier for heterogeneous catalysis and has been applied to a wide variety of important catalytic reactions, including CO oxidation, O2 reduction, CO2 reduction, N2 reduction and H2O electrolysis.[18-26] Compared with traditional catalysts, single–atom catalysts can have high specific activity and significantly reduce the use of precious metals. Among the large number of SACs, single–atom catalysts based on 2D materials have attracted particular attention because of their large specific surface area, exposed active sites, stable structure, and excellent catalytic activity. Transition metal phthalocyanine monolayers (TM–Pc, M = Fe, Mn, Cu)[27-29] have been successfully prepared as novel 2D materials which have been widely used in the field of spintronics[30, 31] and for gas capture.[32] In the TM–Pc monolayer, the transition metal single atoms are distributed uniformly across the substrate, so these monolayers can be used as single-atom catalysts. For example, Zhao et al. studied the Cr–Pc monolayer for catalytic CO oxidation, and the results show that the Cr–Pc monolayer exhibits excellent catalytic performance[33]. Wang et al. studied the catalytic performance of Fe–Pc monolayers for oxygen reduction theoretically.[34] These monolayers also exhibited excellent catalytic stability.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 23

Although the TM–Pc monolayer shows excellent catalytic activity in reactions such as CO oxidation and oxygen reduction, there have been no reports on the reduction of CO2. In this paper, the catalytic performance of TM–Pc monolayers for the first transition metal series on CRR was systematically studied using spin–polarized density functional theory (DFT). Our results show that the transition metal phthalocyanine (TM–Pc, M=Cr, Mn, Fe, Co, Zn) monolayers exhibit excellent catalytic activity and strong stability. Therefore, we predict that TM–Pc monolayers could contribute to the next generation of low–cost, high–stability electroreduction catalysts.

Computational methods The electronic structure and total energy were calculated using DFT with the all–electron method within the generalized gradient approximation (GGA) for the exchange-correlation term as implemented in the DMol3 code.[35, 36] The double numerical plus polarization (DNP) basis set and Perdew–Burke–Ernzerh of (PBE) functional were adopted in all calculations.[37] The DFT semi–core pseudopotential (DSPP) method[38] (which replaces core electrons with a single effective potential and introduces some degree of relativistic correction into the core) was used for transition metal atoms. To ensure high quality results, the DFT–D2 method with the Grimme vdW correction[39] was employed to accurately describe the long-range electrostatic interactions of CRR species and catalysts. The convergence threshold was 10−6 eV in energy. The Conductor-like Screening Model (COSMO) with dielectric constant of ε = 78.54 was used to simulate an H2O solvent environment throughout the whole process. Monolayer TM–Pc sheets were aligned with the xy plane. The z direction was perpendicular to the layer plane, and a vacuum space of 15Å in the z direction was used to avoid interactions between adjacent layers. The Brillouin zone was sampled with a 6×6×1 Γ−point centered Monkhorst–Pack (MP) K–point grid for geometry optimizations, while a 12×12×1 grid was used for the electronic structure calculations. The adsorption energy (Eads) for CO2 on the TM– Pc periodic monolayer was defined as: Eads = ETM-Pc-CO2-ETM-Pc-ECO2

(1)

where ETM-Pc-CO2, ETM-Pc and ECO2 are the total energy for the CO2 adsorbed on the TM–Pc monolayer, the pure TM–Pc monolayer, and the CO2 molecule, respectively. The change in

ACS Paragon Plus Environment

Page 5 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Gibbs free energy (∆G) for each of the CRR steps for the TM–Pc monolayer was evaluated according to the computational hydrogen electrode (CHE) model suggested by NØrskov and coworkers to account for the energy of a proton–electron pair in aqueous solution[40,41]. The Gibbs free energy was defined as: ∆G = ∆E + ∆EZEP - T∆S + ∆GpH + ∆GU

(2)

Where ∆E, ∆EZPE and ∆S are the DFT total energies, zero point energy and the entropy change between the products and reactants, respectively. T is the system temperature (T = 298.15 K in this paper). ∆GpH = 2.303kBTpH is the correction of the free energy due to the difference of the H+ concentration, and in this paper pH was assumed to be zero for acidic medium. ∆GU = –neU, where n is the number of transferred electrons, e is the electron charge, and U is the applied potential. The limiting potential (UL) of CRR can be obtained from the free energy change (∆GMax) by using the relation UL = -∆GMax/ne. The overpotential (ƞ) was evaluated as the difference between the equilibrium potential and the limiting (or onset) potential. Therefore, the overpotential was defined as: ƞ = Uequilibrium -Uonset.

