Next Article in Journal
InterTwin: Deep Learning Approaches for Computing Measures of Effectiveness for Traffic Intersections
Next Article in Special Issue
Evaluation of Active Layer Thickness Influence in Long-Term Stability and Degradation Mechanisms in CsFAPbIBr Perovskite Solar Cells
Previous Article in Journal
An Analysis of the Use of Feed-Forward Sub-Modules for Transformer-Based Image Captioning Tasks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New van der Waals Heterostructures Based on Borophene and Rhenium Sulfide/Selenide for Photovoltaics: An Ab Initio Study

by
Michael M. Slepchenkov
1,
Dmitry A. Kolosov
1 and
Olga E. Glukhova
1,2,*
1
Institute of Physics, Saratov State University, Astrakhanskaya Street 83, 410012 Saratov, Russia
2
Laboratory of Biomedical Nanotechnology, I.M. Sechenov First Moscow State Medical University, Trubetskaya Street 8-2, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(24), 11636; https://doi.org/10.3390/app112411636
Submission received: 30 October 2021 / Revised: 22 November 2021 / Accepted: 6 December 2021 / Published: 8 December 2021
(This article belongs to the Special Issue Novel Organic-Inorganic Photovoltaic Materials)

Abstract

:
One of the urgent tasks of modern materials science is the search for new materials with improved optoelectronic properties for various applications of optoelectronics and photovoltaics. In this paper, using ab initio methods, we investigate the possibility of forming new types of van der Waals heterostructures based on monolayers of triangulated borophene, and monolayers of rhenium sulfide (ReS), and rhenium selenide (ReSe2), and predict their optoelectronic properties. Energy stable atomic configurations of borophene/ReS2 and borophene/ReSe2 van der Waals heterostructures were obtained using density functional theory (DFT) calculations in the Siesta software package. The results of calculating the density of electronic states of the obtained supercells showed that the proposed types of heterostructures are characterized by a metallic type of conductivity. Based on the calculated optical absorption and photocurrent spectra in the wavelength range of 200 to 2000 nm, it is found that borophene/ReS2 and borophene/ReSe2 heterostructures demonstrate a high absorption coefficient in the near- and far-UV(ultraviolet) ranges, as well as the presence of high-intensity photocurrent peaks in the visible range of electromagnetic radiation. Based on the obtained data of ab initio calculations, it is predicted that the proposed borophene/ReS2 and borophene/ReSe2 heterostructures can be promising materials for UV detectors and photosensitive materials for generating charge carriers upon absorption of light.

1. Introduction

The discovery of graphene and the identification of its unique properties stimulated the search for other 2D materials of atomic thickness with controllable properties, capable of overcoming the limitations inherent in graphene; in particular, providing an opening of the forbidden gap in the electronic structure of the material. Thus, the attention of researchers was drawn to isolated monolayers and multilayer crystals of hexagonal boron nitride (hBN), molybdenum disulfide (MoS2), other transition metal dichalcogenides (TMD), layered oxides, and elements of IV and V groups [1,2,3,4,5]. The interest in atomically thin two-dimensional (2D) layered structures has recently increased, in connection with their potential application in photonics, electronics, energy conversion, and storage, with the possibility of replacing traditional materials [6,7]. Along with the study of graphene-like 2D materials, the research field related to the design of vertical heterostructures, by stacking different 2D crystals on top of each other in a precisely selected sequence, has actively developed [8,9,10]. Strong covalent bonds ensure the stability of 2D crystals in the plane, and the action of van der Waals forces is sufficient to hold the stack of layers together. Such structures, called van der Waals heterostructures, are obtained in practice using the technique of direct mechanical assembly (micromechanical superposition) [11], as well as chemical vapor deposition (CVD) [12] and physical epitaxy [13] methods. The created van der Waals heterostructures find their application in the development of nano- and optoelectronic devices, including field-effect tunneling transistors [14], memory devices [15], photodetectors [16], solar cells [17], and rechargeable batteries [18]. Photovoltaics is considered one of the most promising fields of application for van der Waals heterostructures. Certain successes have already been achieved in the field of creating solar cells based on TMD heterostructures, built mainly on the basis of well-studied molybdenum disulfide (MoS2) and molybdenum diselenide (MoSe2), as well as tungsten disulfide (WS2) and tungsten diselenide (WSe2) [19,20,21,22,23]. However, the photoelectric cells created on the basis of MoS2 and WS2 are mainly focused on operating in a narrow region of the visible range of the electromagnetic spectrum, and only slightly cover the region of the near-infrared range [24,25,26,27]. At the same time, the emergence of new representatives of TMD, such as ReS2, ReSe2, TiS2, TiSe2, and HfS2 [28,29], exhibiting pronounced semiconductor properties, opens new possibilities in the direction of creating broadband photoelectric cells based on TMD.
As is known, when constructing a layered heterostructure, the key point is the matching of the crystal lattices of the combined 2D materials. Borophene, which is currently being successfully synthesized, can be considered a promising material for creating new 2D heterostructures [30,31]. Due to borophene polymorphism, various configurations of bonds between the boron atoms should relax the crystallographic lattice matching requirements for the combined 2D materials. In addition, the unique combination of properties of borophene, such as lightness, flexibility, mechanical strength, high conductivity, and optical transparency, makes it one of the most promising single-element 2D materials for nano- and optoelectronic purposes [32]. The first successes have already been achieved in the field of creating and studying borophene-based heterostructures. Liu and Hersam [33] reported on the successful experimental implementation of vertical borophene/graphene heterostructures. During synthesis in an ultrahigh vacuum chamber, graphene was first grown on a Ag (111) metal substrate at a higher temperature, and then boron was deposited on the same substrate. Despite the imperfect crystallographic lattice and symmetry matching, the created graphene/borophene heterostructures have almost perfect atomic consistency. An alternative approach to obtaining vertical heterostructures based on borophene and graphene was proposed by Hou et al. [34]. A new borophene/graphene heterostructure was obtained by heating a mixture of sodium borohydride and multilayer graphene, with the subsequent stepwise thermal decomposition of sodium borohydride, in situ in a high-purity hydrogen medium. Testing the humidity sensing behavior of the heterostructure under ambient conditions has shown that the borophene/graphene heterostructure exhibits high sensitivity (4200%) and a fast response/recovery time (10.5 s/8.3 s), with long-term stability and flexibility, which makes it suitable for creating highly efficient humidity sensors. Based on the results of ab initio calculations, Yu et al. [35] predicted that borophene/graphene heterostructures are a promising anode material, demonstrating a high adsorption energy of lithium (−2.959 eV) and a high theoretical specific capacity (1469.35 mA h/g). At the same time, the assessment of the prospects for the use of borophene-based van der Waals heterostructures for the creation of photovoltaic devices is only just beginning to be carried out. In 2020, Katoch et al. presented the results of an ab initio study of the prospects for using borophene in combination with MX2 (M = Mo, W and X = S, Se) to form metal-semiconductor contact. Scientists have shown that compressively strained borophene β12/MX2 heterostructures can be used to implement tunable Schottky barriers [36]. However, this topic needs further development.
In this paper, using ab initio methods, we study the possibility of constructing new types of van der Waals heterostructures based on 2D sheets of borophene and rhenium sulfide/selenide, and give a preliminary estimate of their optoelectronic properties.

