Next Article in Journal
Effect of Chemical Composition of Metal–Organic Crosslinker on the Properties of Fracturing Fluid in High-Temperature Reservoir
Next Article in Special Issue
Binding Energies and Optical Properties of Power-Exponential and Modified Gaussian Quantum Dots
Previous Article in Journal
Overview on the Development of Electrochemical Immunosensors by the Signal Amplification of Enzyme- or Nanozyme-Based Catalysis Plus Redox Cycling
Previous Article in Special Issue
Ethylene Elimination Using Activated Carbons Obtained from Baru (Dipteryx alata vog.) Waste and Impregnated with Copper Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Dimensional GeC/MXY (M = Zr, Hf; X, Y = S, Se) Heterojunctions Used as Highly Efficient Overall Water-Splitting Photocatalysts

1
School of Electronic Information Engineering, Key Laboratory of Extraordinary Bond Engineering and Advanced Materials Technology of Chongqing, Yangtze Normal University, Chongqing 408100, China
2
School of Electronic Engineering, Xi’an University of Posts and Telecommunications, Xi’an 710121, China
3
School of Physical Science and Technology, Southwest University, Chongqing 400715, China
4
School of Electronic and Information Engineering, Anshun University, Anshun 561000, China
5
School of Physics, University of Electronic Science and Technology of China, Chengdu 610054, China
6
Science, Mathematics and Technology, Singapore University of Technology and Design, Singapore 487372, Singapore
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(12), 2793; https://doi.org/10.3390/molecules29122793
Submission received: 26 March 2024 / Revised: 16 May 2024 / Accepted: 21 May 2024 / Published: 12 June 2024
(This article belongs to the Special Issue Two-Dimensional Materials: From Synthesis to Applications)

Abstract

:
Hydrogen generation by photocatalytic water-splitting holds great promise for addressing the serious global energy and environmental crises, and has recently received significant attention from researchers. In this work, a method of assembling GeC/MXY (M = Zr, Hf; X, Y = S, Se) heterojunctions (HJs) by combining GeC and MXY monolayers (MLs) to construct direct Z-scheme photocatalytic systems is proposed. Based on first-principles calculations, we found that all the GeC/MXY HJs are stable van der Waals (vdW) HJs with indirect bandgaps. These HJs possess small bandgaps and exhibit strong light-absorption ability across a wide range. Furthermore, the built-in electric field (BIEF) around the heterointerface can accelerate photoinduced carrier separation. More interestingly, the suitable band edges of GeC/MXY HJs ensure sufficient kinetic potential to spontaneously accomplish water redox reactions under light irradiation. Overall, the strong light-harvesting ability, wide light-absorption range, small bandgaps, large heterointerfacial BIEFs, suitable band alignments, and carrier migration paths render GeC/MXY HJs highly efficient photocatalysts for overall water decomposition.

Graphical Abstract

1. Introduction

In recent years, environmental pollution and shortages of non-renewable energy have become increasingly severe. Photocatalytic water decomposition (PCWD) for hydrogen generation is considered as an effective approach to alleviating the energy crisis and environmental pollution [1,2,3,4,5,6,7,8,9,10]. In 1972, TiO 2 was first used as a photocatalyst (PC) to decompose water into hydrogen and oxygen [11]. The PCWD process typically comprises three steps: light capture, photoinduced carrier separation/transfer, and water redox reactions occurring on the surfaces of the catalysts [12,13,14]. The required water-splitting PC (WSPC) must have a band edge exceeding the water redox level; specifically, the H+/H2 and H2O/O2 levels must fall between the valence band maximum (VBM) and conduction band minimum (CBM) [15,16,17]. Consequently, the bandgap of a WSPC for PCWD should be greater than 1.23 eV. Furthermore, considering the energy loss during the process of photoinduced carriers transferring to the catalyst’s surfaces and the kinetic potentials required to drive the water redox reactions, the bandgap of the WSPC is typically required to be greater than 1.8 eV [12,18]. In addition to the bandgap requirement, the activity of WSPCs strongly depends on other factors such as photostability, light capture ability, trapping of and photoinduced carrier recombination, and the catalyst’s surface reactivities towards the hydrogen/oxygen evolution reaction (HER/OER) [19]. Although researchers have developed a series of PCs; a few single WSPCs simultaneously possess the advantages of wide light response extent, good carrier mobility, high photoexcited carrier separation, spatially separated reaction sites, strong redox capacity, and lower overpotentials for the HER and OER processes. Thus, it is still urgent to explore new photocatalytic mechanisms and develop highly efficient WSPCs.
Inspired by photosynthesis in green plants, a direct Z-scheme mechanism was constructed to overcome the shortcomings of single WSPCs [20,21,22]. A typical direct Z-scheme WSPC is usually composed of two parts: the hydrogen production PC (HPPC) and oxygen production PC (OPPC) [23,24]. The photoinduced electrons and holes recombine at the interface between the HPPC and OPPC, resulting in remnant electrons at the HPPC and excess holes at the OPPC. This process leads to the efficient spatial separation of photogenerated carriers, thus obtaining strong redox capacity to drive water-splitting. Up to now, the direct Z-scheme mechanism has been experimentally realized in a series of composites, including TiO 2 /ZnIn2S4 [25], aza-CMP/C2N, [26], Cd0.5Zn0.5/BiVO 4 [27], α -Fe2O3/g-C3N4 [28], black P/BiVO 4 [29], CdS/MoS 2 [30], CdS/Co1−xS [31], and TiO 2 /CuO 2 [32]. In particular, direct Z-scheme two-dimensional (2D) van der Waals (vdW) heterojunction (HJ) PCs (HJPCs) exhibit excellent photocatalytic performance due to their highly specific surface area, abundant active sites, good carrier mobility, and tunable interfaces [26,33,34]. In addition, strong electron–hole coupling and charge transfer around such heterointerfaces have been experimentally observed [35,36,37] and theoretically proposed [38,39,40,41,42]. This facilitates interlayer carrier recombination and helps to achieve the Z-scheme photocatalytic mechanism.
The graphene-like hexagonal GeC monolayer (ML) receives considerable attention due to its excellent electronic, mechanical, magnetic, and optical properties [43,44,45,46]. Especially, it possesses lower stiffness and a bigger Poisson’s ratio compared to graphene [47]. Therefore, the excellent characteristics of the GeC ML promote it to achieve device applications in the fields of electronics, optoelectronics, and photovoltaic [48]. More excitingly, GeC thin films have been experimentally synthesized by the chemical vapor deposition (CVD) [49] and laser ablation [50] methods. Since it is known that a large variety of 2D layers can be fabricated using the mechanical exfoliation and CVD [51,52,53] methods, we can speculate that the GeC ML may also be synthesized using similar preparation methods. The GeC ML not only has a stable plane structure, but also shares a similar honeycomb structure and lattice constants with many other 2D materials; thus, some GeC-based HJs have been designed and studied [54,55,56,57,58,59,60,61,62,63]. Although many literature works have explored the photocatalytic performance of the GeC ML and GeC-based type-II HJPCs, there are still relatively few reports on the direct Z-scheme mechanism of GeC-based HJs, which remains an open question thus far. It is interesting and meaningful to find suitable 2D materials to construct direct Z-scheme HJPCs with GeC MLs. Recently, MXY (M = Zr, Hf; X, Y = S, Se) MLs with a stable 1T phase have been demonstrated to exhibit excellent mechanical, thermal, thermoelectric, piezoelectricity, optical, and catalytic properties [64,65,66,67,68,69,70]. In particular, HfS 2 , HfSe 2 , ZrS 2 , and ZrSe 2 MLs have been experimentally verified [71,72,73,74,75], as well as Janus MoSSe and WSSe, which have been experimentally synthesized [76,77,78]. We can speculate that the Janus HfSSe and ZrSSe could be potentially fabricated by selenizing HfS 2 (or HfSe 2 ) and ZrS 2 (or ZrSe 2 ) MLs, respectively, using the CVD method, which is similar to the method used for synthesizing MoSSe and WSSe MLs. Moreover, the photocatalytic properties of MXY-based HJs have also been explored [79,80,81,82,83,84,85,86]. However, the photocatalytic properties of HJs composed of GeC and MXY MLs have not been reported yet. Therefore, we expect to combine GeC MLs and MXY MLs to construct highly efficient direct Z-scheme HJPCs.
Theoretical calculation is a simple and effective way to screen and design potential direct Z-scheme WSPCs [87,88,89,90,91]. From density functional theory (DFT) calculations, one can determine whether a type-II HJ exhibits a direct Z-scheme or type-II photocatalytic mechanisms based on the carrier migration path, judged according to the built-in electric field (BIEF) direction [28,92]. If the BIEF promotes interlayer carrier recombination, the carrier transfer belongs to a Z-scheme mechanism. Otherwise, the type-II mechanism dominates. Herein, first-principles calculations are performed to explore the possibility of constructing GeC/MXY HJs using GeC and MXY MLs as direct Z-scheme systems. Work functions ( Φ ) and charge density differences (CDDs) indicate that the BIEFs promote all eight GeC/MXY HJs to form the Z-scheme photocatalytic mechanism for overall water-splitting. Furthermore, all these HJs possess strong visible light-absorption capacity and substantial near-infrared light-absorption capacity. Moreover, these HJs can provide sufficient driving forces to overcome the HER and OER overpotentials to perform overall water redox reactions. These results are expected to guide experiment progress in exploring 2D direct Z-scheme WSPCs.

