Xiao-Hua Lia,
Bao-Ji Wang*a,
Xiao-Lin Caia,
Wei-Yang Yua,
Li-Wei Zhanga,
Guo-Dong Wanga and
San-Huang Ke*b
aSchool of Physics and Electronic Information Engineering, Henan Polytechnic University, Jiaozuo 454000, China. E-mail: wbj@hpu.edu.cn
bMOE Key Labortoray of Microstructured Materials, School of Physics Science and Engineering, Tongji University, Shanghai, 200092, China. E-mail: shke@tongji.edu.cn
First published on 15th September 2017
Vertical stacking of two-dimensional materials has recently emerged as an exciting method for the design of novel electronic and optoelectronic devices. In this work, we investigate the structural, electronic, and potential photocatalytic properties of arsenene/Ca(OH)2 van der Waals (vdW) heterostructures using first-principles calculations. It is found that all of the heterostructures are semiconductors with indirect band gaps and present similar electronic properties, almost irrespective of the stacking arrangement. However, among these heterostructures, the β-stacking heterostructure is found to be the most stable and its band gap and band edge position can be tuned by biaxial strain. In particular, comparing the band edge positions with the redox potentials of water shows that the strained β-stacking arsenene/Ca(OH)2 vdW heterostructure is a potential photocatalyst for water splitting. Meanwhile, this heterostructure exhibits significantly improved photocatalytic properties under visible-light irradiation by the calculated optical absorption spectra. Our findings provide a detailed understanding of the physical properties of arsenene/Ca(OH)2 vdW heterostructures and a new way to improve the design of photocatalysts for water splitting.
In addition, it is worth noting that these 2D materials offer a platform that allows creation of van der Waals (vdW) heterostructures comprised of at least two types of chemically different 2D materials, where the one layered 2D material are stacked on the other and held together by vdW forces.17,18 Such stacks are very different from the traditional 3D semiconductor heterostructures, as each layer acts simultaneously as the bulk material and the interface, reducing the amount of charge displacement within each layer. As a consequence, beyond being novel atomic-thick materials, vdW heterostructures have received considerable attention. Recently, a great deal of vertically-stacked vdW heterostructures have been realized experimentally and utilized to create electronic and optoelectronic devices with novel physical properties and distinctive capabilities.19–21 At the same time, many theoretical researchers have also shifted their focus from monocomponent systems to vdW heterostructures. The work by Su et al. showed that the indirect bandgap of arsenene/MoS2 heterostructures is tunable by changing the interlayer distance and increase with the increasing interlayer distance.22 A theoretical study23 demonstrated that the GaS/Ca(OH)2 heterobilayer is a type-II heterojunction where spatially separated charge carriers can be formed. The optical spectra of different stacking types exhibit distinct properties. Our previous work suggests that arsenene/GaS heterostructure is a promising photocatalyst for water splitting and its electronic properties can be continuously tuned by external strain.24 All these findings suggest that stacking 2D materials into vdW heterostructures provides an effective way to design novel artificial materials with special characteristics. This motivates us to consider whether arsenene and Ca(OH)2 monolayer can form a stable arsenene/Ca(OH)2 (A/C) vdW heterostructure, and what interesting electronic or optical properties it can provide. In this work, by performing first-principles calculations, we predict a strain dependence of the electronic and optical properties of the A/C vdW heterostructures. Band edge alignments of the A/C heterostructures under various strains are investigated with respect to water oxidation and reduction potentials. Our results reveal that A/C heterostructures could be a promising material for visible light photocatalysis and electronic and optoelectronic devices.
a = b (Å) | dAs–As (Å) | dH–O (Å) | dCa–O (Å) | Eb (meV per unit) | EPBEg (eV) | EHSEg (eV) | |
---|---|---|---|---|---|---|---|
Arsenene | 3.61 | 2.51 | 1.60 | 2.21 | |||
Ca(OH)2 | 3.58 | 0.97 | 2.36 | 3.67 | 5.17 | ||
α-Stacking | 3.60 | 2.51 | 0.97 | 2.37 | −166 | 1.59 | 2.11 |
β-Stacking | 3.60 | 2.51 | 0.97 | 2.37 | −171 | 1.60 | 2.22 |
γ-Stacking | 3.60 | 2.51 | 0.97 | 2.36 | −163 | 1.55 | 2.15 |
We consider three high-symmetry stacking structures named as α-, β- and γ-stacking, as depicted in Fig. 2(a–c). In the α-stacking, As atoms of the two sublattices are located on top of OH groups, while in the β (γ)-stacking one As sublattice is positioned on top of Ca atoms and the other is on top of upper (lower) OH group. To compare the relative stability of these heterostructures, we calculate their interface binding energies, Eb = EA/C − EA − EC, where EA/C, EA, and EC represent total energies of the A/C heterostructure, arsenene and Ca(OH)2 monolayer, respectively. By this definition, a lower Eb value means better stability of the heterostructure. As shown in Table 1, the predicted Eb values per unit cell for the three heterostructures are very close to each other and follow the order of Eβ (−171 meV) < Eα (−166 meV) < Eγ (−163 meV), which means that the three considered heterostructures possess very similar stability, although the β-stacking one is the most stable one. Moreover, the lattice parameters, bond lengths, and interlayer distances of the three A/C heterostructures are almost identical, resulting in very similar band structures (see Fig. 3(a)) and strain tunable electronic properties (see ESI: Fig. S1†). Similar behavior is also found in arsenene/MoS2 heterostructures22 and germanene/BeO heterostructures.38 Thus, in the following sections, only the most stable A/C heterostructures with β configuration is focused and discussed.