2. Methods and Approaches

2.1. Calculation Details

The ab initio study of van der Waals heterostructures was carried out within the framework of the density functional theory (DFT) implemented in the Siesta 4.1.5 software package [37,38,39]. To describe exchange-correlation effects, generalized gradient approximation (GGA) was used in the parametrization of Perdew, Burke and Ernzerhof (PBE) [40]. To describe the van der Waals interaction (vdW) between heterostructure layers, we used the correction scheme proposed by Grimme, where vdW interactions are represented by a pair force field DFT + D2 [41]. When optimizing the geometry of the structure, we used the basis set of split valence orbitals DZ2P (double zeta 2 polarization), which includes polarization functions. The integration of the Brillouin zone was carried out using the Monkhorst-Pack scheme [42] with a 12 × 6 × 1 k-points mesh. The relaxation of the structure was carried out until the maximum interatomic force was less than 0.025 eV/Å. The effective Broyden-Pulay mixing scheme was used to minimize the energy of the electronic subsystem [43]. In order to avoid interaction between neighboring structures in calculations with periodic boundary conditions, the translation vector along the z-axis of the supercell was set to be greater than 20 Å. The real space mesh cutoff was chosen to be 300 Ry throughout the calculations.
To study the optical properties of van der Waals heterostructures, the complex frequency-dependent dielectric function was calculated in SIESTA software package. The calculations were carried out in the framework of the first-order time-dependent perturbation theory [44]. For this purpose, the self-consistent ground state energies and eigenfunctions were first calculated, and then they were used to calculate the transition dipole moment matrix elements. The imaginary part of the dielectric function determined the optical absorption, which was calculated from the transition rate between valance and conduction band states. The equation for calculating the absorption coefficient is as follows:
α ( ω ) = ω c n ( ω ) ε 2 ( ω ) ,
where ε2(ω) is the imaginary part of the complex dielectric function, n(ω) is the refractive index, c is the speed of light. The calculation of ε2(ω) was carried out according to the following equation:
ε 2 ( ω ) = e 2 π m 2 ω 2 ν , c B Z d k | ψ c k | e ^ , p | ψ ν k | 2 δ ( E c ( k ) E ν ( k ) ω ) ,
where the summation was carried out over each pair of states in the valence band (filled) and the conduction band (unoccupied), and integration was performed over all k-points in the Brillouin zone; the indices c and ν refer to the states in the conduction and valence bands, respectively, E(c, ν) (k) and ψ(c, ν),k are the energy and eigenfunction of these states, respectively. The electronic dipole transition matrix element is between the occupied and unoccupied states, where e ^ is the polarization vector and p is the momentum operator. The optical characteristics of the van der Waals heterostructures were calculated for two different directions of light polarization (vector E is parallel to the x-axis; vector E is perpendicular to the x-axis) for the energy range from 0.6 eV to 20 eV using optical broadening of 0.05 eV. The optical characteristics of the heterostructures under study were calculated by integrating the Brillouin zone with a 114 × 65 × 1 k-points mesh.
Based on the absorption spectrum, the photocurrent spectrum was calculated. The maximum value of the photocurrent is calculated by the following equation:
I m a x = e ω 1 ω 2 a ( ω ) P o w e r s o l a r ( ω ) h ν d ω   ,
where e is the electron charge; a(ω) is the absorption coefficient; Powersolar (ω) is the solar radiation power; hν is the energy of a quantum of solar radiation. Equation (3) is applicable for the case of an internal quantum yield of 100%, when each absorbed photon generates an electron.