2. Results and Discussion

Before constructing GeC/MXY HJs, the geometric and electronic properties of GeC and MXY MLs are first investigated. The corresponding structural models for GeC, ZrS 2 , ZrSe 2 , ZrSSe, HfS 2 , HfSe 2 , and HfSSe MLs are plotted in Figure S1. The obtained E g values for GeC, ZrS 2 , ZrSe 2 , ZrSSe, HfS 2 , HfSe 2 , and HfSSe MLs are, respectively, 2.87, 2.02, 1.19, 1.46, 2.13, 1.33, and 1.56 eV, and the corresponding lattice parameters are, respectively, 3.235, 3.685, 3.800, 3.743, 3.645, 3.768, and 3.705 Å (see Table 1). Furthermore, the bond lengths of Ge–C in GeC, Zr–S in ZrS 2 , Zr–Se in ZrSe 2 , Zr–S (or Zr–Se) in ZrSSe, Hf–S in HfS 2 , Hf–Se in HfSe 2 , and Hf–S (or Hf–Se) in HfSSe are 1.868, 2.574, 2.706, 2.568 (or 2.713), 2.552, 2.685, and 2.550 (or 2.687) Å, respectively (see Table 1). GeC possesses a direct bandgap with both the VBM and CBM located at the K point, while all the MXY MLs are indirect bandgap semiconductors with the VBM and CBM, respectively, located at the Γ and M points (see Figure S2). All these results agree well with previous reports [54,55,56,57,58,64,65,66,67,93], as displayed in Table 1, indicating that our calculations are reliable.
Although the lattice constants of GeC and MXY are obviously different, the 2 × 2 GeC supercell could match well with the 3 × 3 MXY supercell. Herein, we define the lattice mismatch as [2 × | L sGeC L sMXY | / ( L sGeC + L sMXY ) ] × 100%, where L sGeC and L sMXY are the lattice constants for the GeC and MXY supercells, respectively. The calculated lattice mismatches between GeC and MXY to construct various GeC/MXY HJs are 1.38%, 1.71%, 0.20%, 0.20%, 2.47%, 0.87%, 0.82%, and 0.82%, respectively. These small lattice mismatches are favorable for the direct growth of GeC/MXY HJs by CVD or physical epitaxy [94]. Considering that the ZrSSe (or HfSSe) ML possesses two different surfaces, we loaded 3 × 3 ZrS 2 , ZrSe 2 , ZrSSe, HfS 2 , HfSe 2 , and HfSSe MLs onto a 2 × 2 GeC ML to construct eight different GeC/MXY HJs, i.e., GeC/ZrS 2 , GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS. The corresponding models of GeC/MXY HJs are shown in Figure 1. Note that, here, the average of the lattice parameters for GeC and MXY is used to build GeC/MXY HJs, and the lattice parameters for GeC/MXY HJs are illustrated in Table 2.
The thermodynamic stability of GeC/MXY HJs is assessed by calculating the interface formation energies ( E f ) as follows:
E f = ( E GeC / MXY T E GeC T E MXY T ) / S ,
where E GeC / MXY T , E GeC T , and E MXY T , respectively, denote the total energies of GeC/MXY HJs, GeC ML, and MXY ML. The calculated E f values for all the considered HJs are negative, which means that the construction of all these GeC/MXY HJs release heat and tend to be thermodynamically stable. The E f values in Table 2 range from −29.4 to −18.5 meV/Å 2 , suggesting that these HJs are formed via the interaction between vdW and the MLs [95]. Moreover, the interlayer distances of GeC/MXY HJs vary from 3.367 to 3.519 Å (see Table 2), aligning with the results of some other typical vdW structures [96,97,98,99,100]. Consequently, all the examined GeC/MXY structures are classified as vdW HJs. The interfacial formation energy can be directly defined as: E f = E GeC / MXY T E GeC T E MXY T . In this case, the E f values for GeC/ZrS 2 , GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS HJs are −0.66, −1.04, −1.05, −1.07, −0.89, −1.04, −1.00, and −1.03 eV, respectively. The E f value of the GeC/SZrSe (or GeC/SHfSe) HJ is sightly more negative than that of the GeC/SeZrS (or GeC/SeHfS) HJ, indicating that the formation of the GeC/SZrSe (or GeC/SHfSe) HJ is energetically slightly more favorable. During experimental preparation, both GeC/SZrSe (or GeC/SHfSe) and GeC/SeZrS (or GeC/SeHfS) HJs are likely to be prepared. The difference in their preparation lies in the fact that the Janus ZrSSe (or HfSSe) ML contacts the GeC ML with different surfaces.
The band structures for various GeC/MXY HJs, computed using the HSE06 hybrid functional, are illustrated in Figure 2. The orange color denotes the contribution from the GeC layer, while the green color represents the contribution from the MXY layers. All the GeC/MXY HJs are indirect bandgap semiconductors, as their VBMs are located at the K point, while the CBMs are positioned at the M point. The corresponding bandgaps for GeC/MXY HJs are 0.45 (0.446), 0.45 (0.453), 0.55, 0.43, 0.53, 0.59, 0.66, and 0.54 eV, respectively (see Table 3), which are significantly lower than those of the corresponding MLs. Consequently, these HJs are expected to achieve high solar energy utilization. The CBMs originate from the MXY layer, whereas the VBMs come from the GeC layer, confirming the staggered type-II nature of all the examined GeC/MXY HJs. These facilitate the spatial separation of the photoinduced carriers. Furthermore, the band alignments of the GeC and MXY layers in the HJs retain the primary characteristic of their isolated MLs, suggesting that the vdW interaction at the heterointerface does not significantly influence the electronic properties of the layers.
The work function ( Φ ) and CDD are crucial in determining the BIEF direction at the heterointerface, a factor that holds a decisive role in the design of Z-scheme PCs [101]. The Φ values can be obtained as follows:
Φ = E vac E F ,
where E vac and E F refer to the vacuum and Fermi energy levels, respectively. The electrostatic potentials (EPs) of the relative MLs and HJs are depicted in Figure 3 and Figure S3. The vacuum levels of the two surfaces in GeC, ZrS 2 , ZrSe 2 , HfS 2 , and HfSe 2 are identical, meaning that the electrostatic potential differences (EPDs) ( Δ E ) between the two sides are all zero. The corresponding Φ values for these MLs are 4.80, 6.55, 5.54, 6.46, and 5.68 eV, respectively. The difference in the electronegativity between the S and Se atoms at the two opposing sides of the Janus ZrSSe and HfSSe results in inherent BIEFs perpendicular to the plane, causing the vacuum levels on both surfaces to differ. Consequently, the work functions are naturally distinct on the surfaces of both ZrSSe and HfSSe. The corresponding Δ E values are 0.13 and 0.11 eV for ZrSSe and HfSSe, respectively. The Φ values for the S-side (Se-side) for ZrSSe and HfSSe are 6.07 (5.93) and 5.95 (5.84) eV, respectively. Evidently, GeC exhibits a lower Φ value compared to the MXY MLs. Once GeC and MXY come into contact to form a GeC/MXY HJ, electrons will transfer from the material with a lower work function to the one with a higher work function until dynamic equilibrium is achieved. Consequently, a BIEF is established across the GeC/MXY heterointerface, pointing from GeC towards MXY. Additionally, the vacuum levels on both sides of the GeC/MXY HJs also differ. The calculated values Δ E for the various GeC/MXY HJs are 0.16, 0.15, 0.25, 0.05, 0.09, 0.10, 0.20, and 0.02 eV, respectively. Furthermore, the work functions for the respective GeC/MXY HJs are 5.12 (5.29), 5.13 (5.28), 5.08 (5.34), 5.12 (5.17), 5.06 (5.14), 5.05 (5.15), 5.00 (5.20), and 5.04 (5.06) eV. This indicates that, when GeC and MXY contact to form a GeC/MXY HJ, electrons migrate from GeC to MXY to reach the same Fermi level.
Moreover, we analyzed the charge transfer at the heterointerface region in GeC/MXY HJs by calculating the visual charge density difference (VCDD) based on the following relationship [102,103]:
Δ ρ = ρ GeC / MXY ρ GeC ρ MXY ,
where ρ GeC / MXY , ρ GeC , and ρ MXY represent the charge densities for the GeC/MXY HJ, GeC ML, and MXY ML, respectively. The yellow (or cyan) region denotes charge accumulation (or consumption). Additionally, the planar-averaged CDD (PACDD) along the z-direction is obtained by the following equation [103,104]:
Δ ρ ( z ) = ρ GeC / MXY dxdy ρ GeC dxdy ρ MXY dxdy ,
where ρ GeC / MXY dxdy , ρ GeC dxdy , and ρ MXY dxdy represent the planar-averaged charge densities of the GeC/MXY HJ, GeC ML, and MXY ML, respectively. The positive (or negative) value indicates the charge accumulation (or consumption). It can be clearly seen from Figure 4 that the charge around the heterointerfaces of all the GeC/MXY HJs is redistributed. Charge accumulation primarily occurs at the heterointerface region near MXY, while charge consumption mainly takes place at the heterointerface region near GeC. This further confirms that electrons migrate from GeC to MXY in all the GeC/MXY HJs. The Bader charge analysis also suggests that 0.11, 0.09, 0.11, 0.09, 0.08. 0.07, 0.08, and 0.07 e, respectively, migrate from GeC to MXY in the various GeC/MXY HJs. The charge transfer at the heterointerfaces of GeC/MXY HJs could cause the BIEF to point away from GeC toward MXY, which commonly promotes the spatial separation of carriers, thus extending the lifetime of photoexcited carriers and enhancing the photocatalytic activity.
As is well known, the type-II band alignment corresponds to both the type-II and direct Z-scheme photocatalytic mechanisms based on different charge transfer pathways. For a type-II HJPC, the band edges of its two components must simultaneously straddle the water redox potentials. Thus, a type-II HJPC usually does not provide sufficient driving force for water redox processes. For a direct Z-scheme HJPC, the VBM of one component should be lower than the water oxidation potential (WOP), while the CBM of the other component should be higher than the water reduction potential (WRP) [105,106]. Thus, a direct Z-scheme HJPC is usually capable of proving sufficient driving force for redox reactions. Next, we arrange the band edges of the GeC ML, MXY MLs, and GeC/MXY HJs in contrast to the water redox levels in Figure 5 and Figure S4, in order to further determine the photocatalytic mechanisms of the considered GeC/MXY HJs. It is known that the water redox levels are determined by the electrochemical potentials relative to the vacuum level, so the difference in the vacuum energy levels on the two surfaces of PCs causes the movement of H + /H 2 and H2O/O2 levels between the two surfaces. The band edges of GeC only span the WOP, indicating that GeC is only suitable for the HER. Conversely, the band edges of MXY MLs solely cross the WRP, meaning that MXY MLs only serve for the OER. Thus, neither GeC nor MXY alone could achieve overall PCWD.
Since the photocatalytic mechanisms for all GeC/MXY HJs are similar, we will use GeC/ZrS2 as an illustrative example for a detailed discussion. As the GeC ML and ZrS2 ML approach each other to form the GeC/ZrS2 HJ, electrons migrate from GeC to ZrS2 due to the smaller work function of GeC compared to ZrS2. Consequently, the GeC and ZrS2 layers become positively and negatively charged, respectively. This results in a BIEF that is directed away from GeC toward ZrS2 across the GeC/ZrS2 heterointerface. Electrons in GeC are repelled by the negatively charged ZrS2, causing GeC’s bands to bend upward. Similarly, ZrS2’s bands will bend downward near the heterointerface due to the same mechanism [107,108]. To simplify the discussion, we have omitted the band bending in the band-alignment diagram of GeC/MXY HJs. When exposed to sunlight, both GeC and ZrS2 can absorb photons with greater energy than their respective bandgaps. This stimulates electrons to transition from the valence bands (VBs) to the conduction bands (CBs), leaving holes in the VBs. However, GeC is unsuitable for the OER due to its higher VBM than the WOP, while ZrS2 is unsuitable for the HER owing to its lower CBM than the WRP. This implies that the photoinduced holes in the VBs of GeC (or the photoexcited holes in the CBs of ZrS2) cannot directly participate in the OER (or HER) process. The calculated conduction band offset (CBO) and valence band offset (VBO) are 2.59 and 1.70 eV, respectively (see Table 3). Due to the BIEF directed from GeC to ZrS2, the migration of photoinduced electrons from the CBs of GeC to the CBs of ZrS2 and the migration of photoinduced holes from the VBs of ZrS 2 to the VBs of GeC are hindered. Conversely, the photoexcited electrons are encouraged to migrate from the CBs of ZrS 2 to the VBs of GeC, where they recombine with the photoexcited holes. Furthermore, the interlayer bandgap of 0.45 eV is significantly smaller than both the CBO and the VBO, favoring the interlayer electron–hole recombination. Consequently, GeC (or ZrS 2 ) accumulates more photoinduced electrons (or holes). Naturally, the superfluous electrons on the CBs for GeC can achieve the HER, while the excess holes on the VBs of ZrS 2 can realize the OER. Evidently, the migration path of photoexcited carriers is like a “Z”. Thus, the GeC/ZrS 2 HJ constitutes a direct Z-scheme system. The spatial separation of photoexcited electrons and holes contributes to enhancing photocatalytic efficiency. Additionally, schematic diagrams illustrating the photocatalytic mechanisms of all considered MLs and HJs versus the normal hydrogen electrode (NHE) are presented in Figures S5 and S6.
Furthermore, the sufficient kinetic potentials ( U e and U h ) provided by photoexcited electrons and holes are crucial for driving the OERs and HERs. The U e and U h values affect the number of active electrons and holes participating in water redox reactions, thereby influencing the photocatalytic activity. Here, U e (or U h ) is defined as the potential difference between the CBM and the H + /H 2 level (or between the H + /H 2 level and the VBM). Given that the water redox levels depend on the p H values, U e and U h can be expressed as follows [109]:
U e = U e ( p H = 0 ) p H × 0.059 V , U h = U h ( p H = 0 ) + p H × 0.059 V .
For the sake of simplicity, we will only discuss the U e (or U h ) value at p H = 0. The calculated U e and U h values are 2.14 and 2.75 V, respectively, which are comparable to some previously studied Z-scheme PCs (see Figure 6) [24,40,41,42,89,90,91,109]. Consequently, GeC/ZrS 2 HJ emerges as a highly efficient Z-scheme WSPC.
The CBO (VBO) values for GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS HJs are 2.30 (0.68), 2.19 (0.92), 2.32 (1.07), 2.34 (1.79), 2.18 (0.71), 2.15 (0.99), and 2.28 (1.12) eV, respectively (see Table 3). Obviously, the bandgaps of these HJs are smaller than their CBOs and VBOs, which is conducive to interlayer electron–hole recombination. Additionally, the BIEF direction is pointing from GeC to MXY. Similarly, the GeC/ZrS 2 , GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS HJs are all Z-scheme WSPCs. Moreover, the obtained U e ( U h ) values are 1.85 (1.73), 1.84 (2.08), 1.87 (1.99), 2.01 (2.74), 1.88 (1.69), 1.94 (2.06), and 1.97 (1.99) V, respectively (see Table 3). These values are also close to those reported for Z-scheme WSPCs (see Figure 6) [24,40,41,42,89,90,91,109]. This indicates that the GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS HJs could supply sufficient dynamic potentials to drive HERs and OERs under light irradiation.
At the initial stage of photocatalytic water-splitting, the light absorption capacity serves as another crucial factor. For highly efficient solar utilization, a wide and intense light absorption spectrum is typically required. Therefore, we investigated the optical absorption curves of GeC, MXY, and GeC/MXY HJs using the HSE06 method. The optical absorption coefficient can be calculated using the following formula [110]:
α ( ω ) = 2 ω c [ ϵ 1 2 ( ω ) + ϵ 2 2 ( ω ) ϵ 1 ( ω ) ] 1 / 2 ,
where ϵ 1 (or ϵ 2 ) represents the real (or imaginary) part of the dielectric function and ω is the frequency of light. As shown in Figure 7, GeC/MXY HJs possess strong visible light absorption ability and non-negligible near-infrared light absorption ability. In addition, GeC/MXY HJs exhibit higher absorption coefficients in the visible and near-infrared light regions, along with a redshift of the absorption spectra, compared to the corresponding MLs. Herein, the GeC/MXY HJs demonstrate excellent light absorption capacity. Furthermore, the proper band alignments and suitable directions of the heterointerface BIEF enable these GeC/MXY HJs to form a Z-scheme photocatalytic mechanism. This facilitates the HERs and OERs to occur in different sublayers and provides sufficient driving force to spontaneously achieve water redox reactions under illumination. Generally speaking, GeC/XYs HJs are promising candidates for direct Z-scheme WSPCs.

