Arsenene-Based Heterostructures: Highly Efficient ... - ACS Publications

Nov 21, 2017 - ABSTRACT: Constructing suitable type II heterostructures is a reliable solution ... The strong and persistent demand for clean and rene...
6 downloads 18 Views 1MB Size
Subscriber access provided by LAURENTIAN UNIV

Article

Arsenene-based heterostructures: high-efficient bifunctional materials for photovoltaics and photocatalytics Xianghong Niu, Yunhai Li, QiongHua Zhou, Huabing Shu, and Jinlan Wang ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b14842 • Publication Date (Web): 21 Nov 2017 Downloaded from http://pubs.acs.org on November 25, 2017

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 free 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 accessible to all readers and 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.

ACS Applied Materials & Interfaces 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 17

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

ACS Applied Materials & Interfaces

Arsenene-based heterostructures: high-efficient bifunctional materials for photovoltaics and photocatalytics Xianghong Niu1, Yunhai Li1, Qionghua Zhou*1, Huabing Shu1,2 and Jinlan Wang*1,3 1

School of Physics, Southeast University, Nanjing 211189, People’s Republic of

China 2

School of Science, Jiangsu University of Science and Technology, Zhenjiang 212003,

People’s Republic of China 3

Synergetic Innovation Center for Quantum Effects and Applications (SICQEA),

Hunan Normal University, Changsha 410081, People’s Republic of China

ABSTRACT: Constructing suitable type II heterostructures is a reliable solution for high-efficient photovoltaic and photocatalytic materials. Arsenene, as a rising member of mono-elemental two-dimensional materials, shows great potential as a building block of heterostructures due to its suitable band gap, high carrier mobility and good optical properties. Based on accurate band structure calculations by combining many-body perturbation GW method with an extrapolation technique, we demonstrate that

arsenene-based

heterostructures

paired

with

molybdenum

disulfide,

tetracyano-quinodimethane or tetracyanonaphtho-quinodimethane can form type-II band alignments. These arsenene-based heterostructures can not only satisfy all the requirements as photocatalysts for photocatalytic water splitting, but also show excellent power conversion efficiency of ~ 20% as potential photovoltaics. KEYWORDS: arsenene-based heterostructures, photovoltaic cell, photocatalytic water splitting, density functional theory, GW method Email: [email protected]; [email protected]

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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 2 of 17

Introduction The strong and persistent demand for clean and renewable energy urges researchers to search for high-efficient photovoltaic and photocatalytic materials.

1-3

An ideal photovoltaic material should satisfy: (I) efficient photon harvesting in visible/UV region; (II) high carrier mobility; (III) good stability under ambient atmosphere; (IV) low recombination rate of photo-generated electron-hole pairs. To be a good photocatalyst, fine-tuned band edges straddling the water redox potential is additionally required. These rigorous requirements make simplex semiconductor photovoltaic and photocatalytic materials rather scare. A possible solution is to construct heterostructures with type-II band alignments, which can effectively facilitate the photogenerated electron-hole pairs migrating in different building blocks,4, 5 and reduce the recombination rate6. Heterostructures may also broaden the range of photoabsorption simultaneously, which is beneficial to the full utilization of solar energy. Therefore, developing type II heterostructures is highly demanding. Arsenene, as a newly synthesized member of mono-elemental two-dimensional (2D) materials, has attracted a surge of research interest very recently.7, moderate band gap (1.66 eV),

8

8

It owns a

high carrier mobility (102-104 cm2V-1s-1)8 and good

visible/UV light absorption capacity,9 which meets the requirements I&II for high-efficient photovoltaic and photocatalytic materials. Moreover, isolate arsenene was reported to be stable at a high temperature of 1000 K in vacuum.8, 10 These excellent electronic and optical properties endow arsenene potential building block of heterostructures to achieve high-efficient photovoltaics and photocatalysis.

ACS Paragon Plus Environment

Page 3 of 17

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

ACS Applied Materials & Interfaces

In this work, we systematically investigate the possibility of arsenene as a potential photovoltaic or photocatalytic material by using density functional theory (DFT) and many-body perturbation GW method in combination with an extrapolation approach. We first evaluate the environmental stability of arsenene. Then, for the counterpart of heterostructures, we consider stable and widely used 2D semiconductors and organic molecule materials, such as molybdenum dichalcogenides (MoS2),11-13 titanium trisulfide

(TiS3),14-18

tetracyano-quinodimethane

(TCNQ),13,

19-21

tetracyanonaphtho-quinodimethane (TCNNQ),22-24 tetracyanoethylene (TCNE)19, and benzyl viologen (BV).