2.2. Atomistic Models of van der Waals Heterostructures

In the framework of this study, the following new atomic configurations of van der Waals heterostructures were considered: (1) the configuration based on monolayers of borophene and rhenium sulfide; (2) the configuration based on monolayers of borophene and rhenium selenide. The formation of heterostructures was carried out with the following considerations: (1) the combined layers of the heterostructure should have a small discrepancy in the vectors of the crystal lattice and differ in the type of conductivity; (2) the atomic configuration of the formed heterostructure must be energetically stable. It is known that borophene is a conductor regardless of the type of allotropic modification [45]. On the contrary, rhenium sulfide (ReS2) and rhenium selenide (ReSe2), depending on the type of crystal system, can exhibit either a metallic character of conductivity, both in the case of a hexagonal crystal system in ReSe2 and a trigonal crystal system in ReS2, and also a semiconducting character of conductivity in the case of a triclinic crystal system for both ReSe2 and ReS2 [46]. In this regard, the unit cells of the ReSe2 and ReS2 of 12 atoms with a triclinic crystal system were chosen to construct the heterostructures. They were taken from the Materials Project open-source database [47]. Among the allotropic forms of borophene, we chose borophene with a triangulated configuration. The triangulated borophene has high energy stability and geometric features that allow it to be combined with ReSe2 and ReS2 layers. The unit cell of triangulated borophene consists of two atoms and has optimized translation vectors, Lx = 1.619 Å; Ly = 2.879 Å, which is in good agreement with the known data [45]. Since triangulated borophene has a crystalline structure with a rectangular unit cell, for its successful combination with the ReSe2 and ReS2 structures, it was necessary to construct rectangular unit cells for them. Such cells, each consisting of 24 atoms, were obtained from the initial unit cells with the triclinic system by the way clearly illustrated in Figure 1. The equilibrium configurations of the obtained unit cells have the following translation vectors: Lx = 6.411 Å; Ly = 11.359 Å for ReS2, and Lx = 6.646 Å; Ly = 11.824 Å for ReSe2. The performed calculations of the density of electronic states (DOS) of the triclinic and rectangular unit cells showed that when passing to a rectangular cell, the features of the electronic structure of ReS2 and ReSe2, primarily the presence of an energy gap, remain.
In order to minimize the lattice mismatch between the combined monolayers of borophene and ReS2/ReSe2, the sizes of the unit cells of each monolayer were increased until the difference in the parameters of the lattice vectors was no more than 2–3%. To build borophene/ReSe2 and borophene/ReS2 heterostructures, the initial unit cell of borophene 1 × 2 (the number of atoms along the X and Y axes) was increased 16 times to a size commensurate with rectangular supercells ReSe2 and ReS2. The final unit cell of borophene 8 × 32 was already 32 atoms. When a borophene monolayer was overlaid to a rectangular cell of ReSe2, the difference between the lengths of the translation vectors of the stacked monolayers was ~2.63% for the translation vector Lx and ~2.67% for translation vector Ly. When a borophene monolayer was overlaid to a rectangular cell of ReS2, the difference between the lengths of the translation vectors was even smaller, as follows: ~1% for Lx; ~1.38% for Ly. Figure 2 and Figure 3 show the process of stacking up monolayers during the formation of the borophene/ReS2 and borophene/ReSe2 heterostructures.
The optimized values of the translation vectors of the borophene/ReS2 heterostructure supercell were Lx = 6.485 Å and Ly = 11.404 Å; for the borophene/ReSe2 heterostructure supercell, Lx = 6.553 Å and Ly = 11.669 Å. The distance between the borophene and ReS2 layers along the z-axis was 2.969 Å, between the borophene and ReSe2 layers it was 3.028 Å.
To estimate the energy stability of the constructed supercells of the borophene/ReS2 and borophene/ReSe2 heterostructures, the binding energies Eb were calculated. The calculation was carried out according to the following equation:
E b   =   [ E B - t r / R e S 2 ( R e S e 2 )   E B - t r   E R e S 2 ( R e S e 2 )   ] / N ,
where EB-tr/ReS2(ReSe2) is the total energy of the borophene/ReS2 (borophene/ReSe2) heterostructure, E(B-tr) is the total energy of the isolated layer of triangulated borophene, EReS2 (ReSe2) is the total energy of the isolated ReS2(ReSe2) layer, N is the number of atoms in the heterostructure. According to the calculation results, the binding energy for both heterostructures is ~−0.05 eV/atom. A small value of Eb indicates that the layers in the heterostructure are bound by van der Waals forces. The negative value of the binding energy indicates that the structures are stable in energy, which means that they can be implemented in practice.