3. Computational Details

In this work, the GeC/MXY (M = Zr, Hf; X, Y = S, Se) HJs are constructed by stacking the 3 × 3 MXY supercell onto a 2 × 2 GeC supercell with an 18 Å vacuum layer to eliminate the image interaction between adjacent layers. Additionally, dipole correction is introduced along the z-direction [111]. All DFT calculations were carried out using VASP5.4 [112,113], and the electron–ion interactions were described using the projector-enhanced wave (PAW) method [114]. The generalized gradient approximation (GGA) [115] of Perdew–Burke–Ernzerhof (PBE) [116] was employed for the exchange correlation functional. Furthermore, Grimme’s DFT-D3 [117,118] method was employed to account for weak vdW interactions. The Monkhorst–Pack k-point grid for the first Brillouin zone was set to 13 × 13 × 1 (or 7 × 7 × 1) for MLs (or HJs). The energy cutoff was set to 500 eV, and all structures were sufficiently optimized with an energy (or force) tolerance of 10 5 eV (10 2 eV/Å). Given that GGA-PBE tends to underestimate the bandgaps [119], the Heyd–Scuseria–Ernzerhof functional (HSE06) was applied to accurately compute the electronic and optical properties [120]. The optical absorption spectra were computed based on the imaginary part of the dielectric functional, following the Kramers–Kronig dispersion relationship [110], and the band alignments of MLs and HJs were referenced to a common vacuum level.