We first investigate the electronic properties of the A/C vdW heterostructure. The projected band structure given by the HSE06 calculation is illustrated in Fig. 3(b), in which the circles filled by red and green indicate the contributions from the arsenene and Ca(OH)2 layers, respectively. Obviously, the A/C heterostructure is a semiconductor with an indirect bandgap of 2.22 eV and its CBM and VBM are all mainly contributed from arsenene, indicating the formation of type-I heterostructure is formed. To show more details, the total and partial density of states (PDOS) are also plotted in Fig. 3(c). One can see that the states around VBM are mainly dominated by the p(x,y) states of As atoms, whereas the dominant contributions to the CBM originate from lone-pair pz states of As atoms, which induces the formation of the π-like states in the interlayer region and are beneficial to bring the arsenene and Ca(OH)2 layers together.22,39
To analyze the charge transfer between the arsenene and Ca(OH)2 layers, in Fig. 4(a) we plot the deformation charge density (DCD) Δρ(z) of the heterostructure along the z direction normal to the surface as , where ρA/C(x,y,z), ρA(x,y,z) and ρC(x,y,z) are the charge density at (x,y,z) points in the heterostructure, the pristine arsenene and pristine monolayer Ca(OH)2, respectively. Therefore, positive values indicate charge accumulation while negative values mean charge depletion. One can see that the Ca(OH)2 layer donates electrons to the arsenene layer, which leads to p-doping in Ca(OH)2 and n-doping in arsenene. The isosurface of the DCD is added in Fig. 4(a) as an inset, where the loss and the accumulation of electrons across the interface are directly depicted. The amount of transferred electrons up to z point is given by . The results in Fig. 4(a) (right y-axis) show that about 0.017|e| per unit cell is transferred from the Ca(OH)2 to arsenene layer, which is determined by the value of ΔQ(z) at the A/C interface (defined as the plane of zero charge variation as shown in the Δρ(z) curve).40 Fig. 4(b) shows the plane-averaged electrostatic potential along z-direction normal to the surface. The arsenene layer has a deeper potential than that of the Ca(OH)2 layer, driving electrons to move from the Ca(OH)2 layer to arsenene, which can be understood by considering the different electronegativities of the atoms. The potential drop (ΔVA/C) across the bilayer is found to be 4.68 eV. Such a large potential difference implies a strong electrostatic field across the interface, which may significantly influence the carrier dynamics and charge injection if the Ca(OH)2 is served as the electrode. In addition, the excitonic behavior of A/C heterostructure can be quite different from that of the isolated Ca(OH)2 layer as the gradient of the potential across the interface may facilitate the separation of electrons and holes.40
Tunability of electronic properties of a material by external control, such as strain, is very beneficial for its applications in nanoelectronics or optoelectronics.41 Here, we consider in-layer biaxial strain effects on the electronic structures of the A/C heterostructure, where the hexagonal unit cell is enlarged or shrunk symmetrically with specific ratios. Thus, 2D-isotropic deformation (ε) is defined as ε = (L − L0)/L0, where L and L0 are the lattice constants for the strained and unstrained structures, respectively. According to this definition, a negative value means a compressive strain, while a positive value indicates a tensile stress. The band-gap evolution given by the vdW-DFT/PBE under various biaxial strains is shown in Fig. 5(a) together with the strain energy. One can see that the bandgap reaches its maximum value (∼1.62 eV) at the strain of about +2% and then decreases with the compressive or tensile strain increasing. Then, the system goes from semiconducting to metallic for strains beyond −10%. We note that DFT/PBE usually underestimates bandgaps of semiconductors because of the lacking of derivative discontinuity in the energy functional. However, our purpose here is to see the trend in bandgap variation under different strains instead of the absolute values of bandgap, therefore the results can still be expected to be meaningful. This justification is also supported by previous calculations using the PBE functional and the HSE06 hybrid functional, showing consistent strain-induced effects on band structures.42
The dependencies of the highest valence and lowest conduction bands on the biaxial strains are shown in Fig. 5(b) and (c), which clearly demonstrate that the size and characteristic of the band-gap is determined by the strain-induced band-energy shifts at different k points. Similar to the HSE06 result, the free-standing heterostructure possesses an indirect bandgap between VBM at Γ point and CBM at k1(0.32, 0, 0) denoted by A in Fig. 5(b) or (c). Under an increasing small tensile strain, the VBM is located at Γ point while the energy of the conduction band bottom at Γ point shifts downwards more rapidly than the original CBM and becomes the new CBM, leading to an indirect–direct (I–D) transition at ε ≃ +2%. When the tensile strain is further increased (ε ≥ +4%), the CBM remains at Γ point, but the k2(0.1, 0.1, 0) denoted by B in Fig. 5(b) evolves gradually into the new VBM, which results in a D–I transition. In this case, the direct bandgaps at Γ point are very