Reaction Selectivity for Plasmon-Driven Carbon Dioxide Reduction on

2 days ago - Density functional theory is employed to investigate the plasmon-driven CO2 reduction at the active sites of metallic silver clusters. Th...
0 downloads 0 Views 3MB Size
Subscriber access provided by UNIV OF LOUISIANA

C: Plasmonics; Optical, Magnetic, and Hybrid Materials

Reaction Selectivity for Plasmon-Driven Carbon Dioxide Reduction on Silver Clusters: A Theoretical Prediction Xia-Guang Zhang, Yuxiu Liu, Chao Zhan, Xi Jin, Qijin Chi, De-Yin Wu, Yi Zhao, and ZhongQun Tian J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b01448 • Publication Date (Web): 12 Apr 2019 Downloaded from http://pubs.acs.org on April 12, 2019

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 41 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

The Journal of Physical Chemistry

Reaction Selectivity for Plasmon-Driven Carbon Dioxide Reduction on Silver Clusters: A Theoretical Prediction Xia-Guang Zhang†, Yuxiu Liu†, Chao Zhan†, Xi Jin†, Qijin Chi‡, De-Yin Wu*†, Yi Zhao*†, Zhong-Qun Tian†

†State

Key Laboratory of Physical Chemistry of Solid Surface, Collaborative

Innovation Center of Chemistry for Energy Materials, and Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen, 361005, P. R. China ‡Department

of Chemistry, Technical University of Denmark, DK-2800 Kongens

Lyngby, Denmark

Corresponding Author E-mail: D.Y. W. ([email protected]); Y. Z. ([email protected]).

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Abstract: Density functional theory is employed to investigate the plasmondriven CO2 reduction at the active sites of metallic silver clusters. The results predict that CO2 prefers to adsorb at the bridge site of silver clusters and the C–O bond is difficult to break due to a high activation barrier at electronic ground state. However, as the photo-generated plasmon energy of silver nanoparticles matches with the dissociation energy of the C–O bond, the CO2 easily dissociates into CO and an adsorption oxygen atom. Moreover, our calculated results demonstrate that the generated CO strongly adsorbed on silver clusters is favorable to its succeeding reduction reaction with hydrogen to CH3OH and CH4. The reaction barrier of the generation of CH3OH is lower than that of the formation of CH4, because the proton combines with the carbon more easily than with the oxygen atom at the initial reaction step. It well accounts for the experimental observation that CH3OH can be formed from the CO2 reduction on silver nanoparticles under visible light irradiations.

2 ACS Paragon Plus Environment

Page 2 of 41

Page 3 of 41 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

The Journal of Physical Chemistry

Introduction Carbon dioxide reduction to CO and hydrocarbons is becoming a prevalent research field since it is significant to complete carbon cycle for carbon sequestration and generating energy-rich products. In general, CO2 is an inert molecule (ΔGf0 = –394.4 kJ/mol),1 which requires considerable energy for chemical activation. Thus, multiple steps of electron and proton transfers on catalytic surfaces are difficulty of carbon dioxide reduction reaction (CO2RR).2 For electrochemical CO2RR on different electrodes which has been systematically studied by Hori groups,3 the reduction potentials and product distribution strongly depend on experimental conditions, electrolytic solutions and electrode materials. Catalyst morphology4-6, alloying7-8 and doping9 are possible to influence on the reaction activation barrier, reaction efficiency and selectivity. Recently, more and more attentions have been paid to use renewable energy like solar energy to assist photocatalytic or photoelectrochemical process of CO2RR10-11 and tremendous progress has been made. Among the various photo-catalysts, TiO2 as a semiconductor is commonly used due to its high reactivity.12-15 But, the large band gap (~3.2 eV) of TiO2 limits its application in ultraviolet light region,16 which contains only 5% of total solar energy. These drawbacks greatly limit the applications of TiO2 in photocatalytic CO2RR. Coinage metal nanoparticles, such as, gold, silver, and cooper have a good optical property due to localized surface plasmon resonances (LSPR), which 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

can enhance the photocatalytic efficiency under visible light irradiations.17-22 The LSPR can significantly induce the formation of hot carriers participating in CO2RR.23-29 Multi-electron transfer can happen in LSPR mediated CO2 reduction, where the C–O bond of chemisorbed CO2 on metal surfaces is directly broken by the LSPR effect.25-26,

30-33

In previous studies, silver

nanoparticles were often used to study LSPR reactions which could increase electrochemical reaction rate and reduce the energy barrier from CO2 to CO.3437

For CO2 electroreduction on a silver electrode, CO is the main product, and

the Faradaic efficiency is above 81.5%.3 Some mechanisms for the formation of CO on silver surfaces have been proposed. One is CO2 in the chemisorption state directly dissociates into CO,38 another one involves CO2 simultaneously obtains a proton and an electron produces CO via intermediate *COOH.3, 39-40 Recently, researchers also found that CO2 could be reduced to hydrocarbon and alcohol compounds through *CHO and *COH intermediates on nanostructured silver surface under visible light,32 and *CHO is a thermodynamically stable structure on silver nanoparticle.41 While the reaction diagram of LSPR induced CO2RR is perplexing, especially, how to understanding photo-generated hot electron transfers from metal to CO2. In this work, we conduct density functional theory (DFT) to first investigate the interaction between CO2 and silver clusters, simulate active sites and explore the CO2 reduction mechanism on silver electrodes. We find that excited plasmons on silver clusters can promote hot electron transfer to LUMO of CO2, 4 ACS Paragon Plus Environment

Page 4 of 41

Page 5 of 41 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

The Journal of Physical Chemistry

and then relax to the charge transfer state, due to the strong orbital coupling of silver clusters and CO2. Under light excitation, the strong LSPR effect of silver nanoparticles is able to efficiently promote the CO2 adsorption and subsequent dissociation to CO. Then CO is reduced to CH3OH and CH4 due to the high activity of silver clusters. The schematic of reaction path and the characteristic vibrations of Raman spectra are analyzed and drawn for a better understanding of the CO2RR on silver nanoparticles. This will be helpful to identify the reaction intermediates and speculate reaction mechanisms. Computational Methods To gain deep insight on the interaction of CO2 on silver electrodes, metallic cluster models were adopted according to previous work of Pereiro’s group,42 where CO2 chemisorbed and physisorbed at active sites of silver surfaces (Figure S1 and S2). All DFT calculations were performed by Gaussian 09 program.43 The hybrid exchange-correlation functional PBE1PBE44 was employed to optimize all structures of the ground states and transition states. For H, C, and O atoms, the basis set Aug–cc–pVTZ45 was adopted here. For the metal atom Ag, the outer and inner shell valence electrons are described by the effective pseudo potential double–ζ (LANL2DZ) basis set. The solvation model of density (SMD) was adopted for all calculations in solvent water with a dielectric constant of 78.4.46 The tight option of the cutoffs on forces and step size are used to determine convergence. Vibrational frequency calculations show that all optimized structures are minima on the potential energy surfaces 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 41