4. Conclusions

In summary, the potential applications of GeC/MXY (M = Zr, Hf; X, Y = S, Se) HJs have been investigated through the calculation of their geometric, electronic, optical properties, band arrangement, and interface binding energies. Based on first-principles calculations, we analyzed their photocatalytic mechanism. All the considered GeC/MXY HJs, namely GeC/ZrS 2 , GeC/ZrSe 2 , GeC/SZrSe, GeC/SeZrS, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe, and GeC/SeHfS, were found to be direct Z-scheme photocatalytic systems with band edges spanning the water redox potentials. Charge redistribution at the heterointerface results in the formation of a BIEF pointing from GeC to MXY, enhancing the separation of the photoinduced carriers. Excitingly, the GeC/MXY HJs exhibit strong redox capacity for photocatalytic water decomposition, ensuring that the HER and OER processes occur spontaneously under light irradiation. Furthermore, the GeC/MXY HJs demonstrated strong visible light absorption and some near-infrared light absorption, guaranteeing efficient utilization of solar energy. These theoretical findings indicate that these GeC/MXY HJs are all promising WSPCs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29122793/s1, Figure S1: Top and side view of optimized geometries of GeC, ZrS 2 , ZrSe 2 , ZrSSe, HfS 2 , HfSe 2 and HfSSe, respectively; Figure S2: Band structures of GeC, ZrS 2 , ZrSe 2 , ZrSSe, HfS 2 , HfSe 2 and HfSSe, respectively; Figure S3: Electrostatic potential diagrams of GeC, HfS 2 , HfSe 2 , HfSSe, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe and GeC/SeHfS, respectively; Figure S4: Schematic diagrams of the photocatalytic mechanisms for GeC, HfS 2 , HfSe 2 , HfSSe, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe and GeC/SeHfS versus vacuum level; Figure S5: Schematic diagrams of the photocatalytic mechanisms for GeC, ZrS 2 , ZrSe 2 , ZrSSe, GeC/ZrS 2 , GeC/ZrSe 2 , GeC/SZrSe and GeC/SeZrS versus NHE; Figure S6: Schematic diagrams of the photocatalytic mechanisms for GeC, HfS 2 , HfSe 2 , HfSSe, GeC/HfS 2 , GeC/HfSe 2 , GeC/SHfSe and GeC/SeHfS versus NHE; Table S1: POSCAR file for the optimized GeC/ZrS 2 ; Table S2: POSCAR file for the optimized GeC/ZrSe 2 ; Table S3: POSCAR file for the optimized GeC/SZrSe; Table S4: POSCAR file for the optimized GeC/SeZrS; Table S5: POSCAR file for the optimized GeC/HfS 2 ; Table S6: POSCAR file for the optimized GeC/HfSe 2 ; Table S7: POSCAR file for the optimized GeC/SHfSe; Table S8: POSCAR file for the optimized GeC/SeHfS.

