Data-Driven Systematic Search of Promising Photocatalysts for Water

2 days ago - ... efficient optical absorption, and appropriate band edge positions, which are suitable for photocatalytic water splitting in the visib...
0 downloads 0 Views 2MB Size
Letter Cite This: J. Phys. Chem. Lett. 2019, 10, 5211−5218

pubs.acs.org/JPCL

Data-Driven Systematic Search of Promising Photocatalysts for Water Splitting under Visible Light Hao Jin,† Huijun Zhang,† Jianwei Li,† Tao Wang,† Langhui Wan,*,† Hong Guo,†,‡ and Yadong Wei*,† †

Downloaded via NOTTINGHAM TRENT UNIV on August 24, 2019 at 01:26:14 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Shenzhen Key Laboratory of Advanced Thin Films and Applications, College of Physics and Optoelectronic Engineering, Shenzhen University, 518060 Shenzhen, P. R. China ‡ Centre for the Physics of Materials and Department of Physics, McGill University, H3A 2T8 Montréal, Canada ABSTRACT: Searching for highly efficient, environmentally friendly, and noble metal free photocatalysts is an important topic in the photocatalytic field. The combination of data-driven high-throughput screening and density functional theory (DFT) might be one of the most effective ways. To this end, we have carried out a systematic search of the Materials Project Database. A high-throughput screening method is employed, and six criteria are selected as the indicators. The screening reduces the list of candidates from 83989 to 22 structures that show promise as photocatalysts. The electronic and optical properties are then determined by DFT calculations. Our results indicate that ZnSe, Ga2Se3, and Na2Zn2O3 show suitable direct band gaps, efficient optical absorption, and appropriate band edge positions, which are suitable for photocatalytic water splitting in the visible light region. We believe that this work not only proposes novel photocatalysts but also provides an efficient way of searching for advanced photocatalysts for water splitting.

T

semiconductors have been identified as promising photocatalysts for overall water splitting.24 The materials community usually employs a guess-and-check approach to explore new materials, which includes low efficiencies, high costs, and long periods of time.25,26 In addition, the previous studies have so far mainly focused on a single material or a single prototype, while the overall picture of photocatalytic materials is still lacking.27 The disadvantages of traditional material exploration impede the fast progress of booming materials science. Note that in the past century, people have accumulated knowledge (e.g., crystal structures and physical properties) and recorded it in a database such as the Materials Project (MP) Database,28,29 the Inorganic Crystal Structure Database (ICSD),30 the Crystallographic Open Database (COD),31 etc. Though there are thousands of semiconductors in these databases, most of the materials are untested, and some of them might be excellent photocatalysts. To take advantage of these databases and accelerate the process for materials exploration, a data-driven highthroughput screening method has been developed in recent years, which is expected to significantly improve the efficiency of material design.25,32,33 These techniques have been successfully employed in the search for two-dimensional (2D) materials, solid lithium-ion conductor materials, inorganic photovoltaic materials, ferromagnetic materials, light-absorbing materials, etc.25−27,34−41

he energy crisis and environmental pollution are two major issues in modern society. As such, developing clean and renewable energy resources has become imperative.1 Since the pioneering work of Fujishima, Honda, and co-workers,2 photocatalysis has been considered as an attractive and promising strategy for converting solar energy into environmentally friendly production of H2.3−6 The first proposed photocatalyst is TiO2, which, however, has a relatively large band gap (3.2 eV) and absorbs only a small proportion (∼5%) of solar energy.7,8 Therefore, to harvest sunlight, a considerable number of studies have been dedicated to developing novel photocatalysts,9−11 and many semiconductors such as SrTiO3, AgCl, and transition metal dichalcogenides (TMDs) have been proposed in recent years.12−17 However, the solar to hydrogen (STH) conversion efficiency is too low to meet the practical need, which is mainly ascribed to the poor catalytic activities, low light absorption efficiency, and inefficient photogenerated charge carrier separations.18−20 For any semiconductor that can be applied in photocatalytic water splitting, several criteria must be met.21 First, the proposed photocatalytic materials should have suitable direct band gaps in the visible light range to harvest solar energy efficiently. Second, the band edge positions of the photocatalysts can perfectly straddle the water redox potential, in which the conduction band minimum (CBM) is above the reduction potential of H+/H2 (−4.44 eV vs vacuum at pH 0), while the valence band maximum (VBM) is below the oxidation potential of H2O/O2 (−5.67 eV vs vacuum at pH 0). In addition, the photogenerated electrons and holes should be efficiently separated to guarantee a high photocatalytic efficiency.22,23 Because of these strict criteria, searching for highly efficient photocatalysts is a considerable challenge. In the past few decades, only a handful of © XXXX American Chemical Society