19, 25

20

The accurate band structure calculations suggest that

As/MoS2, As/TCNQ and As/TCNNQ form type II heterostructures with suitable band gap, perfect band edge alignment and high efficient electron-hole separation for photocatalytic water splitting. Moreover, these heterostructures can reach the power conversion efficiency up to ~20% as photovoltaic solar cells, which make them potential bifunctional materials for both photovoltaics and photocatalysis.

Computational details The DFT calculations were carried out with the Quantum-ESPRESSO26 package. The general gradient approximation (GGA) of Perdew, Burke, Ernzerhof (PBE) exchange-correlation functional27 was employed to obtain the wave functions and energies of the ground states of materials. To avoid the unphysical interaction between periodic images, the supercell technique was employed with vacuum layers larger than 15 Å perpendicular to the surface for 2D crystal, and along all three dimensions for

organic

molecules.

Electron

nucleus

interaction

ACS Paragon Plus Environment

was

treated

using

ACS Applied Materials & Interfaces

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

Page 4 of 17

norm-conserving pseudopotentials.28 The kinetic energy cutoff was set as 150 Ry for TiS3, 80 Ry for other 2D crystal and organic molecules, respectively, which assure the convergence of total energy within 0.01 eV/atom. The Monkhorst-Pack k-grid of 14×10×1 for rectangle crystal, 12×12×1 for hexagonal crystal and only the gamma point for organic molecules was adopted for the Brillouin zone integration. Structure relaxation was performed until the total energy difference smaller than 10-4 eV and the residual forces on each atom less than 0.01 eV/Å. The single-shot G0W0 approximation29 was employed to obtain accurate quasi-particle (QP) valence and conduction band edges. The Coulomb interaction was truncated at the edge of the Wigner-Seitz cell to boost the convergence on the thickness of the vacuum layer.30 The convergence of QP band gap within 0.05 eV were carefully examined with respect to the Monkhosrt-Pack grid and the size of dielectric matrix. However, the accurate determination of QP energies involves summation over a large number of unoccupied bands.31-33 The more the number of unoccupied states, the higher accuracy the QP energies. Nevertheless, as the number of unoccupied states increases, the convergence of correlation part of self-energy will become very, very slow. Hence, we employed the extrapolation technique,31-33 which the energy of conduction band maximum (CBM, the lowest unoccupied molecular orbital in molecules) or valence band minimum (VBM, the highest occupied molecular orbital in molecules) E QP ( N ) is calculated according the following formula, E QP ( N ) = E QP (∞ ) +

a . (b + N )

ACS Paragon Plus Environment

(1)

Page 5 of 17

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

ACS Applied Materials & Interfaces

N is the number of unoccupied bands, a and b are fitting parameters, and E QP (∞ ) is the extrapolated energy of CBM or VBM. In actual calculations, we first computed a series of band energies versus the number of unoccupied bands, then fitted the data using equation (1) and determined the convergence energy of band edges from the fitting parameters. The G0W0 calculations were performed with the YAMBO code.34 Ab initio molecular dynamics (AIMD) simulations were carried out in the canonical ensemble with a time step of 1.0 fs. The temperature was set to 300 K using the Nosé thermostat method.35 The spin-polarization and dipole correction were employed to cancel the errors of total energy, electrostatic potential and atomic force, caused by periodic boundary condition.36 vdW interaction was included by the vdW-D2 level.37 The (4×5) supercell was used to minimize the constraints induced by the periodic model. Under ambient atmosphere, the AIMD simulation of interaction between O2 or H2O and arsenene was based on the norm-conserving pseudopotential 28

in the SIESTA program package.38 The climbing-image nudged elastic band method

(CI-NEB) incorporated with spin-polarization was employed to locate the minimum-energy path of the oxidation of arsenene.

Results and discussion

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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. AIMD simulations of interaction between arsenene and (a) H2O or (b) O2. dmin is the minimum distance between arsenene and molecules. Insets show snapshots of simulation. (c) Energy barrier of O2 dissociation on arsenene along two possible paths. (d) Optical absorption spectra of arsenene for incident light polarized along the armchair direction at the level of G0W0. The inset is the dominating electronic transition channel of the first absorption peak. The spectrum is broadened using Lorentzian-type broadening of 0.1 eV. O, H and As atoms are labeled as red, white and orange, respectively.