3. Results and Discussion

3.1. Electronic Structure of Borophene/ReS2 and Borophene/ReSe2 van der Waals Heterostructures

The next step in our study was the analysis of the electronic structure of the formed borophene/ReS2 and borophene/ReSe2 van der Waals heterostructures. To this end, calculations of the band structure were carried out, and the DOS distributions were constructed based on these calculation results. Figure 4 shows the calculated DOS distributions for both borophene/ReS2 (Figure 4a) and borophene/ReSe2 (Figure 4b) heterostructures, and for isolated monolayers, in order to reveal the regularities in the formation of the DOS profile for van der Waals heterostructures. As can be observed from the graphs in Figure 4, the contours of the DOS profiles of the borophene/ReS2 and borophene/ReSe2 heterostructures largely repeat the contours of the DOS profiles of the ReS2 and ReSe2 isolated monolayers, respectively. At the same time, both heterostructures demonstrate a metallic type of conductivity, as evidenced by the absence of an energy gap between the valence band and conduction band.
Therefore, we can say that borophene makes a decisive contribution to the type of conductivity of both heterostructures. In order to explain the obtained result, we calculated the distribution of the electron charge density over the atoms of the borophene/ReS2 and borophene/ReSe2 supercells according to the Mulliken procedure [48]. Based on the obtained distributions, it was revealed that a charge transfer occurred from borophene to ReS2/ReSe2. The total value of the transferred charge to the ReS2 layer was 0.76e, and to the ReSe2 layer, it was 0.11e. The difference in the value of the charge transferred by borophene is explained, on the one hand, by the difference in the distance between the borophene and ReS2/ReSe2 layers along the z-axis. In the case of the borophene/ReS2 heterostructure, the layers are closer to each other (2.96 Å), which means that the charge transfer is more intense. In the case of the borophene/ReSe2 heterostructure, the layers are farther apart (3.03 Å). On the other hand, the difference in the value of the charge transferred by borophene can be caused by the fact that sulfur has higher electronegativity, according to Mulliken (6.22 eV), as compared to selenium (5.89 eV).

3.2. Optoelectronic Properties of Borophene/ReS2 and Borophene/ReSe2 van der Waals Heterostructures

Next, we analyze the optoelectronic properties of borophene/ReS2 and borophene/ReSe2 van der Waals heterostructures. This analysis was carried out on the basis of the calculated absorption spectra in the wavelength range of electromagnetic radiation, from 200 to 2000 nm, and spectra of the photocurrent. The calculated absorption spectra are shown in Figure 5. To reveal the regularities in the formation of the profile of the absorption spectra of the heterostructures, the absorption spectra of isolated monolayers of borophene and ReS2/ReSe2 were calculated, which are also shown in Figure 5.
Analysis of the absorption spectra profiles showed that both heterostructures are characterized by the appearance of two noticeable peaks in the UV range. In the absorption spectrum of the borophene/ReS2 heterostructure, these peaks are located in the far (λ = 130 nm) and near-UV ranges (λ = 268 nm). Moreover, the peak in the near-UV region is due to the contribution of ReS2, and the peak in the far-UV region is due to the contribution of borophene. Consequently, there is a synergistic effect of the combination of borophene and ReS2, leading to an increase in the absorption capacity of the borophene/ReS2 heterostructure in the UV range. At the same time, in the near-IR region (0.75–2 µm), the absorption coefficient is minimal and is no more than 3%. The presence of absorption in this region is due to the contribution of borophene, since there is no absorption in ReS2 in the range 0.75–2 µm. A similar picture was observed for the borophene/ReSe2 heterostructure. One of the characteristic absorption peaks is located in the far-UV region (λ = 134 nm) and is due to the contribution of borophene, and the other peak is located in the near-UV region (λ = 300 nm) and is explained by the contribution of ReSe2. In the near-IR region (0.75–2 µm), absorption is due to the presence of borophene in the heterostructure, but the absorption coefficient is minimal and does not exceed 4%. Nevertheless, it should be noted that the borophene/ReS2 and borophene/ReSe2 heterostructures are characterized by greater absorption in the IR range, as compared, for example, with molybdenum disulfide MoS2 and molybdenum diselenide MoSe2, which absorb no more than 1% [49].
Based on the obtained absorption spectra and solar radiation spectra on the Earth’s surface (AM1.5) and outside the Earth’s atmosphere (AM0), the photocurrent spectra were calculated. The solar radiation spectra AM0 and AM1.5 were taken from the National Renewable Energy Laboratory (NREL) website [50], which shows solar spectra in the wavelength range of 280–2000 nm. The photocurrent spectra of the borophene/ReS2 and borophene/ReSe2 van der Waals heterostructures are shown in Figure 6. The photocurrent values are given for a surface area of 1 cm2. Figure 6 shows that the maximum value of the photocurrent for the borophene/ReS2 heterostructure falls at a wavelength of 600 nm, and is 9.36 mA cm−2 μm−1 in the case of the spectrum AM0 and 7.81 mA cm−2 μm−1 in the case of the spectrum AM1.5. For the borophene/ReSe2 heterostructure, the photocurrent maximum is shifted closer to the left edge of the visible range and falls at a wavelength of 480 nm. In the case of the AM0 spectrum, this value is 10.19 mA cm−2 μm−1, and in the case of the AM1.5 spectrum, it is 7.98 mA cm−2 μm−1. Thus, for both heterostructures, the peak of the photocurrent falls within the visible range of electromagnetic radiation.
One of the most important physical characteristics of optoelectronic devices is the magnitude of the integrated photocurrent or the density of the photocurrent. It is calculated as the integral of the photocurrent spectrum. Table 1 shows the calculated integral values of the photocurrent for the entire solar radiation spectrum, and for the visible range of the solar spectrum, 380–780 nm, as well as the maximum value of the photocurrent at a wavelength of 550 nm, which corresponds to the maximum power of solar radiation. For comparison, Table 1 also shows the calculated values of the above optoelectronic characteristics for graphene and borophene with the allotropic form β12.
Analysis of the data in Table 1 shows that the borophene/ReS2 and borophene/ReSe2 heterostructures are approximately 10 times higher than graphene and borophene β12, in terms of the integral value of the photocurrent in the visible range for the AM0 and AM1.5 spectra. The value of the maximum photocurrent at a wavelength of 550 nm is also higher in van der Waals heterostructures. In terms of this indicator, they exceed graphene by more than 4 times, and borophene β12 by more than 10 times. It should also be noted that the integral photocurrent of borophene/ReS2 and borophene/ReSe2 heterostructures in the visible range of radiation correlates well with the photocurrent density of the graphene/MoS2 heterostructure (3 mA/cm2) [51] and multilayer WS2 (4.10 mA/cm2) [52] obtained in a real experiment.