Received: July 8, 2019 Accepted: August 22, 2019 Published: August 22, 2019 5211

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters Here, we attempt to systematically explore nontoxic, lowcost, stable, and highly efficient photocatalytic materials. We perform this search on the basis of the MP database, and six criteria are applied to rule out the eligible photocatalyst. From the 83989 entries in the MP Database, 22 potential photocatalytic materials are identified. Then, we perform density functional theory (DFT) calculations to precisely study the electronic and optical properties of these selected candidates. The results show that nine materials are finally suitable for photocatalytic water splitting in the visible light region. We believe that this work provides an efficient way to explore novel photocatalysts, which could stimulate further theoretical and experimental investigations of material design for the production of hydrogen from water splitting. All calculations are performed on the basis of the DFT as implemented in the Vienna ab initio Simulation Package (VASP).42 The projector augmented wave (PAW) scheme with the Perdew−Burke−Ernzerhof (PBE) generalized gradient approximation (GGA) exchange correlation functional is employed.43−46 The cutoff energy for the basis set is chosen to be 500 eV, and the Brillouin zone is sampled using a Monkhorst−Pack grid with a separation of ∼0.03 Å−1.47 The convergence criteria are set to be 10−5 eV in energy and 0.01 eV Å−1 in force. For accurate band gap calculations, we employ the Heyd−Scuseria−Ernzerh (HSE06) of the hybrid density functional, in which the hybrid functional is mixed with 25% exact Hartree−Fock (HF) exchange.48 To find cheaper and nontoxic photocatalysts, we first do a formula screening from the MP database, in which up to 83989 materials are examined. In the formula screening step, we require that a candidate should not contain any radioactive element, toxic element, or noble element. This guarantees that the selected candidates contain environmentally friendly and noble metal free elements, which is crucial for their practical applications. Then, we compile data for a stability screening, in which a candidate must meet the following prerequisite criteria. (i) It should have ordered structure and have been recorded in the ICSD. This eliminates those disordered structures as well as a number of hypothetical materials that have not been observed by the experiments. (ii) The selected candidate is thermodynamically stable. Here, we employ energy above the convex hull Ehull to quantify the stability of the materials, which is tabulated in the MP Database. We regard those compounds with Ehull values of >20 meV/atom as thermodynamically unstable and exclude them from the screening. These precriteria reduce 83989 entries in the MP Database to 5073 candidates as shown in Figure 1. As has been discussed above, a potential photocatalyst should possess suitable band edge positions. To this end, we predict the redox potentials of 5073 candidates. At the zero charge point, conduction band edge E0CB and valence band edge E0VB can be empirically estimated by49 0 ECB = χ (S) − Ee −

1 Eg 2

(1)

0 E VB = χ (S) − Ee +

1 Eg 2

(2)

Figure 1. Schematic diagram to illustrate the search procedure for photocatalytic materials. Through formula and stability screens (Screen 1), 83989 entries in the MP Database are reduced to 5073 candidates. Then, the band gap and band edge position screens are employed (Screen 2), which further reduces the data set to 169 candidates. After the application of the screen of effective mass (Screen 3), 22 candidates finally remain.

can be given by the geometric mean of the electronegativities of the constituent atoms:50 χ (S ) =