Arsenene has a puckered honeycomb structure with two sub-layers of As atoms, similar to black phosphorus (BP). Each As is covalently bonded to three adjacent As atoms to form rectangular lattice. The lattice constants along the zigzag and armchair direction are 4.79 Å and 3.69 Å, respectively, in good agreement with earlier results.10, 39

Isolated arsenene was reported to possess excellent dynamic and thermal

stability.39,10 However, this does not mean it is still stable under certain environment.

ACS Paragon Plus Environment

Page 6 of 17

Page 7 of 17

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

ACS Applied Materials & Interfaces

BP is a typical example. The isolated BP is stable, but the presence of oxygen and humidity causes serious degradation.40 Since arsenene has similar geometric structure to BP and belongs to the group V, it may also suffer the degradation problem under ambient condition. To clarify this point, we investigate the interaction between arsenene

and

H2O/O2.

The

binding

energies,

defined

as

Eb = E As/( H 2O or O2 ) − E As − E( H 2O or O2 ) , are about -189 and -146 meV for H2O and O2,

respectively, indicating that the interaction between arsenene and H2O/O2 is physisorption. AIMD simulations further show that the H2O or O2 molecules drift away from the arsenene surface within 5 ps at room temperature, as displayed in Figure 1a and b. The mean minimum distance between H2O or O2 and arsenene is about 2.5 Å, which is larger than the generally bond length of As-H (1.56 Å) or As-O (1.71 Å). Therefore, the H2O or O2 molecules do not form any chemical bonding with arsenene. Moreover, in view of the degradation of BP deriving from oxidation,40 we computed the energy barrier of O2 dissociation on arsenene. The energy barriers are 1.08 and 1.37 eV for two possible dissociation paths as shown in Figure 1c, which are far higher than that of BP oxidation (0.56 eV).41 Obviously, the oxidation of perfect arsenene is hard to occur at room temperature. The high stability of arsenene derives from the fact that the d electrons in arsenene weaken the bonding ability of lone electron pair on each As atom, different from BP.42 Arsenene exhibits good photo-absorption ability as well due to the large oscillator strength of the low excited states as shown in Figure 1d. The low excited states are

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

active under incident beam polarized along the armchair direction and they are mainly contributed by the transition channel between the highest valence and the lowest conduction band at Г point and around Y point along the Y-Г direction. This outstanding optical property endows arsenene a potential candidate in photovoltaics and photocatalysis.

Figure 2. Comparison of band edges between experiment and DFT, G0W0 or extrapolated G0W0 approach for TCNE. The red and black lines or dot lines correspond to CBM or VBM values respectively. The magenta circle is the experiment value of TCNE (-3.1±0.2 eV). The inset is TCNE, which C and N atoms are labeled as grey and blue, respectively.

To construct type II heterostructure, the prerequisite is to acquire the accurate band edges of the corresponding building blocks. It is well-known that DFT is not sufficient to reproduce accurate band structure due to insufficient inclusion of many-body effect and generally underestimates the band gap of materials. Taking TCNE as an example, the difference of CBM between experiment and DFT is as large as 2.6 eV (see Figure 2). In principle, the many-body perturbation G0W0 approach can reproduce better band gap. However, accurate band edges are hard to obtain by single

ACS Paragon Plus Environment

Page 8 of 17

Page 9 of 17

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

ACS Applied Materials & Interfaces

G0W0 owing to the slow convergence of absolute band edges with respect to the number of unoccupied bands.