4. Conclusions

Based on the results of an ab initio study, the following conclusions can be drawn. The proposed atomic configurations of van der Waals borophene/ReS2 and borophene/ReSe2 heterostructures are energetically stable and characterized by a metallic type of conductivity. The synergistic effect of the combination of a triangulated borophene monolayer and ReS2/ReSe2 monolayers leads to an increased sensitivity of borophene/ReS2 and borophene/ReSe2 heterostructures to radiation in the UV range, which makes them potentially promising materials for the development of UV detectors in the near- and far-UV ranges. It is predicted that the efficiency of such detectors will be high, since both heterostructures are practically insensitive to other wavelengths. Based on the calculated photocurrent spectra in the wavelength range from 200 to 2000 nm, it was predicted that the borophene/ReS2 and borophene/ReSe2 heterostructures have a pronounced ability to generate a photocurrent in the visible wavelength range. It was shown that the integral value of the photocurrent in the visible radiation range of borophene/rhenium sulfide and borophene/rhenium selenide heterostructures is many times higher than the indices of monolayers of graphene and borophene β12, and correlates well with the data of full-scale experiments with other layered vertical 2D structures. At the same time, for the effective use of borophene/rhenium sulfide and borophene/rhenium selenide heterostructures as a material for creating solar cells, it is necessary to solve the problem of opening a band gap in their band structure. This task is the subject of further research.