(3)

Results and discussion Structural features of the TM–Pc monolayer Fig. 1 shows the structure of the TM-Pc material optimization diagram, optimized using a 2 × 2 supercell (see Fig. 1b). Each unit cell (shown in Fig. 1a) contains 20 carbon atoms, 4 hydrogen atoms, 8 nitrogen atoms and one transition metal atom. In the TM–Pc monolayer, each phthalocyanine binds to a transition metal atom as a planar tetradentate anionic ligand through four inwardly protruding N atoms to form a coordination bond. For all 10 transition metals of the first TM series that we considered, all of the atoms are exactly in the plane (with the exception of Sc, which protrudes from the Sc–Pc monolayer as shown in Fig. 1c). The bond length of the metal atom to the nitrogen atom rangesfrom1.93 to 2.11 Å, and the bond length decreases from Sc to Ni. Our calculations also show that most TM–Pc monolayers (except Ni, Cu and Zn) have different degrees of spin–polarized ground state, thus leaving their energy below the non–magnetic state. The local magnetic moment is mainly located at the metal atom, while the N atom combined with the metal atom shows slightly polarized antiferromagnetism. As shown in Table 1, the maximum magnetic moment is for Cr with 3.40 µB. ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 23

Fig.1 Optimized geometric structure for all TM–Pc monolayers. (a) One unit cell, and (c) 2 × 2 supercell top view. (b) Shows a side view of the Sc–Pc monolayer, which is the only case where the TM atom is up out of the plane. The black, blue, white, orange and purple balls represent C, N, H, Sc and TM atoms, respectively.

Table. 1 Energies and structural properties of various TM–Pc systems. Charge on metal atoms QM and the first shell nitrogen atom which nearest to the metal atoms (QN), and the second shell carbon atoms (QC). Spin of the metal atoms, M–N bond length RM–N are listed. TM-Pc

QM / e (Hirshfeld)

Spin (Hirshfeld)

QN / e (Hirshfeld)

QC / e (Hirshfeld)

RM-N / Å

Sc

0.7434

0.0948

-0.1773

0.0898

2.113

Ti

0.6627

-0.0002

-0.1562

0.0842

1.995

V

0.5137

2.1150

-0.1370

0.0889

1.982

Cr

0.4895

3.4034

-0.1525

0.0908

1.984

Mn

0.3451

3.0863

-0.1145

0.0916

1.965

Fe

0.2498

-1.9253

-0.1032

0.0941

1.942

Co

0.2250

-1.0338

-0.0968

0.0947

1.934

Ni

0.1310

0.0000

-0.0860

0.0964

1.932

Cu

0.3910

0.0000

-0.1391

0.0917

1.968

Zn

0.4364

0.0000

-0.1381

0.0911

1.991

ACS Paragon Plus Environment

Page 7 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Stability of single metal atoms embedded in phthalocyanine Stability is an important consideration for the performance of catalytic materials. In order to evaluate the stability of TM-Pc, we calculated the binding energy and formation energy of the materials, and also calculated the cohesive energy of the TM bulk. For the single-atom catalytic material, the stronger the binding energy of the metal atom to the substrate, the better the stability of the material, and the strong binding energy can effectively prevent the metal atoms from agglomerating. It can be seen from Fig. 2 that the absolute value of the binding energy of the single-atom catalysts are in the range of 5.83 ~ 11.21 eV (Table S1), which is significantly larger than the cohesive energy of the corresponding metal bulk (1.40 ~ 5.63 eV) (Table S1). This indicates that the metal atoms can be stably embedded on the substrate, so that metal atoms will be less likely to form clusters. The formation energy reveals the degree of difficulty in preparing catalysts, and the negative value indicates that the preparation process is an exothermic process. Fig. 2 shows that the formation energy of the materials are in the range of –4.70 ~ 0.67 eV (Table S1). The formation energy of Zn-Pc is positive while all of the others are negative. Therefore, we predict that all of the materials (other than Zn-Pc) should be feasible for experimental preparation of TM-Pc catalyst. We also performed a Hirshfeld charge analysis of the TM–Pc materials. The results (see Table 1) indicate that all of the metal atoms carry a partial positive charge, indicating that some electrons from the metal atoms have been transferred to the phthalocyanine, confirming the ionic and covalent interactions between positively charged metal atoms and phthalocyanine.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 2 Ec is the cohesive energy of TM bulk, Eb is the binding energy between TM and Pc in TM–Pc, and Ef is the formation energy of TM–Pc, where TM are the metal atoms of the first transition metal series. The first hydrogenation: selectivity for CRR vs. HER In the electrochemical CO2 reduction reaction (CRR), each step (4)–(6) involves a proton– electron pair (H+ + e−) participating in the reaction. For the first step, protonation may produce two different products, namely C*OOH or O*CHO. Under the same conditions, the catalytic material itself may also form adsorbed H (H*) by consuming one proton–electron pair (6). This may also occur during the hydrogen evolution reaction (HER) which is an undesired side reaction. Therefore, the CO2 reduction reaction is in a competitive relationship with the hydrogen evolution reaction. * + CO2 + H+ + e− → COOH*

(4)

* + CO2 + H+ + e− → OCHO*

(5)

* + H+ + e− → H*

(6)

We first compare the Gibbs free energies of the individual reaction steps that form C*OOH, O*CHO and H*. In Fig. 3 we plot the Gibbs free energy change of the first protonation step in the CO2 reduction reaction versus the Gibbs free energy change of the H2 evolution reaction for all TM–Pc. For all of the transition metals under study (except for Co), ACS Paragon Plus Environment

Page 8 of 23

Page 9 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

O*CHO is more favorable than C*OOH. In addition, the Gibbs free energy of forming C*OOH and O*CHO are smaller than forming H* (except Co–Pc forming O*CHO), which indicates that compared to forming H*, the catalysts are more likely to form C*OOH and O*CHO. However, once the active site of the catalysts are occupied by C*OOH and O*CHO, it will be difficult to form H*, which will cause the hydrogen evolution reaction to be suppressed. So above the dotted line, the reaction is dominated by hydrogen evolution, while below the line, the reaction is dominated by the desired CO2 reduction. The free energy barriers for the first hydrogenation step for Ni and Cu are higher than 1 eV, and so the CRR process will be difficult. Therefore, we will focus on the CRR reaction pathways and product selectivity for Sc, Ti, V, Cr, Mn, Fe, Co and Zn in the remaining discussion. In this paper, we only consider the formation of single–carbon (C1) products, because there are no additional active metal sites around the metal atoms, thus preventing further C–C coupling reactions between the reaction intermediates which would form multi-carbon compounds.

Fig. 3 Gibbs free energy change of the first protonation step in the CO2 reduction reaction (CRR) and H2 evolution reaction (HER) on the TM–Pc. Catalysts below the dotted line are CRR selective.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Reaction pathways for CO2 electrochemical reduction

Scheme 1 Most likely reaction pathways for the electroreduction of CO2 on TM–Pc.

Fig. 4 Key reaction intermediates in CRR. The columns are labeled by the numbers of proton– electron pairs (H+ + e−) transferred to each CO2. Scheme 1 presents a reaction path diagram for the electrochemical reduction of carbon dioxide to a variety of C1 hydrocarbons, including the primary 2e− reduction products (CO and HCOOH) through 8e− reduction products (CH4). Fig. 4 shows key reaction intermediates during the CRR process. In the following, the lowest energy pathways for the reaction of Sc– ACS Paragon Plus Environment

Page 10 of 23

Page 11 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Pc, Ti–Pc, V–Pc, Cr–Pc, Mn–Pc, Fe–Pc, Co–Pc and Zn–Pc will be discussed in terms of the change in Gibbs free energy, overpotential, and product selectivity.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 5 Gibbs free energy profiles for CRR along the most favorable path ways for (a) Sc–Pc, (b) Ti–Pc, (c) V–Pc, (d) Cr–Pc, (e) Mn–Pc, (f) Fe–Pc, (g) Co–Pc and (h) Zn–Pc at zero potential. The free energy zero is set as that of a CO2 molecule in the gas phase and with a clean catalyst surface.

As mentioned above, the first step of hydrogenation will form two intermediates, C*OOH (path1) and O*CHO (path 2). If hydrogenation is continued, C*OOH → C*O + H2O → CO. On the other path, O*CHO → HCO*OH → HCOOH. CO and HCOOH are 2e−reductions of CRR. Whether CO or HCOOH will be produced is primarily determined by their adsorption energy on the catalyst. If the adsorption energy of CO and HCOOH on the catalysts are too large, CO and HCOOH reaction products will not be easily desorbed from the surface of the catalysts, but continue to be reduced as a reaction intermediate. When CO and HCOOH continue to be reduced, the 3 different intermediate states C*HO, C*OH and O*CH are produced. Fig. 5 shows that compared to C*OH and O*CH, the values of ∆GC*HO are lower than ∆GC*OH and ∆GOC*H, so C*O and O*CHOH will undergo path C*O → C*HO and O*CHOH → C*HO + H2O to get the intermediate C*HO. As can be seen from Table S2, the adsorption energies of CO and HCOOH on Sc–Pc, Ti–Pc, V–Pc and Fe–Pc are in the range of −1.835 ~ – 1.140 eV and –1.70 ~ –0.762 eV, respectively, indicating that Sc–Pc, Ti–Pc, V–Pc and Fe–Pc are sufficiently strong to adsorb CO and HCCOH, so that CO and HCOOH will continue to be reduced as reaction intermediates. For the process * + CO2 → C*HO, Fig. 5a, 5b, 5c, and 5f show that the maximum free

ACS Paragon Plus Environment

Page 12 of 23

Page 13 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

energy change (∆Gmax) of path 1 is lower than the maximum change of path 2. Therefore, the optimum reaction path for the four catalysts Sc–Pc, Ti–Pc, V–Pc and Fe–Pc are path 1. The adsorption energies of CO and HCOOH on Cr–Pc, Mn–Pc, Co–Pc and Zn–Pc are in the range of –1.141 ~ –0.208 eV and –0.433 ~ –0.345 eV, respectively (Table S2). Except for Zn–Pc, the ability of the three catalysts to adsorb HCOOH is weaker than that of CO. In addition, by comparing the changes of the free energy (shown in Fig. 5d, 5e, and 5h), the optimum reaction path for Cr–Pc, Mn–Pc and Zn–Pc are path 2. Since the adsorption capacity of Cr–Pc, Mn–Pc and Zn–Pc for HCOOH is relatively weak, the HCOOH reaction product is easily desorbed from the surface of the catalysts. As shown in Fig. 3, the ∆GO*CHO of Co–Pc is greater than ∆GH, so path 1 is more suitable than path 2 for Co–Pc. However, the adsorption capacity of Co-Pc for CO is not sufficiently weak, which allows CO to continue to react as an intermediate. C*HO continues to be reduced to give the intermediate O*CH2. If the adsorption capacity of the catalyst for O*CH2 is sufficiently small, HCHO will become a reduction product desorbed from the catalyst (C*HO → O*CH2 → * + HCHO). The adsorption energies of Sc– Pc, Ti–Pc, V–Pc, Fe–Pc and Co–Pc are –1.708, –1.397, –1.269, –0.972 and –0.456 eV, respectively. Compared to the other four catalysts, the adsorption capacity of HCHO on the surface of Co–Pc is very weak, which makes the catalytic product of Co–Pc mainly HCHO. Since the adsorption energy of HCHO on Sc–Pc, Ti–Pc, V–Pc, Fe–Pc are big enough, it will continue to be reduced. For path 3, O*CH2 → O*CH3 → O* + CH4, the product CH4 can be produced. The product CH3OH can be produced by path4: O*CH2 → O*CH3 → CH3O*H → * + CH3OH. Fig. 5a, 5f show that ∆GO* is greater than ∆GCH3O*H, whereas Figure 5b, 5c shows that ∆GCH3O*H is greater than ∆GO*, so Sc–Pc and Fe–Pc primarily react according to path 4, while Ti–Pc and V– Pc primarily follow path 3. The adsorption energy of CH4 on Sc–Pc and Fe–Pc are only –0.032 and –0.150 eV, so the products of Sc–Pc and Fe–Pc catalytic reduction of CO2 are CH4. The adsorption energy of CH3OH on Ti–Pc and V–Pc are –1.854 and –1.07 eV, which indicates that CH3OH is difficult to desorb as a product. CH3OH continues to be reduced and undergoes the path CH3O*H → O*H + CH4 to obtain the product CH4.[42] The adsorption energies of CH4 on Ti–Pc and V–Pc are –0.041 and –0.108 eV, respectively, so the products of Sc–Pc and Fe–Pc