31-33

We show the convergence behavior of band gap,

CBM and VBM as a function of the number of unoccupied bands. Obviously, although the band gap can quickly achieve convergence around 1600 unoccupied bands, the absolute band edges are far from convergence. In fact, even with 3200 unoccupied bands, the VBM and CBM are not fully converged yet and the computational resource is consumed tremendously. Therefore, we use the extrapolation equation (1) to obtain the band edges at infinite unoccupied bands. As shown in Figure 2, the VBM and CBM are determined as -10.37 and -2.92 eV, respectively. The position of CBM using G0W0 with the extrapolation approach is close to the experiment value of -3.1±0.2 eV. 43 Moreover, compared to the band gap with 1600 (7.50 eV) or 3200 (7.48 eV) unoccupied bands, the extrapolated band gap does not change obviously (7.45 eV), which also confirms the reliability of the extrapolation approach. We adopt the same extrapolation approach to obtain the band edge positions of arsenene and other materials including MoS2, TiS3, TCNQ, TCNNQ and BV in Figure 3. Similarly, the band gaps quickly achieve convergence with the number of unoccupied bands of ~600 for 2D materials and ~1600 for organic molecules. Whereas the convergence of the absolute band edge positions is very slow, which is more evident for organic molecules. Therefore, the band edges are all determined by using G0W0 with the extrapolation approach.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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. Convergence of band edges with respect to the number of unoccupied bands and extrapolated value for (a) TCNQ, (b) TCNNQ, (c) BV, (d) MoS2, (e) TiS3 and (f) As based on G0W0 approach. Black and red points are the calculated band edges under G0W0 approach. (C, black-grey; N, blue; H, white; S, yellow; Mo, green; Ti, grey; As, orange)

When building the vdW heterostructures using supercells, the GW calculations are unaffordable for large sized supercells. Fortunately, the vdW interaction does not cause apparent shifting and hybridization in different building blocks. For example, as shown in Figure S1, the LUMO of TCNQ is only slightly perturbed in As/TCNQ heterostructure compared to the isolated TCNQ, and the energy level of band edges has a very small shift. Similarly, Compared with the isolated arsenene, the band edges of arsenene in As/TCNQ heterostructure is nearly the same. Therefore, we adopt the values of the band edges from isolated calculations as the reference.44 As clearly

ACS Paragon Plus Environment

Page 10 of 17

Page 11 of 17

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

ACS Applied Materials & Interfaces

illustrated in Figure 4, the arsenene can form type II heterostructures with other 2D or organic molecules except for TCNE. The As/TCNE forms a type I heterostructure due to TCNE having higher conduction band edge and lower valence band edge than arsenene and thus cannot avoid the recombination of electron-hole pairs. On the other hand, as a good photocatalyst, the band edges should straddle the water oxidation and reduction potentials. That is, the band gap of heterostructures should be larger than the minimum photocatalytic potential difference, 1.23 eV.45 The band gaps of As/MoS2, As/TiS3, As/TCNQ, As/TCNNQ and As/BV are 1.87, 1.17, 1.94, 1.69 and 1.22 eV, respectively, therefore, the As/TiS3, and As/BV heterostructures are filtered out as well. Therefore, we focus on As/MoS2, As/TCNQ and As/TCNNQ heterostructures in following discussion.

Figure 4. Band alignment of arsenene and other materials based on G0W0-extrapolation approach. The light red and light black are CBM and VBM of arsenene, respectively. The dark red and dark black are band edges of other materials.

For arsenene-based heterostructures paired with MoS2, or TCNQ, or TCNNQ, as illustrated in Figure 5a, the band edges of arsenene are higher than other materials,

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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 17

which is beneficial to the separation of electron-hole pairs. Specifically, under solar illumination, the photoexcited electron-hole pairs will be produced in these heterostructures. Then, the photogenerated electrons in the arsenene can be easily moved to the other layer of heterostructures (MoS2, or TCNQ, or TCNNQ), and holes are shifted in the opposite directions due to the difference of the band alignments. Moreover, the band edges of heterostructures straddle the water redox potential. Thus, the oxidation and redox reactions take place in different layer of heterostructures and the energy-wasting electron-hole recombination can be greatly suppressed.

Figure 5: (a) Schematic illustration of the carrier transfer and separation of arsenene-based heterostructures (As/MoS2, As/TCNQ and As/TCNNQ) in photocatalytic water splitting. Water redox potential positions are denoted by blue lines at pH = 7. (b) Power conversion efficiency contour. Arsenene severs as donor, and TCNNQ, MoS2 and TCNE are acceptors.

We further estimate the maximum power conversion efficiencies (η) of arsenene-based heterostructures paired with MoS2, TCNQ or TCNNQ as photovoltaic solar cells based on model developed by Scharber et al.46 J V β η = sc oc FF = Psolar

P( hϖ ) d ( hϖ ) hϖ ∞ ∫ P(hϖ )d (hϖ )

0.65( Egd − ∆Ec − 0.3) ∫



Eg

0

ACS Paragon Plus Environment

(2)

Page 13 of 17

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

ACS Applied Materials & Interfaces

Where 0.65 is the band filling factor, P(hϖ ) is taken to be AM1.5 solar energy flux d (expressed in W m-2 eV-1) at the photo energy hϖ , Eg is the band gap of donor, the

( Egd − ∆Ec − 0.3) term is an estimation of the maximum open circuit voltage Voc . The integral in the numerator is the short circuit current J sc using a limit external quantum efficiency of 100%, while the denominator is the integrated AM1.5 solar energy flux. In these heterostructures, arsenene is the donor due to its higher band edge, and MoS2, TCNQ or TCNNQ are acceptors. The maximum η values of As/TCNNQ, As/MoS2 and As/TCNQ are ~20%, 22% and 23% as marked in Figure 5b, respectively, which are comparable to the theoretically proposed monolayer or bilayer BP paired with transition metal dichalcogenides.

44, 47

The higher stability of

arsenene respective to BP may endow the arsenene-based heterostructures better potential in photovoltaic solar cells.

Conclusions We have explored a set of arsenene-based heterostructures by obtaining accurate band edges using exact GW approach combining an extrapolation procedure. Our results demonstrate that As/TCNNQ, As/MoS2 and As/TCNQ heterostructures possess type-II band alignment, suitable band edge positions straddling water redox potentials, good stability under ambient atmosphere and efficient photoabsorption, which makes them potential candidates as photocatalysts for water splitting. Simultaneously, the power conversion efficiency of these heterostructures can be as high as ~20% as photovoltaic solar cells. The bi-function, high efficiency and high environmental stability endow these arsenene-based heterostructures very promising in clean and

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

sustainable energy.

Supporting Information The following files are available free of charge. Illustration of the influence of vdW interaction on band edges of heterostructures.

Notes The authors declare no competing financial interest.

Acknowledgments This work is supported by the National Key R&D Program of China (Grant No. 2017YFA0204800), Natural Science Funds of China (21525311, 21373045, 21773027), Jiangsu 333 project (BRA2016353). The Scientific Research Foundation of Graduate School of Southeast University (YBJJ1620) and Jiangsu Innovation Projects for Graduate Student (KYZZ16_0117) in China. The authors thank the computational resources provided by Southeast University.

References 1 Qu Y.; Duan X., Progress, challenge and perspective of heterogeneous photocatalysts. Chem. Soc. Rev., 2013, 42, 2568-2580. 2 Kudo A.; Miseki Y., Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev., 2009, 38, 253-278. 3 Dresselhaus M. S.; Thomas I. L., Alternative energy technologies. Nature, 2001, 414, 332-337. 4 Liu Y.; Weiss N. O.; Duan X.; Cheng H.-C.; Huang Y.; Duan X., Van der Waals Heterostructures and Devices. Nat. Rev. Mater, 2016, 1, 16042. 5 Novoselov K. S.; Mishchenko A.; Carvalho A.; Castro Neto A. H., 2D Materials and van der Waals Heterostructures. Science, 2016, 353, aac9439. 6 Wu Z.; Neaton J. B.; Grossman J. C., Charge Separation via Strain in Silicon Nanowires. Nano Lett., 2009, 9, 2418-2422. 7 Zhang S.; Yan Z.; Li Y.; Chen Z.; Zeng H., Atomically Thin Arsenene and Antimonene: Semimetal-Semiconductor and Indirect-Direct Band-Gap Transitions. Angew. Chem., 2015, 127, 3155-3158. 8 Zhang S.; Xie M.; Li F.; Yan Z.; Li Y.; Kan E.; Liu W.; Chen Z.; Zeng H., Semiconducting Group 15 Monolayers:A Broad Range of Band Gaps and High Carrier Mobilities. Angew.Chem.Int. Ed. , 2016,

ACS Paragon Plus Environment

Page 14 of 17

Page 15 of 17

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

ACS Applied Materials & Interfaces

55, 1666 –1669. 9 Kecik D.; Durgun E.; Ciraci S., Optical properties of single-layer and bilayer arsenene phases. Phys. Rev. B, 2016, 94, 205410. 10 Kecik D.; Durgun E.; Ciraci S., Stability of single-layer and multilayer arsenene and their mechanical and electronic properties. Phys. Rev. B, 2016, 94, 205409. 11 Amani M.; Lien D.-H.; Kiriya D.; Xiao J.; Azcatl A.; Noh J.; R.Madhvapathy S.; Addou R.; KC S.; Dubey M.; Cho K.; Wallace R. M.; Lee S.-C.; He J.-H.; III J. W. A.; Zhang X.; Yablonovitch E.; Javey A., Near-unity photoluminescence quantum yield in MoS2. Science, 2015, 350, 1065-1068. 12 Splendiani A.; Sun L.; Zhang Y.; Li T.; Kim J.; Chim C. Y.; Galli G.; Wang F., Emerging photoluminescence in monolayer MoS2. Nano Lett., 2010, 10, 1271-1275. 13 Mouri S.; Miyauchi Y.; Matsuda K., Tunable photoluminescence of monolayer MoS2 via chemical doping. Nano Lett., 2013, 13, 5944-5948. 14 Dai J.; Zeng X. C., Titanium trisulfide monolayer: theoretical prediction of a new direct-gap semiconductor with high and anisotropic carrier mobility. Angew. Chem. Int. Ed., 2015, 54, 7572-7576. 15 Guilmeau E.; Berthebaud D.; Misse P. R. N.; Hébert S.; Lebedev O. I.; Chateigner D.; Martin C.; Maignan A., ZrSe3-Type Variant of TiS3: Structure and Thermoelectric Properties. Chem. Mater., 2014, 26, 5585-5591. 16 Li M.; Dai J.; Zeng X. C., Tuning the electronic properties of transition-metal trichalcogenides via tensile strain. Nanoscale, 2015, 7, 15385-15391. 17 Molina-Mendoza A. J.; Barawi M.; Biele R.; Flores E.; Ares J. R.; Sánchez C.; Rubio-Bollinger G.; Agraït N.; D'Agosta R.; Ferrer I. J.; Castellanos-Gomez A., Electronic Bandgap and Exciton Binding Energy of Layered Semiconductor TiS3. Adv. Electron. Mater., 2015, 1, 1500126. 18 Abdulsalam M.; Joubert D. P., Electronic and optical properties of MX3(M = Ti, Zr and Hf; X = S, Se) structures: A first principles insight. Phys. Status Solidi B, 2016, 253, 868-874. 19 Jing Y.; Tan X.; Zhou Z.; Shen P., Tuning electronic and optical properties of MoS2 monolayer via molecular charge transfer. J. Mater. Chem. A, 2014, 2, 16892-16897. 20 Chowdhury C.; Jahiruddin S.; Datta A., Pseudo-Jahn-Teller Distortion in Two-Dimensional Phosphorus: Origin of Black and Blue Phases of Phosphorene and Band Gap Modulation by Molecular Charge Transfer. J. Phys. Chem. Lett., 2016, 7, 1288-1297. 21 Abellán G.; Lloret V.; Mundloch U.; Marcia M.; Neiss C.; Görling A.; Varela M.; Hauke F.; Hirsch A., Noncovalent Functionalization of Black Phosphorus. Angew.Chem., 2016, 128, 14777-14782. 22 Nicht S.; Kleemann H.; Fischer A.; Leo K.; Lüssem B., Functionalized p-dopants as self-assembled monolayers for enhanced charge carrier injection in organic electronic devices. Org, Electron, 2014, 15, 654-660. 23 Momblona C.; Gil-Escrig L.; Bandiello E.; Hutter E. M.; Sessolo M.; Lederer K.; Blochwitz-Nimoth J.; Bolink H. J., Efficient vacuum deposited p-i-n and n-i-p perovskite solar cells employing doped charge transport layers. Energy Environ. Sci., 2016, 9, 3456-3463. 24 Lussem B.; Tietze M. L.; Kleemann H.; Hossbach C.; Bartha J. W.; Zakhidov A.; Leo K., Doped organic transistors operating in the inversion and depletion regime. Nat. Commun., 2013, 4, 2775. 25 Kiriya D.; Tosun M.; Zhao P.; Kang J. S.; Javey A., Air-stable surface charge transfer doping of MoS2 by benzyl viologen. J. Am. Chem. Soc., 2014, 136, 7853-7856. 26 Giannozzi P.; Baroni S.; Bonini N.; Calandra M.; Car R.; Cavazzoni C.; Ceresoli D.; Chiarotti G. L.; Cococcioni M.; Dabo I.; Dal Corso A.; de Gironcoli S.; Fabris S.; Fratesi G.; Gebauer R.; Gerstmann U.; Gougoussis C.; Kokalj A.; Lazzeri M.; Martin-Samos L.; Marzari N.; Mauri F.; Mazzarello R.;

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces

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

Paolini S.; Pasquarello A.; Paulatto L.; Sbraccia C.; Scandolo S.; Sclauzero G.; Seitsonen A. P.; Smogunov A.; Umari P.; Wentzcovitch R. M., QUANTUM ESPRESSO: a Modular and Open-source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter, 2009, 21, 395502. 27 Perdew J. P.; Burke K.; Ernzerhof M., Generalized Gradient Approximation Made Simple. Phys. Rev. Lett., 1996, 77, 3865-3868. 28 Goedecker S.; Teter M.; Hutter J., Separable dual-space Gaussian pseudopotentials. Phys. Rev. B, 1996, 54, 1703-1710. 29 Hedin L., New Method for Calculating the One-Particle Green's Function with Application to the Electron-Gas Problem. Phys. Rev., 1965, 139, A796-A823. 30 Ismail-Beigi S., Truncation of periodic image interactions for confined systems. Phys. Rev. B, 2006, 73, 233103. 31 Friedrich C.; Müller M. C.; Blügel S., Erratum: Band convergence and linearization error correction of all-electron GW calculations: The extreme case of zinc oxide [Phys. Rev. B 83, 081101(R) (2011)]. Phys. Rev. B, 2011, 84, 039906. 32 Friedrich C.; Müller M. C.; Blügel S., Band convergence and linearization error correction of all-electronGWcalculations: The extreme case of zinc oxide. Phys. Rev. B, 2011, 83, 081101(R). 33 Nabok D.; Gulans A.; Draxl C., Accurate all-electron G0W0 quasiparticle energies employing the full-potential augmented plane-wave method. Phys. Rev. B, 2016, 94, 035118. 34 Marini A.; Hogan C.; Grüning M.; Varsano D., yambo: An ab initio tool for excited state calculations. Comput. Phys. Commun., 2009, 180, 1392-1403. 35 Nosé S., A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys., 1984, 81, 511-519. 36 Bengtsson L., Dipole Correction for Surface Supercell Calculations. Phys. Rev. B, 1998, 59, 12301-12304. 37 Grimme S., Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem., 2006, 27, 1787-1799. 38 José M. S.; Emilio A.; Julian D. G.; Alberto G.; Javier J.; Pablo O.; Daniel S.-P., The SIESTA method for ab initioorder-Nmaterials simulation. J. Phys.: Condens. Matter, 2002, 14, 2745-2779. 39 Kamal C.; Ezawa M., Arsenene: Two-dimensional buckled and puckered honeycomb arsenic systems. Phys. Rev. B, 2015, 91, 085423. 40 Wang G.; Slough W. J.; Pandey R.; Karna S. P., Degradation of phosphorene in air: understanding at atomic level. 2D Mater., 2016, 3, 025011. 41 Ziletti A.; Carvalho A.; Campbell D. K.; Coker D. F.; Castro Neto A. H., Oxygen Defects in Phosphorene. Phys. Rev. Lett., 2015, 114, 046801. 42 Norman N. C., Chemistry of Arsenic, Antimony and Bismuth. Academic Press, Springer: 1998. 43 Chowdhury S.; Kebarle P., Electron Affinities of Di- and Tetracyanoethylene and Cyanobenzenes Based on Measurements of Gas-Phase Electron-Transfer Equilibria. J. Am. Chem. Soc, 1986, 108, 5453-5459. 44 Ganesan V. D. S. O.; Linghu J.; Zhang C.; Feng Y. P.; Shen L., Heterostructures of phosphorene and transition metal dichalcogenides for excitonic solar cells: A first-principles study. Appl. Phys. Lett., 2016, 108, 122105. 45 Lewerenz H. J.; Peter L., Photoelectrochemical Water Splitting. Springer New York: 2013. 46 Scharber M. C.; Mühlbacher D.; Koppe M.; Denk P.; Waldauf C.; Heeger A. J.; Brabec C. J., Design Rules for Donors in Bulk-Heterojunction Solar Cells-Towards 10 % Energy-Conversion Efficiency. Adv.

ACS Paragon Plus Environment

Page 16 of 17

Page 17 of 17

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

ACS Applied Materials & Interfaces

Mater., 2006, 18, 789-794. 47 Dai J.; Zeng X. C., Bilayer Phosphorene: Effect of Stacking Order on Bandgap and Its Potential Applications in Thin-Film Solar Cells. J. Phys. Chem. Lett., 2014, 5, 1289-1293.

TOC

ACS Paragon Plus Environment