Author Contributions

Conceptualization, M.M.S. and O.E.G.; methodology, M.M.S. and O.E.G.; funding acquisition, M.M.S.; investigation, O.E.G., D.A.K. and M.M.S.; writing—original draft preparation, D.A.K. and M.M.S.; writing—review and editing, O.E.G. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by Russian Science Foundation, grant number 21-72-00082.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mas-Ballesté, R.; Gómez-Navarro, C.; Gómez-Herrero, J.; Zamora, F. 2D materials: To graphene and beyond. Nanoscale 2011, 3, 20–30. [Google Scholar] [CrossRef]
  2. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-Like Two-Dimensional Materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef]
  3. Bhimanapati, G.R.; Lin, Z.; Meunier, V.; Jung, Y.; Cha, J.; Das, S.; Xiao, D.; Son, Y.; Strano, M.; Cooper, V.; et al. Recent Advances in Two-Dimensional Materials beyond Graphene. ACS Nano 2015, 9, 11509–11539. [Google Scholar] [CrossRef]
  4. Miró, P.; Audiffred, M.; Heine, T. An atlas of two-dimensional materials. Chem. Soc. Rev. 2014, 43, 6537–6554. [Google Scholar] [CrossRef]
  5. Kong, X.; Liu, Q.; Zhang, C.; Peng, Z.; Chen, Q. Elemental two-dimensional nanosheets beyond grapheme. Chem. Soc. Rev. 2017, 46, 2127–2157. [Google Scholar] [CrossRef]
  6. Xia, F.; Wang, H.; Xiao, D.; Dubey, M.; Ramasubramaniam, A. Two-dimensional material nanophotonics. Nat. Photon 2014, 8, 899–907. [Google Scholar] [CrossRef]
  7. Khan, K.; Tareen, A.; Aslam, M.; Wang, R.; Zhang, Y.; Mahmood, A.; Ouyang, Z.; Han, Z.; Guo, Z. Recent developments in emerging two-dimensional materials and their applications. J. Mater. Chem. C 2020, 8, 387–440. [Google Scholar] [CrossRef]
  8. Geim, A.K.; Grigorieva, I.V. Van der Waals heterostructures. Nature 2013, 499, 419–425. [Google Scholar] [CrossRef]
  9. Novoselov, K.S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A.H. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439-11. [Google Scholar] [CrossRef] [Green Version]
  10. He, J.; Wang, C.; Zhou, B.; Zhao, Y.; Tao, L.; Zhang, H. 2D van der Waals heterostructures: Processing, optical properties and applications in ultrafast photonics. Mater. Horiz. 2020, 7, 2903–2921. [Google Scholar] [CrossRef]
  11. Dean, C.; Young, A.; Meric, I.; Lee, C.; Wang, L.; Sorgenfrei, S.; Watanabe, K.; Taniguchi, T.; Kim, P.; Shepard, K.L.; et al. Boron nitride substrates for high-quality graphene electronics. Nat. Nanotechnol. 2010, 5, 722–726. [Google Scholar] [CrossRef]
  12. Robinson, J.A. Growing Vertical in the Flatland. ACS Nano 2016, 10, 42–45. [Google Scholar] [CrossRef]
  13. Liu, X.; Balla, I.; Bergeron, H.; Campbell, G.P.; Bedzyk, M.J.; Hersam, M.C. Rotationally Commensurate Growth of MoS2 on Epitaxial Graphene. ACS Nano 2016, 10, 1067–1075. [Google Scholar] [CrossRef] [Green Version]
  14. Li, J.; Chen, X.; Zhang, D.; Zhou, P. Van der Waals Heterostructure Based Field Effect Transistor Application. Crystals 2018, 8, 8. [Google Scholar] [CrossRef] [Green Version]
  15. Liu, C.; Zhou, P. Memory devices based on van der Waals heterostructures. ACS Mater. Lett. 2020, 2, 1101–1105. [Google Scholar] [CrossRef]
  16. Yan, F.; Hu, C.; Wang, Z.; Lin, H.; Wang, K. Perspectives on photodetectors based on selenides and their van der Waals heterojunctions. Appl. Phys. Lett. 2021, 118, 190501. [Google Scholar] [CrossRef]
  17. Furchi, M.M.; Höller, F.; Dobusch, L.; Polyushkin, D.K.; Schuler, S.; Mueller, T. Device physics of van der Waals heterojunction solar cells. NPJ 2D Mater. Appl. 2018, 2, 3. [Google Scholar] [CrossRef]
  18. King’ori, G.W.; Ouma, C.N.M.; Mishra, A.K.; Amolo, G.O.; Makau, N.W. Two-dimensional graphene–HfS2 van der Waals heterostructure as electrode material for alkali-ion batteries. RSC Adv. 2020, 10, 30127–30138. [Google Scholar] [CrossRef]
  19. Bellus, M.Z.; Li, M.; Lane, S.D.; Ceballos, F.; Cui, Q.; Zeng, X.C.; Zhao, H. Type-I van der Waals heterostructure formed by MoS2 and ReS2 monolayers. Nanoscale Horiz. 2017, 2, 31–36. [Google Scholar] [CrossRef]
  20. Hong, X.; Kim, J.; Shi, S.-F.; Zhang, Y.; Jin, C.; Sun, Y.; Tongay, S.; Wu, J.; Zhang, Y.; Wang, F. Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures. Nat. Nanotech. 2014, 9, 682–686. [Google Scholar] [CrossRef] [Green Version]
  21. Gong, Y.; Lin, J.; Wang, X.; Shi, G.; Lei, S.; Lin, Z.; Zou, X.; Ye, G.; Vajtai, R.; Yakobson, B.; et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 2014, 13, 1135–1142. [Google Scholar] [CrossRef] [Green Version]