of electronic ground states. All the transition states have only an imaginary frequency, and the intrinsic reaction coordinate (IRC) calculations prove the validity of transition state. Nature bond orbital (NBO) analysis47-48 has been used to investigate the bonding mechanism between adsorbent and metal clusters. Absorption spectra and nature transition orbital (NTO) analysis could be obtained from time–dependent DFT (TDDFT) calculation based on optimized structures of electronic ground states. The excited state energies calculated by using TDDFT are known to have errors dependent on all functionals.49 In order to verify that the results of PBE1PBE are consistent with other functionals and basis set, we have repeated the calculations above using the other commonly used functional B3LYP and basis set 6-311+G** for C and O atoms and LANL2DZ basis set for Ag atoms.50 The results are provided in the Supporting Information, showing that the two functionals have similar performance. All activation energies have been corrected by considering zeropoint energies and enthalpies. The reduced two–state model was used to predict the electronic coupling by the overlap of the two diabatic states, which were assumed to be the LUMOs of the two isolated neutral molecules.51-53 The electronic coupling is 1 h12  (h11  h22 )S12 2 VDA = 2 1  S12

(1)

where hij = , Sij = , and the ϕi and ϕj are corresponding orbital wave functions of molecule and the silver clusters respectively. 6 ACS Paragon Plus Environment

Page 7 of 41 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

The Journal of Physical Chemistry

The absolute Raman intensities were calculated by using the differential Raman scattering cross section.54 The calculated harmonic frequencies were scaled by scaling factors of 0.981 for below 2000 cm–1, and 0.967 for above 2000 cm–1.55 In this case, the Raman scatting intensity was calculated using54 IRaman

 2π  

2 4   dα 2  dγ    ω0  ωi  h  45    7   45 8π 2cωi 1  exp  hcωi / kBT    dQi  dQi     

=

 2π 

4

 ω0  ωi  h Si 45 8π cωi 1  exp  hcωi / kBT  4

4

(2)

2

where h, c, kB and T are the Planck constant, light speed, Boltzmann constant, and Kelvin temperature, respectively. Si is the Raman scattering factor (in Å4/amu) that can be calculated using Gaussian 09 at the equilibrium geometry, and it is only an expression of derivatives is the static isotropic and anisotropic polarizabilities with respect to the given normal coordinate. Here ω0 and ωi denote the frequency (in cm–1) of the incident light and vibrational frequency of the ith mode. The simulated Raman spectra were presented in terms of a Lorentzian expansion with a line width 10 cm–1. The differential Raman scattering cross section values of different vibrational modes were calculated from the Raman scattering factor under the double-harmonic approximation, with a specific excitation wavelength of 532 nm at room temperature. Results and Discussion Adsorption Structure of CO2 at Silver Nanoparticles. For CO2RR on silver electrodes, firstly, an electron transfers from silver to the CO2, forming the CO2–. This process needs a high negative electrode potential about -1.9 V vs. normal 7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 8 of 41

hydrogen electrode (NHE).56 In general, the formation of CO2– is very difficult on low-index single crystal surface of silver with a weak surface activity.57

Table 1. Structural parameters, NBO charges (qNBO, a.u.), and adsorption energies (Eads, eV) of CO2 adsorbed on silver clusters.a Physisorptionb Ag2 Ag4 Ag6 Ag8 Ag10 Ag12 Ag14 Ag16 Ag18 Ag20

Chemisorption

NBO

Eads

C=O1c

C=O2d

NBO

Eads

–0.009 0.016 0.019 0.031 0.044 0.030 0.029 0.027 0.048 0.051

–0.14 –0.16 –0.17 –0.22 –0.16 –0.16 –0.17 –0.17 –0.17 –0.18

1.237 1.245 1.250 1.252 1.257 1.258 1.260 1.259 1.251 –

1.237 1.276 1.280 1.293 1.283 1.289 1.289 1.294 1.270 –

–1.058 –1.178 –1.091 –1.201 –1.216 –1.237 –1.275 –1.257 –1.030 –

–2.26 –3.40 –3.00 –3.90 –3.51 –4.03 –4.37 –4.15 –3.25 –

a) The unit of bond length is Å; b) the bond length of C=O for physisorption CO2 is 1.158 Å, and the bond length of C=O for free CO2 is 1.157 Å (the experimental value is 1.163 Å58); c) C=O1, the O atom is far away from silver surfaces; d) C=O2, the O atom is close to silver surfaces.

According to the NBO analysis of adsorbed CO2 molecules, the physisorption CO2 moiety possesses zero charge, whereas the chemisorption CO2 carries one negative charge, (i.e., CO2–) (Figure S1, Table 1). The geometric structure of free CO2– is a bending configuration, in which the OCO is 134.6º and the C=O bond lengths are 1.232 Å. For the CO2– adsorbed on silver clusters, the geometric structure parameters are listed in Table 1. CO2 prefers to bind at the 8 ACS Paragon Plus Environment

Page 9 of 41

bridge site of the Ag2 cluster, while for other silver clusters, the C and O of CO2 adsorbed at silver top sites and the C–O bond is parallel to the Ag–Ag bond. When CO2 adsorbs on silver clusters, the interaction between CO2 and silver clusters would lead to the change of geometric structures of clusters.59 The calculated Ag–Ag bond length of clusters (~3.0 Å) with adsorbed CO2 is longer than that of Ag(100) and Ag(111) surfaces (2.889 Å).60 It shows that decreasing coordination number and increasing Ag–Ag bond length of silver clusters can increase the surface activity of silver clusters.61 50

Raman Intensity/10-29 cm2/sr

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

The Journal of Physical Chemistry

1152

700

40

30

20

740

1164

744

1184

1356

Ag14

738

1174

1360

Ag12

736

1188

742

1160

712

10

734 612

0 400

Ag18

1414 1352

600

800

Ag16

1376

Ag10

1380

Ag8

1168

1412 Ag6

1178

1440 Ag4

1162

1585 Ag2

1000

1200

1400

1600

-1

Raman Shift /cm

Figure 1. Calculated Raman spectra of chemisorbed CO2 on different silver clusters. The Raman intensity is expanded on the basis of the Lorentzian line shape with the linewidth of 10 cm–1. For free CO2–, our calculated results show that it has three characteristic vibrational frequencies at 719, 1317, and 1654 cm–1, in accordance with the experimental data at 714,62 1298,63 and 1658 cm–1.62,