N

χ1r χ2 s ...χn − 1 p χnq

(3)

where χn, n, and N are the electronegativity of the constituent atom, the number of species, and the total number of atoms in the compound, respectively. Although this method cannot give exact absolute values, it may serve as a reference to give a rough estimation of the redox ability for those potential photocatalysts. For overall photocatalytic water splitting, the band edge positions should satisfy the relationships E0CB ≤ 0 and E0VB ≥ 1.23 eV [vs the normal hydrogen electrode (NHE)]. At the limit condition, Eg = 3.0 eV. We find χ(S) should approximately be in the range from 4.0 to 6.0 eV. The distribution of χ(S) is plotted in Figure 2a. To efficiently harvest solar energy, the band gap of the potential photocatalyst should be within the visible light range, i.e., 1.6−3.0 eV, which accounts for ∼50% of the solar energy. Notably, the electronic structures given in the MP Database are based on the GGA calculations, which usually under-

where Eg and Ee are the band gap energy of the semiconductor and the energy of the free electron on the hydrogen scale (4.44 eV), respectively, and χ(S) is the electronegativity of the compound. On the basis of the bond length arguments, χ(S)

Figure 2. Distributions of (a) electronegativity and (b) band gaps for the screening materials. 5212

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters Table 1. Materials Information for 22 Candidates after Screening ID

formula

space group

Ehull (meV)

volume (Å3)

no. of sites

band gap (PBE) (eV)

band gap (HSE) (eV)

mp-7907 mp-4979 mp-571162 mp-27272 mp-1019576 mp-19215 mp-5795 mp-22914 mp-570081 mp-1340 mp-30979 mp-13852 mp-2242 mp-571440 mp-28371 mp-24428 mp-7798 mp-1018040 mp-8086 mp-5909 mp-2979 mp-1190

Al2ZnSe4 AlCuS2 Ca3(GaN2)2 Ca3PI3 Ca4SiN4 CaCrO4 CaMg2N2 CuCl CuI Ga2Se3 GaPS4 Ge3N4 GeS K2CuBr3 K2Zn3O4 KH2N MgGeN2 MgSe Ga2Se3 Zn2GeO4 ZnGeN2 ZnSe

I4 I42d C2/c I4132 P21/c I41/amd P3m1 F43m P3m1 Cc P21/c P63/m Pnma Pnma C2/c P21/m Pna21 P63mc P21/c R3 Pna21 F43m

0 0 0 0 0 0 0 0 1 0 0 2 0 0 0 0 0 0 4 0 0 0

174.345 149.485 252.189 963.164 545.065 170.896 67.158 39.467 111.213 262.145 648.8 177.605 175.866 748.088 276.34 110.622 193.43 107.893 187.205 584.253 187.944 47.338

7 8 18 28 36 12 5 2 4 10 24 14 8 24 18 8 16 4 14 42 16 2

2.15 1.71 1.91 2.15 1.43 2.19 2.00 0.56 1.65 1.04 2.48 1.86 1.24 1.90 1.41 2.06 2.57 2.58 1.07 1.89 1.61 1.17

3.15 3.07 3.03 3.02 2.38 3.50 3.05 2.37 2.92 1.93 3.69 3.37 1.85 3.56 2.96 3.39 3.86 3.54 2.60 3.70 2.95 2.26

Figure 3. Optimized crystal structures of nine potential photocatalytic materials.

In addition to the suitable band gap and appropriate band edge positions, the separation of photogenerated electron and hole is another important factor that determines the photocatalytic efficiency of the materials. To rule out the eligible candidates, we demand that the difference in the effective mass between the electron and hole must be at least 3 times larger. In this case, the light carrier is easy to move away while the heavy carrier stays, resulting in a high separation efficiency of the electron and hole. The effective masses of electrons (m*e ) and holes (m*h ) are calculated by fitting parabolic functions of the CBM and VBM, which can be expressed as

estimate the band gaps. To include as many potential photocatalytic materials as possible, here we decide that those exhibiting a band gap (Eg) of 0−3.0 eV are chosen for the next screening. Accordingly, close to 1200 candidates remain after this screening. The distribution of the band gaps is shown in Figure 2b. It should be pointed out that in general, the direct-band gap photocatalyst shows a solar energy conversion efficiency better than those of indirect-band gap semiconductors. As such, we further examine the retained candidates and exclude all indirect-band gap semiconductors from the search set. On the basis of the nature of the band gap together with the band edge positions, 169 candidates are kept after the screenings. 5213

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters

Figure 4. Band structures of nine potential photocatalytic materials with HSE06. The Fermi level is set to zero.