Author Contributions

G.W., Y.S.A. and H.Y. designed the project, guided the study, and prepared the manuscript; W.X., J.C. and Y.C. carried out the calculations; S.G., X.L. and L.Z. analyzed the calculated results and produced the illustrations. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China under Grant No. 12304295, the GHfund B under grant No. ghfund202302023082, the Science and Technology Research Program of Chongqing Municipal Education Commission under grant No. KJQN202201405, the China Postdoctoral Science Foundation under grant No. 2022MD723798, and the Special Funding for Postdoctoral Research Projects by Chongqing Municipal Human Resources and Social Security Bureau under grant No. 2022CQBSHTB3002.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, X.; Zhang, Z.; Wu, D.; Zhang, X.; Zhao, X.; Zhou, Z. Computational screening of 2D materials and rational design of heterojunctions for water splitting photocatalysts. Small Methods 2018, 2, 1700359. [Google Scholar] [CrossRef]
  3. Tachibana, Y.; Vayssieres, L.; Durrant, J.R. Artificial photosynthesis for solar water-splitting. Nat. Photonics 2012, 6, 511–518. [Google Scholar] [CrossRef]
  4. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef] [PubMed]
  5. Moniz, S.J.; Shevlin, S.A.; Martin, D.J.; Guo, Z.X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting—A critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  6. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [CrossRef]
  7. Wang, W.; Xu, X.; Zhou, W.; Shao, Z. Recent progress in metal-organic frameworks for applications in electrocatalytic and photocatalytic water splitting. Adv. Sci. 2017, 4, 1600371. [Google Scholar] [CrossRef]
  8. Tran, M.N.; Moreau, M.; Addad, A.; Teurtrie, A.; Roland, T.; De Waele, V.; Dewitte, M.; Thomas, L.; Levêque, G.; Dong, C.; et al. Boosting gas-phase TiO2 photocatalysis with weak electric field strengths of volt/centimeter. ACS Appl. Mater. Interfaces 2024, 16, 14852–14863. [Google Scholar] [CrossRef]
  9. Song, X.; Wei, G.; Sun, J.; Peng, C.; Yin, J.; Zhang, X.; Jiang, Y.; Fei, H. Overall photocatalytic water splitting by an organolead iodide crystalline material. Nat. Catal. 2020, 3, 1027–1033. [Google Scholar] [CrossRef]
  10. Mao, J.; Ta, Q.T.H.; Tri, N.N.; Shou, L.; Seo, S.; Xu, W. 2D MoTe2 nanomesh with a large surface area and uniform pores for highly active hydrogen evolution catalysis. Appl. Mater. Today 2023, 35, 101939. [Google Scholar] [CrossRef]
  11. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  12. Fu, C.F.; Wu, X.; Yang, J. Material design for photocatalytic water splitting from a theoretical perspective. Adv. Mater. 2018, 30, 1802106. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, G.Z.; Chang, J.L.; Tang, W.; Xie, W.; Ang, Y.S. 2D materials and heterostructures for photocatalytic water-splitting: A theoretical perspective. J. Phys. D Appl. Phys. 2022, 55, 293002. [Google Scholar] [CrossRef]
  14. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef]
  15. Faraji, M.; Yousefi, M.; Yousefzadeh, S.; Zirak, M.; Naseri, N.; Jeon, T.H.; Choi, W.; Moshfegh, A.Z. Two-dimensional materials in semiconductor photoelectrocatalytic systems for water splitting. Energy Environ. Sci. 2019, 12, 59–95. [Google Scholar] [CrossRef]
  16. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based heterostructured photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  17. Wang, G.; Tang, W.; Xie, W.; Tang, Q.; Wang, Y.; Guo, H.; Gao, P.; Dang, S.; Chang, J. Type-II CdS/PtSSe heterostructures used as highly efficient water-splitting photocatalysts. Appl. Surf. Sci. 2022, 589, 152931. [Google Scholar] [CrossRef]
  18. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar water splitting cells. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef] [PubMed]
  19. Jiang, C.; Moniz, S.J.; Wang, A.; Zhang, T.; Tang, J. Photoelectrochemical devices for solar water splitting–materials and challenges. Chem. Soc. Rev. 2017, 46, 4645–4660. [Google Scholar] [CrossRef]
  20. Zhou, P.; Yu, J.; Jaroniec, M. All-solid-state Z-scheme photocatalytic systems. Adv. Mater. 2014, 26, 4920–4935. [Google Scholar] [CrossRef]
  21. Li, H.; Tu, W.; Zhou, Y.; Zou, Z. Z-Scheme photocatalytic systems for promoting photocatalytic performance: Recent progress and future challenges. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, R.; Zhang, L.; Zheng, Q.; Gao, P.; Zhao, J.; Yang, J. Direct Z-scheme water splitting photocatalyst based on two-dimensional Van Der Waals heterostructures. J. Phys. Chem. Lett. 2018, 9, 5419–5424. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z.; Takata, T.; Nakabayashi, M.; Shibata, N.; et al. Scalable water splitting on particulate photocatalyst sheets with a solar-to-hydrogen energy conversion efficiency exceeding 1%. Nat. Mater. 2016, 15, 611–615. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, G.; Tang, W.; Xu, C.; He, J.; Zeng, Q.; Xie, W.; Gao, P.; Chang, J. Two-dimensional CdO/PtSSe heterojunctions used for Z-scheme photocatalytic water-splitting. Appl. Surf. Sci. 2022, 599, 153960. [Google Scholar] [CrossRef]
  25. Zuo, G.; Wang, Y.; Teo, W.L.; Xian, Q.; Zhao, Y. Direct Z-scheme TiO2–ZnIn2S4 nanoflowers for cocatalyst-free photocatalytic water splitting. Appl. Catal. B Environ. 2021, 291, 120126. [Google Scholar] [CrossRef]
  26. Wang, L.; Zheng, X.; Chen, L.; Xiong, Y.; Xu, H. Van der Waals heterostructures comprised of ultrathin polymer nanosheets for efficient Z-scheme overall water splitting. Angew. Chem. Int. Ed. 2018, 130, 3512–3516. [Google Scholar] [CrossRef]
  27. Zeng, C.; Hu, Y.; Zhang, T.; Dong, F.; Zhang, Y.; Huang, H. A core–satellite structured Z-scheme catalyst Cd0.5Zn0.5S/BiVO4 for highly efficient and stable photocatalytic water splitting. J. Mater. Chem. A 2018, 6, 16932–16942. [Google Scholar] [CrossRef]
  28. She, X.; Wu, J.; Xu, H.; Zhong, J.; Wang, Y.; Song, Y.; Nie, K.; Liu, Y.; Yang, Y.; Rodrigues, M.T.F.; et al. High efficiency photocatalytic water splitting using 2D α-Fe2O3/g-C3N4 Z-scheme catalysts. Adv. Energy Mater. 2017, 7, 1700025. [Google Scholar] [CrossRef]
  29. Zhu, M.; Sun, Z.; Fujitsuka, M.; Majima, T. Z-scheme photocatalytic water splitting on a 2D heterostructure of black phosphorus/bismuth vanadate using visible light. Angew. Chem. Int. Ed. 2018, 57, 2160–2164. [Google Scholar] [CrossRef]
  30. Yuan, Y.J.; Chen, D.; Yang, S.; Yang, L.X.; Wang, J.J.; Cao, D.; Tu, W.; Yu, Z.T.; Zou, Z.G. Constructing noble-metal-free Z-scheme photocatalytic overall water splitting systems using MoS2 nanosheet modified CdS as a H2 evolution photocatalyst. J. Mater. Chem. A 2017, 5, 21205–21213. [Google Scholar] [CrossRef]
  31. Li, L.; Guo, C.; Shen, J.; Ning, J.; Zhong, Y.; Hu, Y. Construction of sugar-gourd-shaped CdS/Co1−xS hollow hetero-nanostructure as an efficient Z-scheme photocatalyst for hydrogen generation. Chem. Eng. J. 2020, 400, 125925. [Google Scholar] [CrossRef]
  32. Wei, T.; Zhu, Y.N.; An, X.; Liu, L.M.; Cao, X.; Liu, H.; Qu, J. Defect modulation of Z-scheme TiO2/Cu2O photocatalysts for durable water splitting. ACS Catal. 2019, 9, 8346–8354. [Google Scholar] [CrossRef]
  33. Zhao, D.; Wang, Y.; Dong, C.L.; Huang, Y.C.; Chen, J.; Xue, F.; Shen, S.; Guo, L. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  34. Li, Z.; Hou, J.; Zhang, B.; Cao, S.; Wu, Y.; Gao, Z.; Nie, X.; Sun, L. Two-dimensional Janus heterostructures for superior Z-scheme photocatalytic water splitting. Nano Energy 2019, 59, 537–544. [Google Scholar] [CrossRef]
  35. 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. Nanotechnol. 2014, 9, 682–686. [Google Scholar] [CrossRef] [PubMed]
  36. Rivera, P.; Schaibley, J.R.; Jones, A.M.; Ross, J.S.; Wu, S.; Aivazian, G.; Klement, P.; Seyler, K.; Clark, G.; Ghimire, N.J.; et al. Observation of long-lived interlayer excitons in monolayer MoSe2–WSe2 heterostructures. Nat. Commun. 2015, 6, 6242. [Google Scholar] [CrossRef] [PubMed]