64

Figure 1 shows

calculated Raman spectra of CO2– adsorbed on silver clusters, three vibrational 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

frequencies appear in agreement with the experimental values at 750, 1200, and 1410 cm–1.65-66 Furthermore, for the vibrational frequency at 1410 cm–1, it can be attributed to the C=O stretching vibration with the O atom far away the silver clusters. The vibrational frequency at 1160 – 1190 cm–1 with a large Raman activity belongs to the C=O bond stretching vibration on silver surfaces. The vibrational frequency near to 700 cm–1, mainly comes from the O–C–O bending vibration. The above results indicate that the vibrational frequencies and Raman activities of the chemisorption state CO2 can be used to identify the adsorption structures. Therefore, observed Raman spectra provide the information to recognize surface chemical adsorption states and active sites of CO2. Plasmon Driven CO2 Adsorption Dissociation on Silver Clusters. For CO2RR, from the chemisorption CO2 to CO two reaction paths: the direct twoelectron dissociation path38 and two-electron two-proton reaction path3 are proposed. The activation energy barrier of the former is above 2.0 eV (Table S2). The other reaction path is that CO2 is reduced into CO through a *COOH intermediate.59 The reaction energy barrier from adsorbed CO2 to *COOH is above 1.0 eV (Table S3). Both of them need a very negative cathodic polarization in electrochemical interfaces.34 The surface plasmon excitation is the collective excitation of conductance electrons on metal surface, and the photogenerated hot-carriers from LSPR relaxation were proposed to facilitate the C–O bond break of adsorbed CO2 in illuminated on silver and aluminum10 ACS Paragon Plus Environment

Page 10 of 41

Page 11 of 41 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

The Journal of Physical Chemistry

cuprous oxide.30,

38

The conceptual mechanism of LSPR-driven CO2

dissociation is that the hot electron will transfer from metal to the LUMO of adsorbed CO2. In order to understand the detail process of plasmon driven CO2RR, we performed TDDFT calculations to investigate the orbital energy levels of CO2 and silver clusters, and the plasmon-excited charge transfer process. An excellent absorption of visible light on plasmonic nanostructures may be a major reaction channel for the plasmon driven surface catalysis mechanism. 20, 67

Figure S9 shows the calculated absorption spectra of different adsorption

configurations. The simulated results of absorption spectra of silver clusters are in accordance with the results of previous studies.68-69 For all physisorption structures, all the orbitals are localized on the silver clusters in a low energy region (λex > 500 nm). It doesn’t exist any proper orbital for the charge transfer from the silver clusters to CO2. Whereas in the higher photonic energy region (λex < 500 nm), the SPR excitation appears, which originates from excited sp electrons with large oscillator strengths.68-69 In addition, the charge transfer from silver clusters and CO2 can be found in this region. As shown in Figure 2, we analyze the interaction of the Ag18 cluster with CO2. In the energy region of physisorption structures, we notice each of the orbital energies of LUMO+8 ~ LUMO+11 of Ag18 matches well with the orbital energy of the LUMO of CO2 (Table S4), and the orbital energy of CO2 conforms to data in the previous studies.15,

70-71

Meanwhile, the electron excitation state is associated to 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

LUMO+8 ~ LUMO+11 orbitals have large oscillator strengths. Thus, we propose LUMO+8 ~ LUMO+11 orbitals are related to the plasmon-driven electron transfer mechanism. To prove our hypothesis, we further investigate the coupling strength of wave functions between CO2 and Ag18 based on the two–state model.53 Herein, the two coupled wave functions belong to the unoccupied orbital of silver clusters and the LUMO of CO2. The strength of orbital coupling is strong, about 106.2 meV (Table S5), leading to the excited hot-electrons on silver clusters can easily transfer to the LUMO of CO2 (Figure S10). 1

0

L+11 L+10 LUMO L+9 coupling: L+8 106.24 meV

L+9

-1

LUMO

HOMO

HOMO

absorption f=1.2136 Eabs=3.23 eV

-3

LUMO

absorption f=0.8065 Eabs=3.2 eV

-2

absorption f=1.1269 Eabs=3.3 eV

Orbital Energy (eV)

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 12 of 41

HOMO

-4 Ag18

CO2

Physical Adsorption

Cluster

Chemical Adsorption

Figure 2. Orbital coupling between the silver cluster and CO2. The red arrow, orange arrow, and violet arrow represents the highest oscillator strength transition of the Ag18 cluster, CO2 physisorption structure, and CO2 chemisorption structure, respectively.

From the above analysis, we find that the hot-electron can transfer from silver 12 ACS Paragon Plus Environment

Page 13 of 41 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

The Journal of Physical Chemistry

clusters to the LUMO of CO2 due to a large orbital coupling. Then, the configuration of CO2 obtained one electron becomes nonlinear through geometric relaxation. Moreover, the transition energy of the excited state with a large oscillator strength of CO2 chemisorption structure is 3.2 eV (Figure 2 and Table S7). Surprisingly, this just matches to the energy of the charge transfer excited state of CO2 physisorption structure (Figure 2 and Table S6). For the CO2 chemisorption state, we analyze the charge transfer between CO2 and silver clusters according to the nature transition orbital. Upon photoexcitation, the photo-generated hot electron can transfer into the C–O bond, resulting in the C–O cleavage (Figure 2). We also notice that the required energy to break the C–O bond of CO2 chemisorbed on silver is about 2.23 ~ 2.45 eV. However, the energy is in agreement with that of silver surface plasmons which comes from hot electrons formed from the decay of plasmon excitation on silver nanostructures. Thus, it implies that an efficient excited electron can be transferred to the C–O antibonding orbital, leading to the C–O bond cleavage.18, 72

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Figure 3. Mechanism of the photo-generation hot electron transfer from a silver cluster to CO2. The black line and the blue line represent the density of state of Ag18 cluster and CO2, respectively. The energy level is relative to the vacuum level.

Figure 3 displays the detailed picture of LSPR-driven CO2 adsorbed on silver clusters. It is drawn based on our calculated results and several previous studies of the chemical process induced by hot electrons generated by surface plasmons.22, 73-75 We can see that LSPR-driven hot-electron arisen from a silver nanoparticle is transferred to the LUMO of CO2 in Ag*–CO2 (* denotes the excited state), causing the formation of a charge-transfer state . Then, [Ag+– CO2–]* relaxes to the corresponding charge transfer ground state. LSPR-driven the second hot electron breaks the C–O bond. Finally, the new surface species CO and oxygen atom adsorbed on silver surfaces are formed. It is worth to note that the surface hydrogen atoms from H2O reduction reactions can consume surface adsorbed oxygen atom and then produce the hydroxide anions into aqueous solution.30 Therefore, hot-electron induced by LSPR can promote adsorbed CO2 dissociation into CO on silver clusters. Furthermore, we find that CO can be strongly adsorbed on nanostructures (Table S8) and the adsorption energy is close to that of on copper surfaces.76-77 It is favorable for the further reduction of CO2 on silver nanostructures. 14 ACS Paragon Plus Environment

Page 14 of 41

Page 15 of 41 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

The Journal of Physical Chemistry

Reduction of Adsorbed CO on Silver Clusters. Previous studies indicated that the product of CO could be further transformed into methanol and methane compounds on silver nanoparticles under light irradiations.16, 32, 78 Herein, we calculate thermodynamic energies of the CO2RR to explore the feasible mechanism and the reaction selectivity. We find that strong binding of CO on silver clusters results in the reduction of CO to C1 products. And our calculated results also show that when the surface coverage is large, the binding energy of CO decreases, in agreement with previous studies.79-80 Once a single CO is binds individual Ag atoms and it will be isolated from adjacent Ag atoms. Thus, the C–C coupling is difficult to achieve.81 Figure 4 presents different reaction paths of *CO reduction. We first focus on the reaction from *CO to *CHO in path I. The activation energy of the elementary reaction is 0.53 eV with respect to the reactants *CO and H+. The intermediate *CHO has two parallel branching reaction paths to *CH2O and *CHOH with activation energies about 0.45 and 1.01 eV, respectively. It shows that at the parallel steps the *CHOH formation was a harder step than the *CH2O formation. In such a case, the H+ prefers to bind to C atom. The further reduction reactions of *CH2O also follow a similar rule. The activation energy for *CH3O is only 0.15 eV, much smaller than 1.22 eV for *CH2OH. Thus the final stage is the further reduction of *CH3O to CH3OH with an activation energy about 0.87 eV. Comparing with the activation energy of every elementary reaction, we can see that the reaction from *CO to CH3OH is the rate15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

determining step for the formation of methanol.

H+ 1.4 4e V

OH CH

0.29 eV H+

+

H eV 27 1.

OH CH2

OH CH3

O CH2

H+ 0.45 eV

H+ 0.15 eV

0. H + 87 eV

1. H + 22 eV

V 1e

H+

2.86 eV

V

+

O CH

6e

CH4

eV

O C

H+

H

+

O

0.3

H+

72

C

CH3

0.

O

H+

Path III

H+ 0.93 eV

eV

OH C

Path II

CH2

1.12 eV

1.0

Path I

H+

CH

H+ 1.23 eV

07 1.

H+ 2.30 eV

C

H+ 0.53 eV

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 16 of 41

O CH3

Figure 4. Three reaction paths for CO reduction on Ag18 cluster. The green arrows (Path I) and blue arrows (Path II) present two different reaction path with the same reduction product, CH3OH. And the red arrows are a reaction path from CO to CH4 (Path III). The H2O product is omitted in the scheme for simplicity. The numbers represent reaction activation energies at 298.15 K and 1 atm.

The Path II presents the reduction of CO through intermediate *COH, *CHOH, and *CH2OH and finally converted to CH3OH. The first step from CO to *COH intermediate has a very large activation energy barrier about 2.86 eV. Under the circumstances, we think that the LSPR effect plays an indispensable role in overcoming the high energy barrier and facilitating this reaction. Once *COH is formed, it can be further converted into two intermediates, *CHOH and *C, with different reaction barriers about 0.36 and 2.30 eV, respectively. It suggests that the majority of *COH is reduced to *CHOH rather than *C. As shown in Figure 16 ACS Paragon Plus Environment

Page 17 of 41 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

The Journal of Physical Chemistry

4, the activation energies of several steps from *COH to CH3OH along the Path II are 0.36, 0.29, and 0.72 eV, which are much smaller than the first step. So, along the way with CH3OH as the product, the rate-determining step is the formation of *COH. In all reduction steps on silver clusters of Paths I and II, our calculated results as well as previous studies show that surface hydrogen prefers to binding to surface C atom other than O atom because of lower energy barriers.81-82 For path III, CH4 is the final reduction product, But the condition is a little bit different from Path I and Path II, where all elementary reactions have a large activation energy barrier. The smallest barrier is predicted about 0.93 eV along the reaction path. Therefore, it is very difficult for the CH4 formation via ground state reaction of adsorbed CO on silver surfaces.81 Additionally, an alternative scheme which considers the same excited intermediates of Paths II is also proven to be impractical, as shown in Figure 4.83 Simulated Raman Spectra of Key Adsorption States. In order to clearly assign vibration fundamentals of surface adsorption species, we classify the vibrational frequencies as three regions. Concretely, the vibrational frequencies in the region of 200 – 700, 900 – 1600, and 2500 – 3500 cm–1 are denoted as low, middle and high frequencies, respectively. There is no doubt that the high frequency peaks are ascribed to the C–H (2500 – 3000 cm–1)84 and the O–H ones (larger than 3000 cm–1)85. However, the situation is more complicated in low and middle frequency regions, which are usually termed as fingerprint 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry

region of surface adsorption species. So we will particularly pay attention to the low and middle frequency regions for further understanding of the reaction mechanism. 424

a

1056

1256

*CH3

528 496 546 464

290 346 422

1

322

2

*CH *CH2

454

3

0

1562

*CH3O

1544

1362

1000 634

1

218

2

*CHO *CH2O

1402

b 3

320 392

1320

0

c

3184

1448

1086

914

1

420

390

2

*COH *CHOH *CH2OH

1218

970

3

1248

Raman Intensity/10-28 cm2/sr

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 18 of 41

0 200 400 600 800 1000 1200 1400 1600

30003300

-1

Raman Shift /cm

Figure 5. Simulated Raman spectra of the intermediates for CO reduction on Ag18 cluster. a) CO reduction to CH4 through *CH, *CH2, and *CH3; b) CO reduction to CH3OH through *CHO, CH2O, and CH3O; c) CO reduction to CH3OH through *COH, *CHOH, and *CH2OH. The excitation wavelength of 532 nm was used here with a Lorentzian line width of 10 cm–1.

Figure 5a displays the simulated Raman spectra of intermediates for the 18 ACS Paragon Plus Environment

Page 19 of 41 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

The Journal of Physical Chemistry

process of CH4 formation. In the middle frequency region, two characteristic frequencies of 1256 and 1056 cm–1 come from the bending vibrations of *CH2 and *CH3.84, 86-87 And in the low frequency region, two main characteristic peaks of *CH3 are 546 and 528 cm–1, which are considered as a consequence of the C–H rocking vibration. Besides, the characteristic peaks *CH2 at 523 and 464 cm–1 come from the C–H rocking and bending vibrations, respectively. As for the intermediate *CH, its C–H bending vibration reflects on the characteristic peaks of *CH at 454 and 496 cm–1. The simulated Raman spectra presented in Figure5b corresponds to CO reduction to CH3OH through *CHO intermediate (Path I). We note that the characteristic frequencies in the middle frequency region were useful to distinguish the surface species.32 For *CHO intermediate, the peaks at 1544 and 1362 cm–1 come from the C–O stretching vibration and the Ag–C–H bending vibrations, respectively. And the *CH2O intermediate possesses three Raman peaks at 1562, 1402 and 1000 cm–1. The first two peaks can be ascribed to the O–C–H bending vibrations, C=O stretching vibration and Ag– C–H bending vibration, respectively. For *CH3O intermediate (Figure S11), there are three Raman peaks at 1473, 1425, 1398 cm–1, all of them coming from C–H bending, and another peak at 1027 cm–1 coming from C–O stretching.32 In low frequency region, the characteristic frequency 392 cm–1 comes from the C–H rocking of *CH2O. Figure 5c shows the simulated Raman spectra of CO reduction to CH3OH 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