i d2Ek yz zz * = ±ℏ2jjjj me/h j dk 2 zz k {

−1

that in all studied cases, the CBM and VBM are located at the Γ point. The band gaps are in the range of 1.85−2.96 eV, which cover the visible light range. In addition to an appropriate band gap, a good photocatalytic material demands suitable band edge positions. Here, we use the band gaps obtained with HSE06 to predict the band edges. Note that the redox potentials may be affected by the pH value. To examine the ability of overall photocatalytic water splitting, we plot the energy locations of the VBM and CBM with respect to the redox potentials of water splitting at various pH values in the range of 0−7. From the results shown in Figure 5a, it is clear that Na2Zn2O3 and ZnSe have perfect band edge positions, especially for Na2Zn2O3. The oxidation potential of O2/H2O is higher than the VBM of Na2Zn2O3, suggesting photogenerated holes can transfer from the valence band to H2O and oxidize it to O2 through the reaction

(4)

where k is the wave vector and Ek is the energy corresponding to wave vector k. These two parameters can be obtained from the band structures recorded in the MP Database.28,29 After this criterion is applied, 22 candidates finally remain, which are listed in Table 1. To precisely study the electronic and optical properties of the 22 selected materials, we perform DFT calculations, in which hybrid density functional HSE06 is employed to accurately give the band gap value. To efficiently harvest solar energy, the band gap of the potential photocatalyst should be within the visible light range, i.e., 1.6−3.0 eV. Although the rough calculations using PBE suggest 22 candidates, the more accurate calculations given by HSE06 suggest only nine candidates finally meet the 1.6 eV < Eg ≤ 3.0 eV criterion as listed in Table 1. In Figure 3, we plot the optimized crystal structures of nine potential photocatalytic materials. Apart from previously reported water splitting photocatalysts, i.e., GeS and ZnSe, we have proposed novel photocatalytic materials, such as K2Zn3O4, Na2Zn2O3, and Ga2Se3, which are mainly group VI compounds. The calculated band structures are shown in Figure 4. It is interesting to note

2H 2O + 4h+ → O2 + 4H+

(5)

The reduction potential of H+/H2 is lower than the CBM of Na2Zn2O3, leading to the electron transfer from CBM to H2O to further produce H2: 2H+ + 2e− → H 2 5214

(6) DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters

energy, we believe that these materials act as promising visible light-driven photocatalysts for water splitting. From the point view of potential photocatalytic applications, charge separation is another critical factor. Here, we calculate the carrier effective masses, which can present a quantitative description of the carrier transfer ability along the specific directions. The effective masses of electrons (me*) and holes (mh*) are calculated by eq 4. As shown in Table 2, a large difference in effective mass between the electron and hole is observed for these selected materials. In all studied cases, electrons show relatively lighter effective masses as compared with those of holes. In particular, mh* is >25 times heavier than m*e for CuCl. As a result, the photogenerated electrons are much easier to move away while the holes are left behind. The large discrepancy in effective mass between electrons and holes can promote the separation of photogenerated carriers, giving rise to better performance of photocatalytic water splitting. To estimate the photocatalytic performance, we predict the STH efficiency using the following equation:52,53 ηSTH = ηabsηcu (8) where ηabs is the efficiency of light absorption, which is defined as ∞

Figure 5. (a) Calculated edge positions of the VBM and CBM. The vacuum level is set as 0 eV. The redox potentials of water splitting are plotted with dashed lines at pH 0 (green) and pH 7 (orange). (b) Imaginary part of the dielectric function for nine photocatalysts.

ηabs =

ηcu =

α

β

∫0 P(hw) d(hw)

ΔG∫

∞ P(hw) hw

E ∞

(9)

d(hw)

∫E P(hw) d(hw) g

(10)

where ΔG is the potential energy difference for water splitting, i.e., 1.23 eV, and E is the energy of photons available in the process of photocatalysis. According to the definition of STH, ηSTH is calculated for selected materials (values listed in Table 2). It is noteworthy that for ZnSe, Ga2Se3, and Na2Zn2O3, ηSTH reaches values of 11.25%, 7.02%, and 5.33%, respectively. These values are much higher than that of TiO2 (i.e., ∼1.1%),54 suggesting the proposed materials are promising candidates for photocatalytic water splitting. In summary, this work reports a systematic search for photocatalytic materials based on the MP Database. More than 80000 materials are examined to rule out the eligible candidates for photocatalytic water splitting. Six criteria are applied. (1) A candidate should not contain any radioactive element, toxic element, or noble element. (2) It should have ordered structure and have been recorded in the ICSD. (3) The selected candidate is thermodynamically stable, with an Ehull of ≤20 meV/atom. (4) It has suitable band edge positions. (5) It exhibits an appropriate direct band gap. (6) A large difference in effective mass exists between the electron and hole. Consequently, 22 candidates remain. Then, we employ DFT calculations to precisely study the electronic and optical properties of the selected materials. Our results indicate that ZnSe, Ga2Se3, and Na2Zn2O3 show suitable direct band gaps, efficient optical absorption under visible light irradiation, and

4π 2e 2 1 lim ∑ 2ωkδ(εck − εvk − ω) Ω q → 0 q2 c , v , k × ⟨μck + qe |μvk ⟩ × ⟨μvk |μck + qe ⟩

g



where P(hw) is the AM1.5G solar energy flux at photon energy hw and Eg is the band gap of the semiconductor. The integral in the denominator represents the total power density of incident simulated sunlight, while the integral in the numerator represents the absorbed power density of semiconductors. ηcu is the efficiency of carrier utilization, which is given by

Except for CuCl, other materials also satisfy the requirement for photocatalytic water splitting at a certain pH value. For example, the band edges of Ga2Se3 can perfectly straddle the water redox potentials at pH >1.5, while for Ca4SiN4, it has a great reducing strength and can fully split the water in a neutral environment. Though the reduction potential of CuCl cannot satisfy the requirement of water splitting, we note that its oxidizing ability is extremely strong. The calculated oxidation potential is ≤7.3 eV (vs vacuum at pH 0), which is ∼1.0 eV higher than that of H2O2 and O3, and even comparable with that of TiO2. The strong oxidative ability combined with visible light absorption makes it an excellent candidate for further applications in the degradation of organic contaminants. To investigate the visible light absorption ability of these potential photocatalytic materials, we calculate the optical absorption spectrum. The imaginary part of the dielectric constant is determined by a summation over empty states using the following equation:51 ε2αβ (ω) =

∫E P(hw) d(hw)

(7)

where the indices c and v refer to the conduction and valence band states, respectively, and μck represents the cell periodic part of the wave functions at the k point. As shown in Figure 5b, GeS, Ga2Se3, and ZnSe compounds exhibit outstanding optical absorptions in the visible light region. In addition, the absorption peaks of Ga2Se3, CuCl, ZnSe, and Na2Zn2O3 are located at 500, 477, 470, and 431 nm, respectively, which are within the green−blue visible light spectrum. Considering the fact that visible light accounts for approximately 50% of solar 5215

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

2.42%

*E-mail: [email protected]. *E-mail: [email protected]. ORCID 7.02%

X−Γ Γ−N X−Γ Γ−N

The authors declare no competing financial interest.

■ 5.33%

0.26 0.32 0.87 2.36

ACKNOWLEDGMENTS This work is supported by the National Natural Science Foundation of China (11604213 and 11574217), the Shenzhen Key Lab Fund (Grant ZDSYS20170228105421966), and the Natural Science and Engineering Research Council of Canada (NSERC).

E−Γ Γ−Y E−Γ Γ−Y

Ga2Se3

Notes



1.90 0.72 1.69 2.17

Na2Zn2O3

Hao Jin: 0000-0002-5085-6144

1.68%

GeS Y−Γ Γ−U Y−Γ Γ−U 0.33 0.33 2.58 1.02

K2Zn3O4

Y−Γ Γ−M Y−Γ Γ−M 2.52% 0.27 0.27 6.52 3.20 −

CuCl X−Γ Γ−L X−Γ Γ−L 0.29 0.27 4.15 4.15

M−Γ Γ−L M−Γ Γ−L 2.78% 2.37% ηSTH

mh *

CuI

REFERENCES

(1) Chen, X.; Li, C.; Grätzel, M.; Kostecki, R.; Mao, S. S. Nanomaterials for Renewable Energy Production and Storage. Chem. Soc. Rev. 2012, 41 (23), 7909−7937. (2) Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238 (5358), 37−38. (3) Ma, X.-C.; Dai, Y.; Yu, L.; Huang, B.-B. Energy Transfer in Plasmonic Photocatalytic Composites. Light: Sci. Appl. 2016, 5, No. e16017. (4) Qu, Y.; Duan, X. Progress, Challenge and Perspective of Heterogeneous Photocatalysts. Chem. Soc. Rev. 2013, 42 (7), 2568− 2580. (5) Chen, X.; Shen, S.; Guo, L.; Mao, S. S. Semiconductor-based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110 (11), 6503−6570. (6) Hisatomi, T.; Kubota, J.; Domen, K. Recent Advances in Semiconductors for Photocatalytic and Photoelectrochemical Water Splitting. Chem. Soc. Rev. 2014, 43 (22), 7520−7535. (7) Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 Photocatalysis: A Historical Overview and Future Prospects. Jpn. J. Appl. Phys. 2005, 44 (12), 8269−8285. (8) Xu, H.; Ouyang, S.; Liu, L.; Reunchan, P.; Umezawa, N.; Ye, J. Recent advances in TiO2-based photocatalysis. J. Mater. Chem. A 2014, 2 (32), 12642−12661. (9) Jin, H.; Dai, Y.; Huang, B.-B. Design of Advanced Photocatalysis System by Adatom Decoration in 2D Nanosheets of Group-IV and III−V Binary Compounds. Sci. Rep. 2016, 6, 23104. (10) Liang, Y.; Long, C.; Li, J.; Jin, H.; Huang, B.; Dai, Y. InSe Monolayer: Promising Cocatalyst of g-C3N4 for Water Splitting under Visible Light. ACS Appl. Energy Mater. 2018, 1 (10), 5394−5401. (11) Long, C.; Liang, Y.; Jin, H.; Huang, B.; Dai, Y. PdSe2: Flexible Two-Dimensional Transition Metal Dichalcogenides Monolayer for Water Splitting Photocatalyst with Extremely Low Recombination Rate. ACS Appl. Energy Mater. 2019, 2 (1), 513−520. (12) Wang, P.; Huang, B.; Zhang, X.; Qin, X.; Jin, H.; Dai, Y.; Wang, Z.; Wei, J.; Zhan, J.; Wang, S.; et al. Highly Efficient Visible-Light Plasmonic Photocatalyst Ag@AgBr. Chem. - Eur. J. 2009, 15 (8), 1821−1824. (13) Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Wei, J.; Whangbo, M.-H. Ag@AgCl: A Highly Efficient and Stable Photo-

0.59 0.45 4.24 4.24

Ca4SiN4

AUTHOR INFORMATION

Corresponding Authors

S−Γ Γ−U S−Γ Γ−U

ZnSe

X−Γ 0.12 Γ−L 0.11 X−Γ 2.79 Γ−L 2.79 11.25% 0.19 0.20 0.43 2.22



0.20 0.19 2.83 2.93

ZnGeN2

appropriate band edge positions that meet the requirements of the redox potentials in photocatalytic water splitting. The large difference in carrier effective masses promotes the separation of photogenerated charge carriers. As such, we believe these materials are promising candidates for high-performance photocatalytic water splitting. Our work not only proposes novel photocatalysts but also provides a useful approach to searching for advanced photocatalysts for water splitting.

E−Γ Γ−Y E−Γ Γ−Y me*

Table 2. Effective Masses for Electrons (m*e ) and Holes (m*h ) along Different Paths in Reciprocal Space and Predicted Solar-to-Hydrogen Efficiencies (ηSTH) for Selected Candidates

The Journal of Physical Chemistry Letters

5216

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters catalyst Active under Visible Light. Angew. Chem., Int. Ed. 2008, 47 (41), 7931−7933. (14) Wei, W.; Dai, Y.; Guo, M.; Zhu, Y.; Huang, B. Density Functional Theory Study of Ag Adsorption on SrTiO3 (001) Surface. J. Phys. Chem. C 2010, 114 (24), 10917−10921. (15) Yi, J.; Li, H.; Gong, Y.; She, X.; Song, Y.; Xu, Y.; Deng, J.; Yuan, S.; Xu, H.; Li, H. Phase and Interlayer Effect of Transition Metal Dichalcogenide Cocatalyst toward Photocatalytic Hydrogen Evolution: The Case of MoSe2. Appl. Catal., B 2019, 243, 330−336. (16) Lu, Q.; Yu, Y.; Ma, Q.; Chen, B.; Zhang, H. 2D TransitionMetal-Dichalcogenide-Nanosheet-Based Composites for Photocatalytic and Electrocatalytic Hydrogen Evolution Reactions. Adv. Mater. 2016, 28 (10), 1917−1933. (17) Zhang, S.; Jin, H.; Long, C.; Wang, T.; Peng, R.; Huang, B.; Dai, Y. MoSSe Nanotube: a Promising Photocatalyst with an Extremely Long Carrier Lifetime. J. Mater. Chem. A 2019, 7 (13), 7885−7890. (18) Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-Photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24 (2), 229−251. (19) Kudo, A.; Miseki, Y. Heterogeneous Photocatalyst Materials for Water Splitting. Chem. Soc. Rev. 2009, 38 (1), 253−278. (20) Kubacka, A.; Fernández-García, M.; Colón, G. Advanced Nanoarchitectures for Solar Photocatalytic Applications. Chem. Rev. 2012, 112 (3), 1555−1614. (21) Fu, C.-F.; Wu, X.; Yang, J. Material Design for Photocatalytic Water Splitting from a Theoretical Perspective. Adv. Mater. 2018, 30 (48), 1802106. (22) Li, Y.; Li, Y.-L.; Sa, B.; Ahuja, R. Review of Two-Dimensional Materials for Photocatalytic Water Splitting from a Theoretical Perspective. Catal. Sci. Technol. 2017, 7 (3), 545−559. (23) Zhao, P.; Ma, Y.; Lv, X.; Li, M.; Huang, B.; Dai, Y. Twodimensional III2-VI3 Materials: Promising Photocatalysts for Overall Water Splitting under Infrared Light Spectrum. Nano Energy 2018, 51, 533−538. (24) Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J. M.; Domen, K.; Antonietti, M. A Metal-Free Polymeric Photocatalyst for Hydrogen Production from Water under Visible Light. Nat. Mater. 2009, 8, 76. (25) Zhang, X.; Chen, A.; Zhou, Z. High-Throughput Computational Screening of Layered and Two-Dimensional Materials. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2019, 9 (1), No. e1385. (26) Sendek, A. D.; Yang, Q.; Cubuk, E. D.; Duerloo, K.-A. N.; Cui, Y.; Reed, E. J. Holistic Computational Structure Screening of More Than 12 000 Candidates for Solid Lithium-Ion Conductor Materials. Energy Environ. Sci. 2017, 10 (1), 306−320. (27) Zhu, Y.; Kong, X.; Rhone, T. D.; Guo, H. Systematic Search for Two-Dimensional Ferromagnetic Materials. Phys. Rev. Mater. 2018, 2 (8), 081001. (28) Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1 (1), 011002. (29) Hattrick-Simpers, J. R.; Gregoire, J. M.; Kusne, A. G. Perspective: Composition-Structure-Property Mapping in HighThroughput Experiments: Turning Data into Knowledge. APL Mater. 2016, 4 (5), 053211. (30) Belsky, A.; Hellenbrandt, M.; Karen, V. L.; Luksch, P. New developments in the Inorganic Crystal Structure Database (ICSD): accessibility in support of materials research and design. Acta Crystallogr., Sect. B: Struct. Sci. 2002, 58 (3−1), 364−369. (31) Gražulis, S.; Daškevič, A.; Merkys, A.; Chateigner, D.; Lutterotti, L.; Quirós, M.; Serebryanaya, N. R.; Moeck, P.; Downs, R. T.; Le Bail, A. Crystallography Open Database (COD): an openaccess collection of crystal structures and platform for world-wide collaboration. Nucleic Acids Res. 2012, 40 (D1), D420−D427. (32) Rickman, J. M.; Lookman, T.; Kalinin, S. V. Materials Informatics: From the Atomic-Level to the Continuum. Acta Mater. 2019, 168, 473−510.

(33) Afzal, M. A. F.; Hachmann, J. High-Throughput Computational Studies in Catalysis and Materials Research, and their Impact on Rational Design. arXiv 2019, 1902.03721. (34) Howard, J. M.; Tennyson, E. M.; Neves, B. R. A.; Leite, M. S. Machine Learning for Perovskites’ Reap-Rest-Recovery Cycle. Joule 2019, 3 (2), 325−337. (35) Liu, H.; Sun, J.-T.; Liu, M.; Meng, S. Screening Magnetic TwoDimensional Atomic Crystals with Nontrivial Electronic Topology. J. Phys. Chem. Lett. 2018, 9 (23), 6709−6715. (36) Zhang, X.; Zhang, Z.; Yao, S.; Chen, A.; Zhao, X.; Zhou, Z. An Effective Method to Screen Sodium-based Layered Materials for Sodium Ion Batteries. npj Comput. Mater. 2018, 4 (1), 13. (37) Mounet, N.; Gibertini, M.; Schwaller, P.; Campi, D.; Merkys, A.; Marrazzo, A.; Sohier, T.; Castelli, I. E.; Cepellotti, A.; Pizzi, G.; et al. Two-Dimensional Materials from High-Throughput Computational Exfoliation of Experimentally Known Compounds. Nat. Nanotechnol. 2018, 13 (3), 246−252. (38) Yu, L.; Zunger, A. Identification of Potential Photovoltaic Absorbers Based on First-Principles Spectroscopic Screening of Materials. Phys. Rev. Lett. 2012, 108 (6), 068701. (39) Cheon, G.; Duerloo, K.-A. N.; Sendek, A. D.; Porter, C.; Chen, Y.; Reed, E. J. Data Mining for New Two- and One-Dimensional Weakly Bonded Solids and Lattice-Commensurate Heterostructures. Nano Lett. 2017, 17 (3), 1915−1923. (40) Schmidt, J.; Shi, J.; Borlido, P.; Chen, L.; Botti, S.; Marques, M. A. L. Predicting the Thermodynamic Stability of Solids Combining Density Functional Theory and Machine Learning. Chem. Mater. 2017, 29 (12), 5090−5103. (41) Fabini, D. H.; Koerner, M.; Seshadri, R. Candidate Inorganic Photovoltaic Materials from Electronic Structure-Based Optical Absorption and Charge Transport Proxies. Chem. Mater. 2019, 31 (5), 1561−1574. (42) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for ab initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54 (16), 11169−11186. (43) Kresse, G.; Furthmüller, J. Efficiency of ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6 (1), 15−50. (44) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77 (18), 3865− 3868. (45) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59 (3), 1758−1775. (46) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50 (24), 17953−17979. (47) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13 (12), 5188−5192. (48) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118 (18), 8207−8215. (49) Morrison, S. R. Electrochemistry at Semiconductor and Oxidized Metal Electrodes; Plenum Press: New York, 1980. (50) Sanderson, R. T. Chemical Periodicity; Reinhold: New York, 1960. (51) Gajdoš, M.; Hummer, K.; Kresse, G.; Furthmü ller, J.; Bechstedt, F. Linear Optical Properties in the Projector-Augmented Wave methodology. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73 (4), 045112. (52) Chen, Z.; Jaramillo, T. F.; Deutsch, T. G.; KleimanShwarsctein, A.; Forman, A. J.; Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; et al. Accelerating Materials Development for Photoelectrochemical Hydrogen Production: Standards for Methods, Definitions, and Reporting Protocols. J. Mater. Res. 2010, 25 (1), 3−16. (53) Fu, C.-F.; Sun, J.; Luo, Q.; Li, X.; Hu, W.; Yang, J. Intrinsic Electric Fields in Two-dimensional Materials Boost the Solar-toHydrogen Efficiency for Photocatalytic Water Splitting. Nano Lett. 2018, 18 (10), 6312−6317. 5217

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218

Letter

The Journal of Physical Chemistry Letters (54) Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R. C.; Wang, C.; Zhang, J. Z.; Li, Y. Hydrogen-Treated TiO2 Nanowire Arrays for Photoelectrochemical Water Splitting. Nano Lett. 2011, 11 (7), 3026−3033.

5218

DOI: 10.1021/acs.jpclett.9b01977 J. Phys. Chem. Lett. 2019, 10, 5211−5218