  37. Chiu, M.H.; Zhang, C.; Shiu, H.W.; Chuu, C.P.; Chen, C.H.; Chang, C.Y.S.; Chen, C.H.; Chou, M.Y.; Shih, C.K.; Li, L.J. Determination of band alignment in the single-layer MoS2/WSe2 heterojunction. Nat. Commun. 2015, 6, 7666. [Google Scholar] [CrossRef] [PubMed]
  38. Long, R.; Prezhdo, O.V. Quantum coherence facilitates efficient charge separation at a MoS2/MoSe2 van der Waals junction. Nano Lett. 2016, 16, 1996–2003. [Google Scholar] [CrossRef] [PubMed]
  39. Zhou, Z.; Niu, X.; Zhang, Y.; Wang, J. Janus MoSSe/WSeTe heterostructures: A direct Z-scheme photocatalyst for hydrogen evolution. J. Mater. Chem. A 2019, 7, 21835–21842. [Google Scholar] [CrossRef]
  40. Niu, X.; Bai, X.; Zhou, Z.; Wang, J. Rational design and characterization of direct Z-scheme photocatalyst for overall water splitting from excited state dynamics simulations. ACS Catal. 2020, 10, 1976–1983. [Google Scholar] [CrossRef]
  41. Jiang, X.; Gao, Q.; Xu, X.; Xu, G.; Li, D.; Cui, B.; Liu, D.; Qu, F. Design of a noble-metal-free direct Z-scheme photocatalyst for overall water splitting based on a SnC/SnSSe van der Waals heterostructure. Phys. Chem. Chem. Phys. 2021, 23, 21641–21651. [Google Scholar] [CrossRef] [PubMed]
  42. Fu, C.F.; Li, X.; Yang, J. A rationally designed two-dimensional MoSe2/Ti2CO2 heterojunction for photocatalytic overall water splitting: Simultaneously suppressing electron–hole recombination and photocorrosion. Chem. Sci. 2021, 12, 2863–2869. [Google Scholar] [CrossRef] [PubMed]
  43. Ji, Y.; Dong, H.; Hou, T.; Li, Y. Monolayer graphitic germanium carbide (g-GeC): The promising cathode catalyst for fuel cell and lithium–oxygen battery applications. J. Mater. Chem. A 2018, 6, 2212–2218. [Google Scholar] [CrossRef]
  44. Şahin, H.; Cahangirov, S.; Topsakal, M.; Bekaroglu, E.; Akturk, E.; Senger, R.T.; Ciraci, S. Monolayer honeycomb structures of group-IV elements and III-V binary compounds: First-principles calculations. Phys. Rev. B 2009, 80, 155453. [Google Scholar] [CrossRef]
  45. Pan, L.; Liu, H.; Wen, Y.; Tan, X.; Lv, H.; Shi, J.; Tang, X. First-principles study of monolayer and bilayer honeycomb structures of group-IV elements and their binary compounds. Phys. Lett. A 2011, 375, 614–619. [Google Scholar] [CrossRef]
  46. Hao, A.; Yang, X.; Wang, X.; Zhu, Y.; Liu, X.; Liu, R. First-principles investigations on electronic, elastic and optical properties of XC (X = Si, Ge, and Sn) under high pressure. J. Appl. Phys. 2010, 108, 063531. [Google Scholar] [CrossRef]
  47. Peng, Q.; Liang, C.; Ji, W.; De, S. A first-principles study of the mechanical properties of g-GeC. Mech. Mater. 2013, 64, 135–141. [Google Scholar] [CrossRef]
  48. Ren, K.; Sun, M.; Luo, Y.; Wang, S.; Xu, Y.; Yu, J.; Tang, W. Electronic and optical properties of van der Waals vertical heterostructures based on two-dimensional transition metal dichalcogenides: First-principles calculations. Phys. Lett. A 2019, 383, 1487–1492. [Google Scholar] [CrossRef]
  49. Wu, X.; Zhang, W.; Yan, L.; Luo, R. The deposition and optical properties of Ge1−xCx thin film and infrared multilayer antireflection coatings. Thin Solid Film. 2008, 516, 3189–3195. [Google Scholar] [CrossRef]
  50. Yuan, H.; Williams, R.S. Synthesis by laser ablation and characterization of pure germanium-carbon alloy thin films. Chem. Mater. 1993, 5, 479–485. [Google Scholar] [CrossRef]
  51. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-like two-dimensional materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef] [PubMed]
  52. Li, Z.; Lv, Y.; Ren, L.; Li, J.; Kong, L.; Zeng, Y.; Tao, Q.; Wu, R.; Ma, H.; Zhao, B.; et al. Efficient strain modulation of 2D materials via polymer encapsulation. Nat. Commun. 2020, 11, 1151. [Google Scholar] [CrossRef] [PubMed]
  53. Li, J.; Chen, M.; Zhang, C.; Dong, H.; Lin, W.; Zhuang, P.; Wen, Y.; Tian, B.; Cai, W.; Zhang, X. Fractal-theory-based control of the shape and quality of CVD-grown 2D materials. Adv. Mater. 2019, 31, 1902431. [Google Scholar] [CrossRef] [PubMed]
  54. Din, H.; Idrees, M.; Albar, A.; Shafiq, M.; Ahmad, I.; Nguyen, C.V.; Amin, B. Rashba spin splitting and photocatalytic properties of GeC-MSSe (M= Mo, W) van der Waals heterostructures. Phys. Rev. B 2019, 100, 165425. [Google Scholar] [CrossRef]
  55. Jiang, X.; Xie, W.; Xu, X.; Gao, Q.; Li, D.; Cui, B.; Liu, D.; Qu, F. A bifunctional GeC/SnSSe heterostructure for highly efficient photocatalysts and photovoltaic devices. Nanoscale 2022, 14, 7292–7302. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, G.; Zhang, L.; Li, Y.; Zhao, W.; Kuang, A.; Li, Y.; Xia, L.; Li, Y.; Xiao, S. Biaxial strain tunable photocatalytic properties of 2D ZnO/GeC heterostructure. J. Phys. D Appl. Phys. 2020, 53, 015104. [Google Scholar] [CrossRef]
  57. Gao, X.; Shen, Y.; Ma, Y.; Wu, S.; Zhou, Z. ZnO/g-GeC van der Waals heterostructure: Novel photocatalyst for small molecule splitting. J. Mater. Chem. C 2019, 7, 4791–4799. [Google Scholar] [CrossRef]
  58. Cao, M.; Luan, L.; Wang, Z.; Zhang, Y.; Yang, Y.; Liu, J.; Tian, Y.; Wei, X.; Fan, J.; Xie, Y.; et al. Type-II GeC/ZnTe heterostructure with high-efficiency of photoelectrochemical water splitting. Appl. Phys. Lett. 2021, 119, 083101. [Google Scholar] [CrossRef]
  59. Huong, P.T.; Idrees, M.; Amin, B.; Hieu, N.N.; Phuc, H.V.; Hoa, L.T.; Nguyen, C.V. Electronic structure, optoelectronic properties and enhanced photocatalytic response of GaN–GeC van der Waals heterostructures: A first principles study. RSC Adv. 2020, 10, 24127–24133. [Google Scholar] [CrossRef]
  60. Yang, Z.; Wang, J.; Hu, G.; Yuan, X.; Ren, J.; Zhao, X. Strain-tunable Zeeman splitting and optical properties of CrBr3/GeC van der Waals heterostructure. Results Phys. 2022, 37, 105559. [Google Scholar] [CrossRef]
  61. Lou, P.; Lee, J.Y. GeC/GaN vdW heterojunctions: A promising photocatalyst for overall water splitting and solar energy conversion. ACS Appl. Mater. Interf. 2020, 12, 14289–14297. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, Y.L.; Shi, Y.; Yang, C.L. Two-dimensional MoSSe/g-GeC van der waals heterostructure as promising multifunctional system for solar energy conversion. Appl. Surf. Sci. 2021, 545, 148952. [Google Scholar] [CrossRef]
  63. Gao, X.; Shen, Y.; Liu, J.; Lv, L.; Zhou, M.; Zhou, Z.; Feng, Y.P.; Shen, L. Boost the large driving photovoltages for overall water splitting in direct Z-scheme heterojunctions by interfacial polarization. Catal. Sci. Technol. 2022, 12, 3614–3621. [Google Scholar] [CrossRef]
  64. Abdulsalam, M.; Joubert, D.P. Optical spectrum and excitons in bulk and monolayer MX2 (M = Zr, Hf; X = S, Se). Phys. Status Solidi B 2016, 253, 705–711. [Google Scholar] [CrossRef]
  65. Abdulsalam, M.; Rugut, E.; Joubert, D. Mechanical, thermal and thermoelectric properties of MX2 (M = Zr, Hf; X = S, Se). Mater. Today Commun. 2020, 25, 101434. [Google Scholar] [CrossRef]
  66. Bera, J.; Betal, A.; Sahu, S. Spin orbit coupling induced enhancement of thermoelectric performance of HfX2 (X = S, Se) and its Janus monolayer. J. Alloys Compd. 2021, 872, 159704. [Google Scholar] [CrossRef]
  67. Dimple; Jena, N.; Rawat, A.; Ahammed, R.; Mohanta, M.K.; De Sarkar, A. Emergence of high piezoelectricity along with robust electron mobility in Janus structures in semiconducting Group IVB dichalcogenide monolayers. J. Mater. Chem. A 2018, 6, 24885–24898. [Google Scholar] [CrossRef]
  68. Shi, W.; Wang, Z. Mechanical and electronic properties of Janus monolayer transition metal dichalcogenides. J. Phys. Condens. Mat. 2018, 30, 215301. [Google Scholar] [CrossRef] [PubMed]
  69. Som, N.N.; Jha, P.K. Hydrogen evolution reaction of metal di-chalcogenides: ZrS2, ZrSe2 and Janus ZrSSe. Int. J. Hydrogen Energy 2020, 45, 23920–23927. [Google Scholar] [CrossRef]
  70. Hoat, D.; Naseri, M.; Hieu, N.N.; Ponce-Pérez, R.; Rivas-Silva, J.; Vu, T.V.; Cocoletzi, G.H. A comprehensive investigation on electronic structure, optical and thermoelectric properties of the HfSSe Janus monolayer. J. Phys. Chem. Solids 2020, 144, 109490. [Google Scholar] [CrossRef]
  71. Kaur, H.; Yadav, S.; Srivastava, A.K.; Singh, N.; Rath, S.; Schneider, J.J.; Sinha, O.P.; Srivastava, R. High-yield synthesis and liquid-exfoliation of two-dimensional belt-like hafnium disulphide. Nano Res. 2018, 11, 343–353. [Google Scholar] [CrossRef]
  72. Yue, R.; Barton, A.T.; Zhu, H.; Azcatl, A.; Pena, L.F.; Wang, J.; Peng, X.; Lu, N.; Cheng, L.; Addou, R.; et al. HfSe2 thin films: 2D transition metal dichalcogenides grown by molecular beam epitaxy. ACS Nano 2015, 9, 474–480. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, M.; Zhu, Y.; Wang, X.; Feng, Q.; Qiao, S.; Wen, W.; Chen, Y.; Cui, M.; Zhang, J.; Cai, C.; et al. Controlled synthesis of ZrS2 monolayer and few layers on hexagonal boron nitride. J. Am. Chem. Soc. 2015, 137, 7051–7054. [Google Scholar] [CrossRef] [PubMed]
  74. Tsipas, P.; Tsoutsou, D.; Marquez-Velasco, J.; Aretouli, K.; Xenogiannopoulou, E.; Vassalou, E.; Kordas, G.; Dimoulas, A. Epitaxial ZrSe2/MoSe2 semiconductor vd Waals heterostructures on wide band gap AlN substrates. Microelectron. Eng. 2015, 147, 269–272. [Google Scholar] [CrossRef]
  75. Zeng, Z.; Yin, Z.; Huang, X.; Li, H.; He, Q.; Lu, G.; Boey, F.; Zhang, H. Single-layer semiconducting nanosheets: High-yield preparation and device fabrication. Ang. Chem. Int. Ed. 2011, 123, 11289–11293. [Google Scholar] [CrossRef]
  76. Zhang, J.; Jia, S.; Kholmanov, I.; Dong, L.; Er, D.; Chen, W.; Guo, H.; Jin, Z.; Shenoy, V.B.; Shi, L.; et al. Janus monolayer transition-metal dichalcogenides. ACS Nano 2017, 11, 8192–8198. [Google Scholar] [CrossRef] [PubMed]
  77. Lu, A.Y.; Zhu, H.; Xiao, J.; Chuu, C.P.; Han, Y.; Chiu, M.H.; Cheng, C.C.; Yang, C.W.; Wei, K.H.; Yang, Y.; et al. Janus monolayers of transition metal dichalcogenides. Nat. Nanotechnol. 2017, 12, 744–749. [Google Scholar] [CrossRef]
  78. Lin, Y.C.; Liu, C.; Yu, Y.; Zarkadoula, E.; Yoon, M.; Puretzky, A.A.; Liang, L.; Kong, X.; Gu, Y.; Strasser, A.; et al. Low energy implantation into transition-metal dichalcogenide monolayers to form Janus structures. ACS Nano 2020, 14, 3896–3906. [Google Scholar] [CrossRef] [PubMed]
  79. Singh, A.; Jain, M.; Bhattacharya, S. MoS2 and Janus (MoSSe) based 2D van der Waals heterostructures: Emerging direct Z-scheme photocatalysts. Nanoscale Adv. 2021, 3, 2837–2845. [Google Scholar] [CrossRef]
  80. Zhu, X.T.; Xu, Y.; Cao, Y.; Zou, D.F.; Sheng, W. Direct Z-scheme arsenene/HfS2 van der Waals heterojunction for overall photocatalytic water splitting: First-principles study. Appl. Surf. Sci. 2022, 574, 151650. [Google Scholar] [CrossRef]
  81. Bai, Y.; Zhang, H.; Wu, X.; Xu, N.; Zhang, Q. Two-dimensional arsenene/ZrS2 (HfS2) deterostructures as direct Z-Scheme photocatalysts for overall water splitting. J. Phys. Chem. C 2022, 126, 2587–2595. [Google Scholar] [CrossRef]
  82. Sun, R.; Yang, C.L.; Wang, M.S.; Ma, X.G. High solar-to-hydrogen efficiency photocatalytic hydrogen evolution reaction with the HfSe2/InSe heterostructure. J. Power Sources 2022, 547, 232008. [Google Scholar] [CrossRef]
  83. Zhang, X.; Meng, Z.; Rao, D.; Wang, Y.; Shi, Q.; Liu, Y.; Wu, H.; Deng, K.; Liu, H.; Lu, R. Efficient band structure tuning, charge separation, and visible-light response in ZrS2-based van der Waals heterostructures. Energy Environ. Sci. 2016, 9, 841–849. [Google Scholar] [CrossRef]
  84. Cao, J.; Zhang, X.; Zhao, S.; Wang, S.; Cui, J. Mechanism of photocatalytic water splitting of 2D WSeTe/XS2 (X = Hf, Sn, Zr) van der Waals heterojunctions under the interaction of vertical intrinsic electric and built-in electric field. Appl. Surf. Sci. 2022, 599, 154012. [Google Scholar] [CrossRef]
  85. Opoku, F.; Akoto, O.; Oppong, S.O.B.; Adimado, A.A. Two-dimensional layered type-II MS2/BiOCl (M = Zr, Hf) van der Waals heterostructures: Promising photocatalysts for hydrogen generation. New J. Chem. 2021, 45, 20365–20373. [Google Scholar] [CrossRef]
  86. Ahmad, S.; Idrees, M.; Khan, F.; Nguyen, C.; Ahmad, I.; Amin, B. Strain engineering of Janus ZrSSe and HfSSe monolayers and ZrSSe/HfSSe van der Waals heterostructure. Chem. Phys. Lett. 2021, 776, 138689. [Google Scholar] [CrossRef]
  87. Fu, C.F.; Luo, Q.; Li, X.; Yang, J. Two-dimensional van der Waals nanocomposites as Z-scheme type photocatalysts for hydrogen production from overall water splitting. J. Mater. Chem. A 2016, 4, 18892–18898. [Google Scholar] [CrossRef]
  88. Gao, Y.; Fu, C.; Hu, W.; Yang, J. Designing direct Z-scheme heterojunctions enabled by edge-modified phosphorene nanoribbons for photocatalytic overall water splitting. J. Phys. Chem. Lett. 2021, 13, 1–11. [Google Scholar] [CrossRef] [PubMed]
  89. Meng, J.; Wang, J.; Wang, J.; Li, Q.; Yang, J. β-SnS/GaSe heterostructure: A promising solar-driven photocatalyst with low carrier recombination for overall water splitting. J. Mater. Chem. A 2022, 10, 3443–3453. [Google Scholar] [CrossRef]
  90. Xiong, R.; Shu, Y.; Yang, X.; Zhang, Y.; Wen, C.; Anpo, M.; Wu, B.; Sa, B. Direct Z-scheme WTe2/InSe van der Waals heterostructure for overall water splitting. Catal. Sci. Technol. 2022, 12, 3272–3280. [Google Scholar] [CrossRef]
  91. Dai, Z.N.; Cao, Y.; Yin, W.J.; Sheng, W.; Xu, Y. Z-scheme SnC/HfS2 van der Waals heterojunction increases photocatalytic overall water splitting. J. Phys. D Appl. Phys. 2022, 55, 315503. [Google Scholar] [CrossRef]
  92. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  93. Wang, G.; Chang, J.; Guo, S.D.; Wu, W.; Tang, W.; Guo, H.; Dang, S.; Wang, R.; Ang, Y.S. MoSSe/Hf(Zr)S2 heterostructures used for efficient Z-scheme photocatalytic water-splitting. Phys. Chem. Chem. Phys. 2022, 24, 25287–25297. [Google Scholar] [CrossRef] [PubMed]
  94. Novoselov, K.; Mishchenko, A.; Carvalho, A.; Castro Neto, A. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439. [Google Scholar] [CrossRef]
  95. Björkman, T.; Gulans, A.; Krasheninnikov, A.V.; Nieminen, R.M. van der Waals bonding in layered compounds from advanced density-functional first-principles calculations. Phys. Rev. Lett. 2012, 108, 235502. [Google Scholar] [CrossRef]
  96. Guo, H.; Zhang, Z.; Huang, B.; Wang, X.; Niu, H.; Guo, Y.; Li, B.; Zheng, R.; Wu, H. Theoretical study on the photocatalytic properties of 2D InX (X = S, Se)/transition metal disulfide (MoS2 and WS2) van der Waals heterostructures. Nanoscale 2020, 12, 20025–20032. [Google Scholar] [CrossRef] [PubMed]
  97. Xu, L.; Huang, W.Q.; Hu, W.; Yang, K.; Zhou, B.X.; Pan, A.; Huang, G.F. Two-dimensional MoS2-graphene-based multilayer van der Waals heterostructures: Enhanced charge transfer and optical absorption, and electric-field tunable Dirac point and band gap. Chem. Mater. 2017, 29, 5504–5512. [Google Scholar] [CrossRef]
  98. Bafekry, A.; Obeid, M.; Nguyen, C.V.; Ghergherehchi, M.; Tagani, M.B. Graphene hetero-multilayer on layered platinum mineral jacutingaite (Pt2HgSe3): Van der Waals heterostructures with novel optoelectronic and thermoelectric performances. J. Mater. Chem. A 2020, 8, 13248–13260. [Google Scholar] [CrossRef]
  99. Zhang, C.F.; Yang, C.L.; Wang, M.S.; Ma, X.G. Z-Scheme photocatalytic solar-energy-to-hydrogen conversion driven by the HfS2/SiSe heterostructure. J. Mater. Chem. C 2022, 10, 5474–5481. [Google Scholar] [CrossRef]
  100. Wang, G.; Li, Z.; Wu, W.; Guo, H.; Chen, C.; Yuan, H.; Yang, S.A. A two-dimensional h-BN/C2N heterostructure as a promising metal-free photocatalyst for overall water-splitting. Phys. Chem. Chem. Phys. 2020, 22, 24446–24454. [Google Scholar] [CrossRef]
  101. Zhou, Z.; Yuan, S.; Wang, J. Theoretical progress on direct Z-scheme photocatalysis of two-dimensional heterostructures. Front. Phys. 2021, 16, 43203. [Google Scholar] [CrossRef]
  102. He, C.; Zhang, J.; Zhang, W.; Li, T. Type-II InSe/g-C3N4 heterostructure as a high-efficiency oxygen evolution reaction catalyst for photoelectrochemical water splitting. J. Phys. Chem. Lett. 2019, 10, 3122–3128. [Google Scholar] [CrossRef] [PubMed]
  103. Xu, L.; Huang, W.Q.; Wang, L.L.; Tian, Z.A.; Hu, W.; Ma, Y.; Wang, X.; Pan, A.; Huang, G.F. Insights into enhanced visible-light photocatalytic hydrogen evolution of g-C3N4 and highly reduced graphene oxide composite: The role of oxygen. Chem. Mater. 2015, 27, 1612–1621. [Google Scholar] [CrossRef]
  104. Liu, J.; Cheng, B.; Yu, J. A new understanding of the photocatalytic mechanism of the direct Z-scheme g-C3N4/TiO2 heterostructure. Phys. Chem. Chem. Phys. 2016, 18, 31175–31183. [Google Scholar] [CrossRef] [PubMed]
  105. Bai, S.; Jiang, J.; Zhang, Q.; Xiong, Y. Steering charge kinetics in photocatalysis: Intersection of materials syntheses, characterization techniques and theoretical simulations. Chem. Phys. Rev. 2015, 44, 2893–2939. [Google Scholar]
  106. Fu, C.F.; Wu, X.; Yang, J. Theoretical design of two-dimensional visible light-driven photocatalysts for overall water splitting. Chem. Phys. Rev. 2022, 3, 011310. [Google Scholar] [CrossRef]
  107. Huang, Z.F.; Song, J.; Wang, X.; Pan, L.; Li, K.; Zhang, X.; Wang, L.; Zou, J.J. Switching charge transfer of C3N4/W18O49 from type-II to Z-scheme by interfacial band bending for highly efficient photocatalytic hydrogen evolution. Nano Energy 2017, 40, 308–316. [Google Scholar] [CrossRef]
  108. Zhang, Z.; Yates, J.T., Jr. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chem. Rev. 2012, 112, 5520–5551. [Google Scholar] [CrossRef] [PubMed]
  109. Fan, Y.; Wang, J.; Zhao, M. Spontaneous full photocatalytic water splitting on 2D MoSe2/SnSe2 and WSe2/SnSe2 vdW heterostructures. Nanoscale 2019, 11, 14836–14843. [Google Scholar] [CrossRef]
  110. Tian, F.; Liu, C. DFT description on electronic structure and optical absorption properties of anionic S-doped anatase TiO2. J. Phys. Chem. B 2006, 110, 17866–17871. [Google Scholar] [CrossRef]
  111. Bengtsson, L. Dipole correction for surface supercell calculations. Phys. Rev. B 1999, 59, 12301. [Google Scholar] [CrossRef]
  112. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef] [PubMed]
  113. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  114. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef] [PubMed]
  115. Singh, D.; Ashkenazi, J. Magnetism with generalized-gradient-approximation density functionals. Phys. Rev. B 1992, 46, 11570. [Google Scholar] [CrossRef] [PubMed]
  116. Ernzerhof, M.; Scuseria, G.E. Assessment of the Perdew–Burke–Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. [Google Scholar] [CrossRef]
  117. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  118. Moellmann, J.; Grimme, S. DFT-D3 study of some molecular crystals. J. Phys. Chem. C 2014, 118, 7615–7621. [Google Scholar] [CrossRef]
  119. Kümmel, S.; Kronik, L. Orbital-dependent density functionals: Theory and applications. Rev. Mod. Phys. 2008, 80, 3. [Google Scholar] [CrossRef]
  120. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
Figure 1. Top and side views for various GeC/MXY HJs.
Figure 1. Top and side views for various GeC/MXY HJs.
Molecules 29 02793 g001
Figure 2. Band structures for various GeC/MXY HJs. The orange (or green) dots denote the contribution from the GeC (or MXY) layer.
Figure 2. Band structures for various GeC/MXY HJs. The orange (or green) dots denote the contribution from the GeC (or MXY) layer.
Molecules 29 02793 g002
Figure 3. Electrostatic potential diagrams of (a) GeC, (b) ZrS 2 , (c) ZrSe 2 , (d) ZrSSe, (e) GeC/ZrS 2 , (f) GeC/ZrSe 2 , (g) GeC/SZrSe, and (h) GeC/SeZrS, respectively.
Figure 3. Electrostatic potential diagrams of (a) GeC, (b) ZrS 2 , (c) ZrSe 2 , (d) ZrSSe, (e) GeC/ZrS 2 , (f) GeC/ZrSe 2 , (g) GeC/SZrSe, and (h) GeC/SeZrS, respectively.
Molecules 29 02793 g003
Figure 4. VCDDs and PACDDs for various GeC/MXY HJs.
Figure 4. VCDDs and PACDDs for various GeC/MXY HJs.
Molecules 29 02793 g004
Figure 5. Band alignments for (a) GeC, (b) ZrS 2 , (c) ZrSe 2 , (d) ZrSSe, (e) GeC/ZrS 2 , (f) GeC/ZrSe 2 , (g) GeC/SZrSe, and (h) GeC/SeZrS versus vacuum level.
Figure 5. Band alignments for (a) GeC, (b) ZrS 2 , (c) ZrSe 2 , (d) ZrSSe, (e) GeC/ZrS 2 , (f) GeC/ZrSe 2 , (g) GeC/SZrSe, and (h) GeC/SeZrS versus vacuum level.
Molecules 29 02793 g005
Figure 6. Comparison of U e and U h values of the proposed GeC/MXY HJs with some other reported HJs [24,40,41,42,89,90,91,109].
Figure 6. Comparison of U e and U h values of the proposed GeC/MXY HJs with some other reported HJs [24,40,41,42,89,90,91,109].
Molecules 29 02793 g006
Figure 7. (ah) Optical absorption curves of various GeC/MXY HJs compared with those of GeC and MXY MLs.
Figure 7. (ah) Optical absorption curves of various GeC/MXY HJs compared with those of GeC and MXY MLs.
Molecules 29 02793 g007
Table 1. Lattice constants (a), bond lengths ( L B ), bandgaps ( E g ), dipole moments ( μ ), and EPDs ( Δ E ) between two sides for GeC and MXY MLs.
Table 1. Lattice constants (a), bond lengths ( L B ), bandgaps ( E g ), dipole moments ( μ ), and EPDs ( Δ E ) between two sides for GeC and MXY MLs.
Systemsa (Å)a (Å) (Refs.) L B  (Å) L B  (Å) (Refs.) E g (eV) E g (eV) (Refs.) μ (D) Δ E (eV)
GeC3.2353.26 [54,55], 3.233 [56]1.8681.882 [54], 1.887 [56]2.873.01 [54], 2.90 [55]00
3.263 [57] 2.88 [56], 2.782 [57]
2.85 [58]
ZrS 2 3.6853.70 [64,65], 3.69 [67]2.5742.58 [67], 2.570 [93]2.021.99 [67], 1.96 [93]00
3.669 [93]
ZrSe 2 3.8003.82 [64,65], 3.75 [67]2.7062.69 [67], 2.702 [93]1.191.07 [67], 1.14 [93]00
3.786 [93]
ZrSSe3.7433.73 [67]2.568 (2.713)2.55 (2.72) [67]1.461.37 [67]0.0430.135
HfS 2 3.6453.66 [64,65], 3.65 [66]2.5522.55 [66], 2.56 [67]2.132.03 [67], 2.07 [93]00
3.65 [67], 3.628 [93] 2.548 [93]
HfSe 2 3.7683.82 [64], 3.78 [65,66]2.6852.69 [66], 2.68 [67]1.331.16 [67], 1.26 [93]00
3.72 [67], 3.751 [93] 2.681 [93]
HfSSe3.7053.71 [66], 3.68 [67]2.550 (2.687)2.55 (2.69) [66], 2.54 (2.69) [67]1.561.45 [67]0.0350.110
Table 2. Lattice parameters (a), interlayer distances ( d i ), interface formation energies ( E f ), dipole moments ( μ ), EPDs ( Δ E ) between two surfaces, and charge transferred from GeC layer ( Δ Q ) in various GeC/MXY HJs.
Table 2. Lattice parameters (a), interlayer distances ( d i ), interface formation energies ( E f ), dipole moments ( μ ), EPDs ( Δ E ) between two surfaces, and charge transferred from GeC layer ( Δ Q ) in various GeC/MXY HJs.
Systemsa (Å) d i  (Å) E f  (meV/Å 2 ) μ (D) Δ E (eV) Δ Q (e)
GeC/ZrS 2 6.4263.367−18.50.160.160.11
GeC/ZrSe 2 6.5273.446−28.10.150.150.09
GeC/SZrSe6.4773.376−28.80.240.250.11
GeC/SeZrS6.4773.436−29.40.040.050.09
GeC/HfS 2 6.3923.468−25.20.080.090.08
GeC/HfSe 2 6.4993.484−28.50.100.100.07
GeC/SZrSe6.4443.421−27.80.190.200.08
GeC/SeZrS6.4443.519−28.50.010.020.07
Table 3. The bandgap ( E g ), CBO, VBO, U e , and U h values for GeC/MXY HJs.
Table 3. The bandgap ( E g ), CBO, VBO, U e , and U h values for GeC/MXY HJs.
Systems E g  (eV)CBO (eV)VBO (eV) U e  (V) U h  (V)
GeC/ZrS 2 0.452.591.702.142.75
GeC/ZrSe 2 0.452.300.681.851.73
GeC/SZrSe0.552.190.921.842.08
GeC/SeZrS0.432.321.071.871.99
GeC/HfS 2 0.532.341.792.012.74
GeC/HfSe 2 0.592.180.711.881.69
GeC/SHfSe0.662.150.991.942.06
GeC/SeHfS0.542.281.121.971.99
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, G.; Xie, W.; Guo, S.; Chang, J.; Chen, Y.; Long, X.; Zhou, L.; Ang, Y.S.; Yuan, H. Two-Dimensional GeC/MXY (M = Zr, Hf; X, Y = S, Se) Heterojunctions Used as Highly Efficient Overall Water-Splitting Photocatalysts. Molecules 2024, 29, 2793. https://doi.org/10.3390/molecules29122793

AMA Style

Wang G, Xie W, Guo S, Chang J, Chen Y, Long X, Zhou L, Ang YS, Yuan H. Two-Dimensional GeC/MXY (M = Zr, Hf; X, Y = S, Se) Heterojunctions Used as Highly Efficient Overall Water-Splitting Photocatalysts. Molecules. 2024; 29(12):2793. https://doi.org/10.3390/molecules29122793

Chicago/Turabian Style

Wang, Guangzhao, Wenjie Xie, Sandong Guo, Junli Chang, Ying Chen, Xiaojiang Long, Liujiang Zhou, Yee Sin Ang, and Hongkuan Yuan. 2024. "Two-Dimensional GeC/MXY (M = Zr, Hf; X, Y = S, Se) Heterojunctions Used as Highly Efficient Overall Water-Splitting Photocatalysts" Molecules 29, no. 12: 2793. https://doi.org/10.3390/molecules29122793

Article Metrics

Back to TopTop