through *COH intermediate along the Path II. Similar to the Path I from *CHO to CH3OH, the middle characteristic frequencies play a significant role in identifying the surface species. The 970 cm–1 vibration of *CHO comes from C– O stretching and O–C–H bending vibration. For *CHOH, the 1320 cm–1 peak comes from the C–O stretching vibration, and 1218 cm–1 does from O–H and C-H bending vibration. *CH2OH owns three main signals at 1248, 1086 and 914 cm–1, which can be assigned to C–O–H bending, CH2 bending and C–O stretching vibrations, respectively. For *CH3OH on silver clusters, the 1360 cm–1 peak comes from the mixture of C–O–H bending and O–C–H bending, and the 1070 cm–1 does from C–O stretching, C–O–H bending, and O–C–H bending (Figure S12), which are in agreement with experimental studies.88-89 Finally, it is worth to emphasize that the characteristic stretching frequencies of the C–Ag bond strongly depend on surface species. As shown in our calculated results, there are a fundamental 404 cm–1 for Ag–CH3, 422 and 322 cm–1 for Ag-CH2, 290, 346 and 424 cm–1 for Ag–CH, and 509, 449, and 254 cm–1 for Ag-C. The fundamentals of 424 and 232 cm–1 arise from Ag–CHO, 218 cm–1 from Ag–CH2O. Finally, the fundamentals of 283, 243 and 410 cm–1 can be assigned to Ag–COH, Ag–CHOH, Ag–CH2OH vibrations, respectively.90 The vibrational analysis shows that the characteristic frequencies of the reaction intermediates are located in the low and middle frequency regions. However, surface infrared spectroscopy is difficult to detect the low frequency vibrational peaks of surface species so far. Thus surface-enhanced Raman 20 ACS Paragon Plus Environment

Page 20 of 41

Page 21 of 41 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

The Journal of Physical Chemistry

spectroscopy may be a powerful technique to detect and identify surface species and the reaction intermediates of CO reduction on silver nanoparticles. Conclusions In summary, we have explored the mechanism of plasmon-driven CO2RR at silver clusters by DFT calculations. From the electron excitation of silver clusters, we find that the LUMO energy of CO2 well matches with sp electrons’ excitation energies of silver clusters, and the excited hot electrons on silver clusters can easily transfer to the LUMO orbital of CO2 because of the strong orbital coupling. The performed charge transfer state, having a chemisorption structure, greatly lowers the energy barriers of the C–O bond cleavage to form CO. In the succeeding reduction reactions of adsorbed CO with hydrogen, we have found that the favorable reduction path to the formation of CH3OH is through the *CHO intermediate, which is consistent with a recent experimental work of CO2RR on silver nanoparticles. Furthermore, we have calculated the Raman spectra for different intermediates to identify the surface species and deduce possible reaction paths. The present work may provide an encouraging motivation for experimental efforts to photo-electrochemical reduction of CO2. Supporting Information Computational details, calculated data of the geometric structures of chemisorbed and physisorbed states for CO2 on silver clusters, and the reaction barriers of C–O dissociation and CO2 with H interaction. Simulated absorption 21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

spectra and electronic coupling, and assignment of the electronic transitions. All calculated results of B3LYP functional calculations. Acknowledgements X.G.Z, C.Z, X.J, D.Y.W and Z.Q.T acknowledge support from the National Natural Science Foundation of China (NSFC) (21533006, 21621091, 21373172), the 973 Program of the Chinese Ministry of Science and Technology (2015CB932303 and Y2018YFC1602802) and Funds of State Key Laboratory of Physical Chemistry of Solid surface and Fujian Science and Technology Office. Y.X.L, Y.Z. thanks the support from NSFC (No. 21833006). D.Y.W thanks the Otto Mønsted fund. The authors thank Sheng-Hisen Lin and Martin Moskovits for helpful discussions. References 1. White, J. L.; Baruch, M. F.; Pander, J. E.; Hu, Y.; Fortmeyer, I. C.; Park, J. E.; Zhang, T.; Liao, K.; Gu, J.; Yan, Y., et al. Light-Driven Heterogeneous Reduction of Carbon Dioxide: Photocatalysts and Photoelectrodes. Chem. Rev. 2015, 115, 12888-12935. 2. Ager, J. W.; Lapkin, A. A. Chemical Storage of Renewable Energy. Science 2018, 360, 707. 3. Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrocatalytic Process of CO Selectivity in Electrochemical Reduction of CO2 at Metal Electrodes in Aqueous Media. Electrochim. Acta 1994, 39, 1833-1839. 4. De Luna, P.; Quintero-Bermudez, R.; Dinh, C.-T.; Ross, M. B.; Bushuyev, 22 ACS Paragon Plus Environment

Page 22 of 41

Page 23 of 41 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

The Journal of Physical Chemistry

O. S.; Todorović, P.; Regier, T.; Kelley, S. O.; Yang, P.; Sargent, E. H. Catalyst Electro-Redeposition Controls Morphology and Oxidation State for Selective Carbon Dioxide Reduction. Nat. Catal. 2018, 1, 103-110. 5. Dinh, C.-T.; Burdyny, T.; Kibria, M. G.; Seifitokaldani, A.; Gabardo, C. M.; García de Arquer, F. P.; Kiani, A.; Edwards, J. P.; De Luna, P.; Bushuyev, O. S., et al. CO2 Electroreduction to Ethylene Via Hydroxide-Mediated Copper Catalysis at an Abrupt Interface. Science 2018, 360, 783. 6. Salehi-Khojin, A.; Jhong, H.-R. M.; Rosen, B. A.; Zhu, W.; Ma, S.; Kenis, P. J. A.; Masel, R. I. Nanoparticle Silver Catalysts That Show Enhanced Activity for Carbon Dioxide Electrolysis. J. Phys. Chem. C 2013, 117, 1627-1632. 7. Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Compositional Dependence of the Stability of AuCu Alloy Nanoparticles. Chem. Commun. 2012, 48, 5626-5628. 8. Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Synergistic Geometric and Electronic Effects for Electrochemical Reduction of Carbon Dioxide Using Gold–Copper Bimetallic Nanoparticles. Nat. Commun. 2014, 5, 4948. 9. Chen, L. D.; Urushihara, M.; Chan, K.; Nørskov, J. K. Electric Field Effects in Electrochemical CO2 Reduction. ACS Catal. 2016, 6, 7133-7139. 10. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic Reduction of Carbon Dioxide in Aqueous Suspensions of Semiconductor Powders. Nature 1979, 277, 637. 11. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A Review of Catalysts for the 23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Electroreduction of Carbon Dioxide to Produce Low-Carbon Fuels. Chem. Soc. Rev. 2014, 43, 631-675. 12. He, Z.; Tang, J.; Shen, J.; Chen, J.; Song, S. Enhancement of Photocatalytic Reduction of CO2 to CH4 over TiO2 Nanosheets by Modifying with Sulfuric Acid. Appl. Surf. Sci. 2016, 364, 416-427. 13. Khan, S. U. M.; Al-Shahry, M.; Ingler, W. B. Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2. Science 2002, 297, 2243. 14. Habisreutinger Severin, N.; Schmidt-Mende, L.; Stolarczyk Jacek, K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372-7408. 15. Tan, S.; Zhao, Y.; Zhao, J.; Wang, Z.; Ma, C.; Zhao, A.; Wang, B.; Luo, Y.; Yang, J.; Hou, J. CO2 Dissociation Activated through Electron Attachment on the Reduced Rutile TiO2(110)-1×1 Surface. Phys. Rev. B 2011, 84, 155418. 16. Liu, E.; Kang, L.; Wu, F.; Sun, T.; Hu, X.; Yang, Y.; Liu, H.; Fan, J. Photocatalytic Reduction of CO2 into Methanol over Ag/TiO2 Nanocomposites Enhanced by Surface Plasmon Resonance. Plasmonics 2014, 9, 61-70. 17. Hartland, G. V. Optical Studies of Dynamics in Noble Metal Nanostructures. Chem. Rev. 2011, 111, 3858-3887. 18. Christopher, P.; Xin, H.; Linic, S. Visible-Light-Enhanced Catalytic Oxidation Reactions on Plasmonic Silver Nanostructures. Nat. Chem. 2011, 3, 467. 19. Kim, Y.; Dumett Torres, D.; Jain, P. K. Activation Energies of Plasmonic 24 ACS Paragon Plus Environment

Page 24 of 41

Page 25 of 41 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

The Journal of Physical Chemistry

Catalysts. Nano Lett. 2016, 16, 3399-3407. 20. Zhang, Y.; He, S.; Guo, W.; Hu, Y.; Huang, J.; Mulcahy, J. R.; Wei, W. D. Surface-Plasmon-Driven Hot Electron Photochemistry. Chem. Rev. 2018, 118, 2927-2954. 21. Long, R.; Prezhdo, O. V. Instantaneous Generation of Charge-Separated State on TiO2 Surface Sensitized with Plasmonic Nanoparticles. J. Am. Chem. Soc. 2014, 136, 4343-4354. 22. Mukherjee, S.; Libisch, F.; Large, N.; Neumann, O.; Brown, L. V.; Cheng, J.; Lassiter, J. B.; Carter, E. A.; Nordlander, P.; Halas, N. J. Hot Electrons Do the Impossible: Plasmon-Induced Dissociation of H2 on Au. Nano Lett. 2013, 13, 240-247. 23. Neaţu, Ş.; Maciá-Agulló, J. A.; Concepción, P.; Garcia, H. Gold–Copper Nanoalloys Supported on TiO2 as Photocatalysts for CO2 Reduction by Water. J. Am. Chem. Soc. 2014, 136, 15969-15976. 24. Choi, C. H.; Chung, K.; Nguyen, T.-T. H.; Kim, D. H. Plasmon-Mediated Electrocatalysis for Sustainable Energy: From Electrochemical Conversion of Different Feedstocks to Fuel Cell Reactions. ACS Energy Lett. 2018, 3, 14151433. 25. Yu, S.; Wilson, A. J.; Kumari, G.; Zhang, X.; Jain, P. K. Opportunities and Challenges of Solar-Energy-Driven Carbon Dioxide to Fuel Conversion with Plasmonic Catalysts. ACS Energy Lett. 2017, 2, 2058-2070. 26. Hou, W.; Hung, W. H.; Pavaskar, P.; Goeppert, A.; Aykol, M.; Cronin, S. B. 25 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Photocatalytic Conversion of CO2 to Hydrocarbon Fuels Via PlasmonEnhanced Absorption and Metallic Interband Transitions. ACS Catal. 2011, 1, 929-936. 27. Gao, D.; Zegkinoglou, I.; Divins, N. J.; Scholten, F.; Sinev, I.; Grosse, P.; Roldan Cuenya, B. Plasma-Activated Copper Nanocube Catalysts for Efficient Carbon Dioxide Electroreduction to Hydrocarbons and Alcohols. ACS Nano 2017, 11, 4825-4831. 28. Ichinohe, Y.; Wadayama, T.; Hatta, A. Electrochemical Reduction of CO2 on Silver as Probed by Surface-Enhanced Raman Scattering. J. Raman Spectrosc. 1995, 26, 335-340. 29. Choi, K. M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C. A.; Barmanbek, J. T. D.; Alshammari, A. S.; Yang, P.; Yaghi, O. M. Plasmon-Enhanced Photocatalytic CO2 Conversion within Metal–Organic Frameworks under Visible Light. J. Am. Chem. Soc. 2017, 139, 356-362. 30. Robatjazi, H.; Zhao, H.; Swearer, D. F.; Hogan, N. J.; Zhou, L.; Alabastri, A.; McClain, M. J.; Nordlander, P.; Halas, N. J. Plasmon-Induced Selective Carbon Dioxide Conversion on Earth-Abundant Aluminum-Cuprous Oxide Antenna-Reactor Nanoparticles. Nat. Commun. 2017, 8, 27. 31. Yu, S.; Wilson, A. J.; Heo, J.; Jain, P. K. Plasmonic Control of Multi-Electron Transfer and C–C Coupling in Visible-Light-Driven CO2 Reduction on Au Nanoparticles. Nano Lett. 2018, 18, 2189-2194. 32. Low, J.; Qiu, S.; Xu, D.; Jiang, C.; Cheng, B. Direct Evidence and 26 ACS Paragon Plus Environment

Page 26 of 41

Page 27 of 41 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

The Journal of Physical Chemistry

Enhancement of Surface Plasmon Resonance Effect on Ag-Loaded TiO2 Nanotube Arrays for Photocatalytic CO2 Reduction. Appl. Surf. Sci. 2018, 434, 423-432. 33. Song, H.; Meng, X.; Dao, T. D.; Zhou, W.; Liu, H.; Shi, L.; Zhang, H.; Nagao, T.; Kako, T.; Ye, J. Light-Enhanced Carbon Dioxide Activation and Conversion by Effective Plasmonic Coupling Effect of Pt and Au Nanoparticles. ACS Appl. Mater. Interfaces 2018, 10, 408-416. 34. Kostecki, R.; Augustynski, J. Photon-Driven Reduction Reactions on Silver. J. Appl. Electrochem. 1993, 23, 567-572. 35. Kostecki, R.; Augustynski, J. Unusually Strong Cathodic Photoeffect at Silver in Contact with Aqueous Solutions Containing Carbon Dioxide. Chem. Phys. Lett. 1992, 194, 386-390. 36. Zhao, S.; Jin, R.; Jin, R. Opportunities and Challenges in CO2 Reduction by Gold- and Silver-Based Electrocatalysts: From Bulk Metals to Nanoparticles and Atomically Precise Nanoclusters. ACS Energy Lett. 2018, 3, 452-462. 37. Kim, Y.; Creel, E. B.; Corson, E. R.; McCloskey, B. D.; Urban, J. J.; Kostecki, R. Surface-Plasmon-Assisted Photoelectrochemical Reduction of CO2 and NO3− on Nanostructured Silver Electrodes. Adv. Energy Mater. 2018, 8, 1800363. 38. Shoichiro, I.; Takehiko, T.; Kaname, I. Selective Formation of Formic Acid, Oxalic Acid, and Carbon Monoxide by Electrochemical Reduction of Carbon Dioxide. Bull. Chem. Soc. Jpn. 1987, 60, 2517-2522. 27 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