  22. Yu, Y.; Hu, S.; Su, L.; Huang, L.; Liu, Y.; Jin, Z.; Purezky, A.; Geohegan, D.; Kim, K.; Zhang, Y.; et al. Equally Efficient Interlayer Exciton Relaxation and Improved Absorption in Epitaxial and Nonepitaxial MoS2/WS2 Heterostructures. Nano Lett. 2015, 15, 486–491. [Google Scholar] [CrossRef] [Green Version]
  23. Yuan, J.; Najmaei, S.; Zhang, Z.; Zhang, J.; Lei, S.M.; Ajayan, P.; Yakobson, B.; Lou, J. Photoluminescence Quenching and Charge Transfer in Artificial Heterostacks of Monolayer Transition Metal Dichalcogenides and Few-Layer Black Phosphorus. ACS Nano 2015, 9, 555–563. [Google Scholar] [CrossRef]
  24. Jia, Z.; Li, S.; Xiang, J.; Wen, F.; Bao, X.; Feng, S.; Yang, R.; Liu, Z. Highly sensitive and fast monolayer WS2 phototransistors realized by SnS nanosheet decoration. Nanoscale 2017, 9, 1916–1924. [Google Scholar] [CrossRef]
  25. Yoo, G.; Choi, S.; Park, S.; Lee, K.-T.; Lee, S.; Oh, M.S.; Heo, J.; Park, H.J. Flexible and Wavelength-Selective MoS2 Phototransistors with Monolithically Integrated Transmission Color Filters. Sci. Rep. 2017, 7, 40945. [Google Scholar] [CrossRef] [Green Version]
  26. Huo, N.; Yang, Y.; Li, J. Optoelectronics based on 2D TMDs and heterostructures. J. Semicond. 2017, 38, 031002. [Google Scholar] [CrossRef]
  27. Wang, F.; Wang, Z.; Yin, L.; Cheng, R.; Wang, J.; Wen, Y.; Shifa, T.A.; Wang, F.; Zhang, Y.; Zhan, X.; et al. 2D library beyond graphene and transition metal dichalcogenides: A focus on photodetection. Chem. Soc. Rev. 2018, 47, 6296–6341. [Google Scholar] [CrossRef]
  28. Hart, L.; Dale, S.; Hoye, S.; Webb, J.L.; Wolverson, D. Rhenium Dichalcogenides: Layered Semiconductors with Two Vertical Orientations. Nano Lett. 2016, 16, 1381–1386. [Google Scholar] [CrossRef] [Green Version]
  29. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  30. Mannix, A.J.; Zhou, X.-F.; Kiraly, B.; Wood, J.D.; Alducin, D.; Myers, B.D.; Liu, X.; Fisher, B.L.; Santiago, U.; Guest, J.R.; et al. Synthesis of borophenes: Anisotropic, two-dimensional boron polymorphs. Science 2015, 350, 1513–1516. [Google Scholar] [CrossRef] [Green Version]
  31. Feng, B.; Zhang, J.; Zhong, Q.; Li, W.; Li, S.; Li, H.; Cheng, P.; Meng, S.; Chen, L.; Wu, K. Experimental realization of two-dimensional boron sheets. Nat. Chem. 2016, 8, 563–568. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, Z.; Penev, E.S.; Yakobson, B.I. Two-dimensional boron: Structures, properties and applications. Chem. Soc. Rev. 2017, 46, 6746–6763. [Google Scholar] [CrossRef]
  33. Liu, X.; Hersam, M.C. Borophene-graphene heterostructures. Sci. Adv. 2019, 5, eaax6444. [Google Scholar] [CrossRef] [Green Version]
  34. Hou, C.; Tai, G.; Liu, B.; Wu, Z.; Yin, Y. Borophene-graphene heterostructure: Preparation and ultrasensitive humidity sensing. Nano Res. 2021, 14, 2337–2344. [Google Scholar] [CrossRef]
  35. Yu, J.; Zhou, M.; Yang, M.; Yang, Q.; Zhang, Z.; Zhang, Y. High-Performance Borophene/Graphene Heterostructure Anode of Lithium-Ion Batteries Achieved via Controlled Interlayer Spacing. ACS Appl. Energy Mater. 2020, 3, 11699–11705. [Google Scholar] [CrossRef]
  36. Katoch, N.; Kumar, A.; Sharma, R.; Ahluwalia, P.K.; Kumar, J. Strain Tunable Schottky Barriers and Tunneling Characteristics of Borophene/MX2 van der Waals Heterostructures. Phys. E Low Dimens. Syst. Nanostruct. 2020, 120, 113842. [Google Scholar] [CrossRef]
  37. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab-initio order-N materials simulation. J. Phys. Condens. Matt. 2002, 14, 2745–2779. [Google Scholar] [CrossRef] [Green Version]
  38. The SIESTA Group. Available online: Departments.icmab.es/leem/siesta/ (accessed on 10 April 2021).
  39. García, A.; Papior, N.; Akhtar, A.; Artacho, E.; Blum, V.; Bosoni, E.; Brandimarte, P.; Brandbyge, M.; Cerdá, J.I.; Corsetti, F.; et al. Siesta: Recent developments and applications. J. Chem. Phys. 2020, 152, 204108. [Google Scholar] [CrossRef]
  40. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  41. Grimme, S. Semiempirical GGA-type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  42. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  43. Pulay, P. Convergence acceleration of iterative sequences. The case of SCF iteration. Chem. Phys. Lett. 1980, 73, 393–398. [Google Scholar] [CrossRef]
  44. Economou, E.N. Green’s Functions in Quantum Physics, 3rd ed.; Springer: Berlin/Heidelberg, Germany, 1983; pp. 55–75. [Google Scholar]
  45. Luo, Z.; Fan, X.; An, Y. First-Principles Study on the Stability and STM Image of Borophene. Nanoscale Res. Lett. 2017, 12, 514. [Google Scholar] [CrossRef] [PubMed]
  46. Hafeez, M.; Gan, L.; Bhatti, A.S.; Zhai, T. Rhenium dichalcogenides (ReX2, X = S or Se): An emerging class of TMDs family. Mater. Chem. Front. 2017, 1, 1917–1932. [Google Scholar] [CrossRef]