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 23

catalytic reduction of CO2 are still CH4.

Table 2 The calculated potential determining steps (PDS), M–O bond length R[M–O] of the intermediate (O*H, O*CHOH, O*CH2 or O*CHO), limiting potentials (UL/V) and overpotentials (ƞ/V) of the CRR. TM–Pc

PDS

RM-O / Å

UL / V

ƞ/V

Sc–Pc

O*H → * + H2O

1.906

–1.707

1.876

Ti–Pc

O*H → * + H2O

1.805

–1.771

1.940

V–Pc

O*H → * + H2O

1.794

–1.248

1.417

Cr–Pc

O*CHOH → * + HCOOH

2.202

–0.293

0.043

Mn–Pc

O*CHOH → * + HCOOH

2.295

–0.267

0.017

Fe–Pc

O*CHOH → C*HO + H2O

1.974

–0.650

0.819

Co–Pc

O*CH2 → * + HCHO

2.223

–0.438

0.368

Zn–Pc

* + CO2 → O*CHO

2.028

–0.728

0.478

The overpotential is an important parameter for the evaluation of the performance of catalysts. To achieve high reactivity and selectivity of carbon dioxide reduction, the catalyst should have less negative limiting potentials. The larger the negative limit potential value is, the smaller the overpotential. We will discuss the overpotential of the eight TM–Pc monolayers in the catalytic reduction of CO2.The rate–determining step determines the limiting potential of the reaction, and the limiting potentials of the different monolayers are listed in detail in Table 2. The results show that the overpotential of the reduced product to CH4 is higher than other reduction products. Among the four catalysts in which the reduction product is CH4 (Sc–Pc, Ti–Pc, V–Pc and Fe–Pc), the lowest overpotential is Fe–Pc (0.819 V), and the rest of the overpotential are greater than 1 V. Therefore, Sc–Pc, Ti–Pc, V–Pc and Fe–Pc require a higher overpotential for the reduction product CH4. The reduction products of Cr–Pc, Mn–Pc and Zn– Pc are HCOOH, and their overpotentials are 0.043 V, 0.017 V and 0.478 V, respectively, which indicates that Cr–Pc, Mn–Pc and Zn–Pc can catalyze the reduction of CO2 at a low

ACS Paragon Plus Environment

Page 15 of 23 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

overpotential. Similarly, the reduction product of Co–Pc is HCHO and its overpotential is 0.368 V, indicating that the reaction can also occur at a lower overpotential. In general, except for Sc–Pc, Ti–Pc and V–Pc which have a higher overpotential, the overpotentials of other catalysts are comparable (even lower) than the most active step surface Cu(211) (ƞ = 0.77 V) and the most active metal surface, Pt(111) (ƞ = 0.46 V).[43] Furthermore, the overpotentials of our designed catalysts are comparable or even smaller than other experimentally known promising catalysts, such as, modified Cu electrode from the reduction of thick Cu2O films (ƞ = 0.5 V),[44] high surface area tin oxide nanocrystals (0.34 V),[45] monodisperse Au nanoparticles (0.26 V),[46] nitrogen-doped carbon nanotubes (0.54 V),[47] carbon nanofibre (0.17 V),[48] and nanoporous silver (