39. Hatsukade, T.; Kuhl, K. P.; Cave, E. R.; Abram, D. N.; Jaramillo, T. F. Insights into the Electrocatalytic Reduction of CO2 on Metallic Silver Surfaces. Phys. Chem. Chem. Phys. 2014, 16, 13814-13819. 40. Deng, W.; Zhang, L.; Dong, H.; Chang, X.; Wang, T.; Gong, J. Achieving Convenient CO2 Electroreduction and Photovoltage in Tandem Using PotentialInsensitive Disordered Ag Nanoparticles. Chem. Sci. 2018, 9, 6599-6604. 41. Peterson, A. A.; Nørskov, J. K. Activity Descriptors for CO2 Electroreduction to Methane on Transition-Metal Catalysts. J. Phys. Chem. Lett. 2012, 3, 251258. 42. Pereiro, M.; Baldomir, D. Structure of Small Silver Clusters and Static Response to an External Electric Field. Phys. Rev. A 2007, 75, 033202. 43. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A., et al. Gaussian 09, Rev. D.01; Gaussian, Inc.: Wallingford, CT, 2013. 44. Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158-6170. 45. Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007-1023. 46. Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent 28 ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41 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

The Journal of Physical Chemistry

Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378. 47. Foster, J. P.; Weinhold, F. Natural Hybrid Orbitals. J. Am. Chem. Soc. 1980, 102, 7211-7218. 48. Reed, A. E.; Weinstock, R. B.; Weinhold, F. Natural Population Analysis. J. Chem. Phys. 1985, 83, 735-746. 49. Maitra, N. T. Undoing Static Correlation: Long-Range Charge Transfer in Time-Dependent Density-Functional Theory. J. Chem. Phys. 2005, 122, 234104. 50. Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Challenges for Density Functional Theory. Chem. Rev. 2012, 112, 289-320. 51. Leng, C.; Qin, H.; Si, Y.; Zhao, Y. Theoretical Prediction of the Rate Constants for Exciton Dissociation and Charge Recombination to a Triplet State in PCPDTBT with Different Fullerene Derivatives. J. Phys. Chem. C 2014, 118, 1843-1855. 52. Valeev, E. F.; Coropceanu, V.; da Silva Filho, D. A.; Salman, S.; Brédas, J.L. Effect of Electronic Polarization on Charge-Transport Parameters in Molecular Organic Semiconductors. J. Am. Chem. Soc. 2006, 128, 9882-9886. 53. Zhang, W.; Liang, W.; Zhao, Y. Non-Condon Effect on Charge Transport in Dithiophene-Tetrathiafulvalene Crystal. J. Chem. Phys. 2010, 133, 024501. 54. Neugebauer, J.; Reiher, M.; Kind, C.; Hess, B. A. Quantum Chemical Calculation of Vibrational Spectra of Large Molecules—Raman and IR Spectra 29 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 30 of 41

for Buckminsterfullerene. J. Comput. Chem. 2002, 23, 895-910. 55. Yu, L.-J.; Pang, R.; Tao, S.; Yang, H.-T.; Wu, D.-Y.; Tian, Z.-Q. Solvent Effect

and

Hydrogen

Bond

Interaction

on

Tautomerism,

Vibrational

Frequencies, and Raman Spectra of Guanine: A Density Functional Theoretical Study. J. Phys. Chem. A 2013, 117, 4286-4296. 56. Schwarz, H. A.; Dodson, R. W. Reduction Potentials of CO2- and the Alcohol Radicals. J. Phys. Chem. 1989, 93, 409-414. 57. Sakurai, M.; Okano, T.; Tuzi, Y. Vibrational Excitation of Physisorbed CO2 on a Ag(111) Surface. J. Vac. Sci. Technol. A 1987, 5, 431-434. 58. Greenwood, N.; Earnshaw, A., Chemistry of the Elements 2nd Edition; Butterworth-Heinemann, 1997. 59. Zhang, X.-G.; Jin, X.; Wu, D.-Y.; Tian, Z.-Q. Selective Electrocatalytic Mechanism of CO2 Reduction Reaction to CO on Silver Electrodes: A Unique Reaction Intermediate. J. Phys. Chem. C 2018, 122, 25447-25455. 60. Kittel, C., Introduction to Solid State Physics. Fifth Edition; Wiley,Sussex, Eng.: United Kingdom, 1976. 61. Calle-Vallejo, F.; Loffreda, D.; Koper, M. T. M.; Sautet, P. Introducing Structural Sensitivity into Adsorption–Energy Scaling Relations by Means of Coordination Numbers. Nat. Chem. 2015, 7, 403. 62. Jacox, M. E.; Thompson, W. E. The Vibrational Spectra of Molecular Ions Isolated in Solid Neon. I. CO2+ and CO2−. J. Chem. Phys. 1989, 91, 1410-1416. 63. Janik, I.; Tripathi, G. N. R. The Nature of the CO2− Radical Anion in Water. 30 ACS Paragon Plus Environment

Page 31 of 41 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

The Journal of Physical Chemistry

J. Chem. Phys. 2016, 144, 154307. 64. Santos, V. O.; Leite, I. R.; Brolo, A. G.; Rubim, J. C. The Electrochemical Reduction of CO2 on a Copper Electrode in 1-N-Butyl-3-Methyl Imidazolium Tetrafluoroborate

(BMI.BF4)

Monitored

by

Surface-Enhanced

Raman

Scattering (SERS). J. Raman Spectrosc. 2016, 47, 674-680. 65. Maynard, K. J.; Moskovits, M. An Electron Energy Loss Study of Carbon Dioxide Adsorption on Alkali Metal Predosed Silver Surfaces. Surf. Sci. 1990, 225, 40-46. 66. Smith, E. L.; Porter, M. D. Structure of Monolayers of Short Chain NAlkanoic Acids (CH3(CH2)nCOOH, N = 0-9) Spontaneously Adsorbed from the Gas Phase at Silver as Probed by Infrared Reflection Spectroscopy. J. Phys. Chem. 1993, 97, 8032-8038. 67. Linic, S.; Christopher, P.; Xin, H.; Marimuthu, A. Catalytic and Photocatalytic Transformations on Metal Nanoparticles with Targeted Geometric and Plasmonic Properties. Acc. Chem. Res. 2013, 46, 1890-1899. 68. Jensen, L.; Zhao, L. L.; Schatz, G. C. Size-Dependence of the Enhanced Raman Scattering of Pyridine Adsorbed on Agn (N = 2−8, 20) Clusters. J. Phys. Chem. C 2007, 111, 4756-4764. 69. Jensen, L.; Aikens, C. M.; Schatz, G. C. Electronic Structure Methods for Studying Surface-Enhanced Raman Scattering. Chem. Soc. Rev. 2008, 37, 1061-1073. 70. Kalhor, M. P.; Chermette, H.; Chambrey, S.; Ballivet-Tkatchenko, D. From 31 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

CO2 to Dimethyl Carbonate with Dialkyldimethoxystannanes: The Key Role of Monomeric Species. Phys. Chem. Chem. Phys. 2011, 13, 2401-2408. 71. Luther III, G. W. Kinetics of the Reactions of Water, Hydroxide Ion and Sulfide Species with CO2, OCS and CS2: Frontier Molecular Orbital Considerations. Aquatic Geochemistry 2004, 10, 81-97. 72. Olsen, T.; Gavnholt, J.; Schiøtz, J. Hot-Electron-Mediated Desorption Rates Calculated from Excited-State Potential Energy Surfaces. Phys. Rev. B 2009, 79, 035403. 73. Li, J.; Cushing, S. K.; Meng, F.; Senty, T. R.; Bristow, A. D.; Wu, N. Plasmon-Induced Resonance Energy Transfer for Solar Energy Conversion. Nature Photonics 2015, 9, 601. 74. Brongersma, M. L.; Halas, N. J.; Nordlander, P. Plasmon-Induced Hot Carrier Science and Technology. Nat. Nanotechnol. 2015, 10, 25. 75. Kale, M. J.; Avanesian, T.; Xin, H.; Yan, J.; Christopher, P. Controlling Catalytic Selectivity on Metal Nanoparticles by Direct Photoexcitation of Adsorbate–Metal Bonds. Nano Lett. 2014, 14, 5405-5412. 76. Vollmer, S.; Witte, G.; Wöll, C. Determination of Site Specific Adsorption Energies of CO on Copper. Catal. Lett. 2001, 77, 97-101. 77. Muttaqien, F.; Hamamoto, Y.; Inagaki, K.; Morikawa, Y. Dissociative Adsorption of CO2 on Flat, Stepped, and Kinked Cu Surfaces. J. Chem. Phys. 2014, 141, 034702. 78. Dutta, A.; Morstein, C. E.; Rahaman, M.; Cedeño López, A.; Broekmann, 32 ACS Paragon Plus Environment

Page 32 of 41

Page 33 of 41 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

The Journal of Physical Chemistry

P. Beyond Copper in CO2 Electrolysis: Effective Hydrocarbon Production on Silver-Nanofoam Catalysts. ACS Catal. 2018, 8, 8357-8368. 79. Asadi, M.; Kim, K.; Liu, C.; Addepalli, A. V.; Abbasi, P.; Yasaei, P.; Phillips, P.; Behranginia, A.; Cerrato, J. M.; Haasch, R., et al. Nanostructured Transition Metal Dichalcogenide Electrocatalysts for CO2 Reduction in Ionic Liquid. Science 2016, 353, 467. 80. Tsai, C.; Abild-Pedersen, F.; Nørskov, J. K. Tuning the MoS2 Edge-Site Activity for Hydrogen Evolution Via Support Interactions. Nano Lett. 2014, 14, 1381-1387. 81. Cheng, M.-J.; Clark, E. L.; Pham, H. H.; Bell, A. T.; Head-Gordon, M. Quantum Mechanical Screening of Single-Atom Bimetallic Alloys for the Selective Reduction of CO2 to C1 Hydrocarbons. ACS Catal. 2016, 6, 77697777. 82. 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. 83. Nie, X.; Luo, W.; Janik, M. J.; Asthagiri, A. Reaction Mechanisms of CO2 Electrochemical Reduction on Cu(111) Determined with Density Functional Theory. J. Catal. 2014, 312, 108-122. 84. Cooney, R. P.; Mahoney, M. R.; Howard, M. W. Intense Raman Spectra of Surface Carbon and Hydrocarbons on Silver Electrodes. Chem. Phys. Lett. 1980, 76, 448-452. 33 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

85. Pockrand, I. A Raman Vibrational Study of Water Adsorption on Silver. Surf. Sci. 1982, 122, L569-L573. 86. Mahojey, M. R.; Howard, M. W.; Cooney, R. P. Carbon Dioxide Conversion to Hydrocarbons at Silver Electrode Surfaces: Raman Spectroscpic Evidence for Surface Carbon Intermediates. Chem. Phys. Lett. 1980, 71, 59-63. 87. Pettinger, B.; Philpott, M. R.; Gordon, J. G. Contribution of Specifically Adsorbed Ions, Water, and Impurities to the Surface Enhanced Raman Spectroscopy (SERS) of Ag Electrodes. J. Chem. Phys. 1981, 74, 934-940. 88. Wang, C.-B.; Deo, G.; Wachs, I. E. Interaction of Polycrystalline Silver with Oxygen, Water, Carbon Dioxide, Ethylene, and Methanol:  In Situ Raman and Catalytic Studies. J. Phys. Chem. B 1999, 103, 5645-5656. 89. Sobocinski, R. L.; Pemberton, J. E. Determination of Alcohol Solvent Orientation and Bonding at Silver Electrodes Using Surface-Enhanced Raman Scattering: Methanol, Ethanol, 1-Propanol, and 1-Pentanol. Langmuir 1992, 8, 2049-2063. 90. Chen, Y.-L.; Panneerselvam, R.; Wu, D.-Y.; Tian, Z.-Q. Theoretical Study of Normal Raman Spectra and SERS of Benzyl Chloride and Benzyl Radical on Silver Electrodes. J. Raman Spectrosc. 2017, 48, 53-63.

34 ACS Paragon Plus Environment

Page 34 of 41

Page 35 of 41 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

The Journal of Physical Chemistry

TOC Graphic

35 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Figure 1. Calculated Raman spectra of chemisorbed CO2 on different silver clusters. The Raman intensity is expanded on the basis of the Lorentzian line shape with the linewidth of 10 cm–1. 222x175mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 36 of 41

Page 37 of 41 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

The Journal of Physical Chemistry

Figure 2. Orbital coupling between the silver cluster and CO2. The red arrow, orange arrow, and violet arrow represents the highest oscillator strength transition of the Ag18 cluster, CO2 physisorption structure, and CO2 chemisorption structure, respectively. 231x191mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Figure 3. Mechanism of the photo-generation hot electron transfer from a silver cluster to CO2. The black line and the blue line represent the density of state of Ag18 cluster and CO2, respectively. The energy level is relative to the vacuum level. 252x191mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 38 of 41

Page 39 of 41 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

The Journal of Physical Chemistry

Figure 4. Three reaction paths for CO reduction on Ag18 cluster. The green arrows (Path I) and blue arrows (Path II) present two different reaction path with the same reduction product, CH3OH. And the red arrows are a reaction path from CO to CH4 (Path III). The H2O product is omitted in the scheme for simplicity. The numbers represent reaction activation energies at 298.15 K and 1 atm. 190x113mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Figure 5. Simulated Raman spectra of the intermediates for CO reduction on Ag18 cluster. a) CO reduction to CH4 through *CH, *CH2, and *CH3; b) CO reduction to CH3OH through *CHO, CH2O, and CH3O; c) CO reduction to CH3OH through *COH, *CHOH, and *CH2OH. The excitation wavelength of 532 nm was used here with a Lorentzian line width of 10 cm–1. 172x255mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 41

Page 41 of 41 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

The Journal of Physical Chemistry

TOC Graphic

ACS Paragon Plus Environment