close in size to the indirect bandgaps (B–Γ) with a maximum energy difference of 0.07 eV. One can expect that the (quasi-)direct interband transitions at the Γ point improve the optical absorption (see later discussion). On the other hand, when an increasing compressive strain is applied, both the conduction and valence bands tend to shift downwards. The VBM is always located at Γ-point while the conduction band bottom at K point drops faster than A, and becomes equal in energy to A at ε ≃ −4.0%, and eventually becomes the new CBM when ε > −4.0%. Moreover, compressions greater than −10% induce a semiconductor–metal transition, where the bands cross the Fermi level at the K (conduction band) and Γ (valence band) points. It is interesting to note that there are two critical strains, +2.0% and −4.0%, under which degenerate conduction band valleys are created. This behavior would have potential applications for valleytronic devices: quasiparticles with the same energy at different positions in momentum space are less susceptible to phonon scattering.43,44
To check if all the strains considered are within the elastic limit, we calculate the strain energy per atom, Es = (Estrained − Eunstrained)/n, with n being the number of atoms in the unit cell. The results in Fig. 5(a) (right y-axis) show that Es varies smoothly as a quadratic function of the strain, indicating that the system is flexible and all the strains considered are within the elastic limit and, therefore, are fully reversible.
Since the bandgap of the A/C heterostructure are mechanically tunable, it is worthwhile exploring its potential for energy conversion applications, e.g., photocatalytic water splitting. To become a promising photocatalyst for water splitting, a material must satisfy two important criteria: (i) the bandgap of the semiconductor must be at least 1.6–1.7 eV and (ii) the band edges must straddle the redox potential energies of water, i.e., −4.44 eV and −5.67 eV for the hydrogen and oxygen evolution, respectively. Fig. 6 compares the band edge positions of the A/C heterostructure under the tensile biaxial strains considered with the redox potentials of hydrogen evolution (H+/H2) and oxygen evolution (O2/H2O) at pH = 0. As one can see, for the unstrained A/C heterostructure, the energies of the CBM and VBM relative to the vacuum level are −3.43 eV and −5.65 eV, respectively. Clearly, the VBM is unable to drive the oxygen evolution. However, a moderate tensile strain in the range from 1% to 6% reduces the VBM to become lower than the O2/H2O energy, rendering it useful in spontaneous photocatalytic water splitting. In particular, when ε ≃ +4% the bandgaps and band edge levels of the heterostructure become much match to the standard requirements for photocatalytic water splitting, which indicates that better photocatalytic performance might be expected in the A/C heterostructure with 4% tensile strain. Nevertheless, we notice that for the strains smaller than 6% the position of the VBM is very close to the water oxidation potential, implying a low oxidizing power. So a cocatalyst would be helpful to achieving the whole water splitting for this heterostructure.45 In fact, large pH values can make the redox reactions energetically more favorable, since the redox potentials of water increase with increased pH values, shifting the water's redox energy levels (as shown in Fig. 6) upward.46 When the tensile strain is further increased to be higher than 6%, the VBM will be raised above the oxidation potential of O2/H2O (being similar to the case of the unstrained heterostructure), making it unsuitable for water splitting anymore.
In addition to an appropriate bandgap and a suitable band alignment, a catalyst is also required to have outstanding optical adsorption in visible region. To this end, the optical absorbance A(ω) of the A/C heterostructure is determined as A(ω) = ωLImε(ω)/c and the results are shown in Fig. 7, where ω, L, Imε(ω), and c are the frequency of light, the length of the cell in the layer-normal direction, the imaginary part of the dielectric constant, and the speed of light in vacuum, respectively.47 Thereinto, Imε(ω) is calculated using the PBE functional. Following a previous studies,48–50 we compensate for the bandgap underestimation of the PBE functional by a rigid shift of the absorption curves upward by the value of the bandgap difference between the HSE06 and PBE functional. As can be seen, biaxial tensile strain red-shifts the optical spectra at uniform incremental intervals in the range of visible light, leading to significantly enhanced optical absorption in the region of [1.66 eV, 2.86 eV]. This result is related to the aforementioned results that strains cause bandgap reductions and a transition from an indirect to a (quasi-)direct,51 as shown in Fig. 5. Similar behaviors are also found in other two stacking cases although their optical absorptions are weaker than that in the β-stacking A/C heterostructures (see ESI: Fig. S2†). Finally, considering the fact that only few existing materials possess both a (quasi-)direct bandgap and suitable band edge positions, our prediction is important that the strained A/C heterostructures, especially the β-stacking one, is a good potential photocatalyst for water splitting.46
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c7ra08029h |
This journal is © The Royal Society of Chemistry 2017 |