  47. The Materials Project. Available online: https://materialsproject.org/ (accessed on 10 June 2021).
  48. Lu, H.; Dai, D.; Yanga, P.; Lic, L. Atomic orbitals in molecules: General electronegativity and improvement of Mulliken population analysis. Phys. Chem. Chem. Phys. 2006, 8, 340–346. [Google Scholar] [CrossRef]
  49. Maurizia, P.; Bernardi, M.; Grossman, J.C. Exciton Radiative Lifetimes in Two-Dimensional Transition Metal Dichalcogenides. Nano Lett. 2015, 5, 2794–2800. [Google Scholar]
  50. National Renewable Energy Laboratory (NREL). Available online: https://www.nrel.gov/ (accessed on 15 September 2021).
  51. Seo, D.-B.; Trung, T.N.; Bae, S.-S.; Kim, E.-T. Improved Photoelectrochemical Performance of MoS2 through MorphologyControlled Chemical Vapor Deposition Growth on Graphene. Nanomaterials 2021, 11, 1585. [Google Scholar] [CrossRef]
  52. Went, C.M.; Wong, J.; Jahelka, P.R.; Kelzenberg, M.; Biswas, S.; Hunt, M.S.; Carbone, A.; Atwater, H.A. A new metal transfer process for van der Waals contacts to vertical Schottky-junction transition metal dichalcogenide photovoltaics. Sci. Adv. 2019, 5, eaax6061. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Method for obtaining rectangular supercells of ReS2 (a) and ReSe2 (b) 2D structures. (Sulfur atoms are marked in yellow. Selenium atoms are marked in orange).
Figure 1. Method for obtaining rectangular supercells of ReS2 (a) and ReSe2 (b) 2D structures. (Sulfur atoms are marked in yellow. Selenium atoms are marked in orange).
Applsci 11 11636 g001
Figure 2. The process of constructing a supercell of van der Waals borophene/ReS2 heterostructures: (a) top view; (b) side view.
Figure 2. The process of constructing a supercell of van der Waals borophene/ReS2 heterostructures: (a) top view; (b) side view.
Applsci 11 11636 g002
Figure 3. The process of constructing a supercell of van der Waals borophene/ReSe2 heterostructures: (a) top view; (b) side view.
Figure 3. The process of constructing a supercell of van der Waals borophene/ReSe2 heterostructures: (a) top view; (b) side view.
Applsci 11 11636 g003
Figure 4. DOS of borophene/ReS2 (a) and borophene/ReSe2 (b) van der Waals heterostructures.
Figure 4. DOS of borophene/ReS2 (a) and borophene/ReSe2 (b) van der Waals heterostructures.
Applsci 11 11636 g004
Figure 5. Optical absorption coefficient of borophene/ReS2 (a) and borophene/ReSe2 (b) van der Waals heterostructures.
Figure 5. Optical absorption coefficient of borophene/ReS2 (a) and borophene/ReSe2 (b) van der Waals heterostructures.
Applsci 11 11636 g005
Figure 6. Photocurrent spectra of van der Waals borophene/ReS2 (a) and borophene/ReSe2 (b) heterostructures on the Earth’s surface (AM1.5) and outside the Earth’s atmosphere (AM0).
Figure 6. Photocurrent spectra of van der Waals borophene/ReS2 (a) and borophene/ReSe2 (b) heterostructures on the Earth’s surface (AM1.5) and outside the Earth’s atmosphere (AM0).
Applsci 11 11636 g006
Table 1. Some optoelectronic characteristics of van der Waals borophene/ReS2 and borophene/ReSe2 heterostructures.
Table 1. Some optoelectronic characteristics of van der Waals borophene/ReS2 and borophene/ReSe2 heterostructures.
Solar Spectrum AM0
The Integral Value of the Photocurrent for the Entire Solar Radiation Spectrum,
mA/cm2
Maximum Photocurrent at a Wavelength of 550 nm,
mA cm−2 μm−1
Integral Value of the Photocurrent for the Visible Spectrum 380–780 nm, mA/cm2
borophene/ReSe22.268.046.3
borophene/ReSe22.038.726.05
graphene2.11.850.677
borophene β121.320.7330.506
Solar Spectrum AM1.5
borophene/ReSe21.676.655.03
borophene/ReSe21.537.214.84
graphene1.51.540.559
borophene β121.00.610.433
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Slepchenkov, M.M.; Kolosov, D.A.; Glukhova, O.E. New van der Waals Heterostructures Based on Borophene and Rhenium Sulfide/Selenide for Photovoltaics: An Ab Initio Study. Appl. Sci. 2021, 11, 11636. https://doi.org/10.3390/app112411636

AMA Style

Slepchenkov MM, Kolosov DA, Glukhova OE. New van der Waals Heterostructures Based on Borophene and Rhenium Sulfide/Selenide for Photovoltaics: An Ab Initio Study. Applied Sciences. 2021; 11(24):11636. https://doi.org/10.3390/app112411636

Chicago/Turabian Style

Slepchenkov, Michael M., Dmitry A. Kolosov, and Olga E. Glukhova. 2021. "New van der Waals Heterostructures Based on Borophene and Rhenium Sulfide/Selenide for Photovoltaics: An Ab Initio Study" Applied Sciences 11, no. 24: 11636. https://doi.org/10.3390/app112411636

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop