Pham D. Trung*a and
Hien D. Tong
b
aYersin University, 27 Ton That Tung, Ward 8, Dalat City, Lam Dong Province, Vietnam. E-mail: phdtrung2018@gmail.com
bFaculty of Engineering, Vietnamese-German University (VGU), Ben Cat City, Binh Duong Province, Vietnam
First published on 17th March 2025
Recently, the quaternary Janus monolayers with the formula A2B2X3Y3 have been shown to be promising candidates for optoelectronic applications, especially in the photocatalytic water splitting reaction. Therefore, first-principles calculations were employed to investigate the photocatalytic properties of Ga2Ge2X3Y3 (X and Y represent S, Se or Te atoms) monolayers. The Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers exhibit dynamic and thermal stability, supported by high cohesive energies (3.78–4.20 eV) and positive phonon dispersion. With a moderate Young's modulus (50.02–65.31 N m−1) and high Poisson's ratio (0.39–0.41), these monolayers offer a balance of stiffness and flexibility, making them suitable for flexible electronic applications. Especially, the difference in work function of 0.27 eV induces an intrinsic electric field in the Ga2Ge2S3Se3 monolayer, making the electronic structure of this material be suitable for the photocatalytic water splitting process. With light irradiation, the oxygen evolution reaction (OER) happened simultaneously, producing electrons and H+ protons for the hydrogen evolution reaction (HER) to happen at a low potential barrier. Moreover, the Ga2Ge2S3Se3 monolayer has a high absorption rate α(ω) of 105–106 cm−1 and a high electron mobility of 430.82–461.50 cm2 V−1 s−1. These characteristics result in a good solar-to-hydrogen of the Ga2Ge2S3Se3 monolayer (14.80%) which is promising for use in photon-driven water splitting.
To further enhance the efficiency of photocatalytic water splitting, researchers are exploring various strategies. Bandgap engineering, which involves modifying the bandgap of materials to better match the solar spectrum, is a key approach. Additionally, strategies to prevent electron–hole recombination, such as introducing co-catalysts or forming heterojunctions, are being investigated.3,10 Surface modification techniques are employed to improve light absorption and charge transfer, while nanostructuring is used to increase surface area and light harvesting efficiency.4,5 These advancements hold the potential to significantly improve the performance of photocatalytic water splitting systems.
Among 2D materials, ternary 2D chalcogenides have emerged as promising materials for photocatalytic water splitting due to their unique electronic and optical properties.11 Compounds like ZnIn2S4 and AgInS2 have garnered significant attention.12 These materials often exhibit suitable bandgap energies for visible light absorption, efficient charge carrier separation, and enhanced stability compared to their binary counterparts.13 However, challenges such as photocorrosion and low carrier mobility still need to be addressed. Ongoing research focuses on optimizing synthesis techniques, incorporating co-catalysts, and developing heterostructures to further improve the photocatalytic performance of these materials.
Two-dimensional (2D) layered transition metal phosphorus trichalcogenides (MPX3, where M represents transition metals and X = S, Se) exhibit high in-plane stiffness and lower cleavage energies than graphite.14–17 This allows them to be exfoliated down to atomic thickness. Recently, these materials have attracted renewed attention due to their unique structural and electronic properties. Notably, chemically exfoliated monolayer FePS3 has demonstrated efficient photocatalytic hydrogen evolution.18 Inspired by these findings, monolayers such as AsGeC3, SbSnC3, and BiPbC3 have been investigated, revealing strong stability and excellent optical properties.19 Additionally, CoGeSe3, GaSnS3, and InSnS3 monolayers show promise as visible-light photocatalysts with high carrier mobility.20,21 The GaGeSe3 monolayer, in particular, exhibits a solar-to-hydrogen conversion efficiency of 11.33% and a charge carrier mobility of 790.65 cm2 V−1 s−1.22 Substitution is an effective strategy for tuning the electronic properties of monolayers, as demonstrated experimentally. For instance, the MoSSe monolayer was synthesized by replacing the top-layer S atoms with Se atoms.23 Building on this concept, many quaternary monolayers have been theoretically designed based on the aforementioned ternary structures. The Zn2P2S3Se3 and Cd2P2S3Se3 monolayers are predicted to exhibit exceptionally high charge carrier mobility of 2772 cm2 V−1 s−1.24 Meanwhile, Janus-type monolayers such as In2Ge2Te3Se3, In2Ge2Se6, and In2Ge2S3Se3 have emerged as promising candidates for thermoelectric applications in high-temperature environments. In this study, the Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers were constructed based on the atomic structures of GaGeS3 and GaGeSe3 monolayers, respectively. The structural, electronic and transport properties of the two new monolayers are studied using first-principles calculation.
![]() | ||
Fig. 1 Optimized atomic structure of Ga2Ge2X3Y3 (X/Y = S, Se, Te; X ≠ Y) monolayers shown in (a) top view and (b) site view together with (c) their phonon dispersions. |
It is worth noting that the Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers are formed by modifying the atomic structures of GaGeS3 and GaGeSe3 monolayers,22 respectively, where heavier atoms Se/Te are substituted for lighter host atoms S/Se. Such substitutions cause lattice enlargement, as shown in Table 1, the lattice constants (a = b) of Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers are 6.29 Å and 6.74 Å, respectively. These lattice constants are about 0.17–0.28 Å longer than the corresponding values of GaGeS3 or GaGeSe3 monolayer.22 Due to the stacking of sublayer of identical elements, the substitution of Se atoms for S atoms replaces the whole S-sublayer in the GaGeS3 monolayer22 (S–Ge–Ga–Ge–S) with the Se-sublayer in the Ga2Ge2S3Se3 monolayer (S–Ge–Ga–Ge–Se). Such a significant change causes noticeable enlargement of all bond-lengths as GaGeS3 monolayer22 transforms into Ga2Ge2S3Se3 monolayer. As shown in Table 1, the Ge–S, Ga–S, and Ge–Ge bond lengths increase by 0.3–0.13 Å. Consequently, the monolayer thickness, which is defined as the distance between the two edge sublayers, also increases by about 0.08 Å. Similarly, the data listed in Table 1 shows that the replacement of Se-sublayer in GaGeSe3 monolayer22 with the Te-sublayer in Ga2Ge2Se3Te3 monolayer results in some increase in the Ge–Se, Ga–Se, and Ge–Ge bond lengths by about 0.04–0.22 Å. Meanwhile, the thickness of Ga2Ge2Se3Te3 monolayer is 0.12 eV.
a (Å) | h (Å) | dGe–X (Å) | dGa–X (Å) | dGe–Ge (Å) | dGe–Y (Å) | dGa–Y (Å) | Ecoh (eV per atom) | EPBEg (eV) | EHSE06g (eV) | C11 (N m−1) | C12 (N m−1) | C66 (N m−1) | Y2D (N m−1) | ν2D | |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Ga2Ge2S3Se3 | 6.29 | 3.28 | 2.37 | 2.66 | 2.37 | 4.20 | 1.02 | 1.83 | 76.69 | 29.55 | 23.57 | 65.31 | 0.39 | ||
Ga2Ge2Se3Te3 | 6.74 | 3.50 | 2.42 | 2.57 | 2.87 | 3.78 | 0.06 | 0.63 | 59.82 | 24.21 | 17.81 | 50.02 | 0.41 |
The significant change in the structural characteristics of the two new monolayers requires further study on their stability. From an energy perspective, a new system is considered stable if its total energy is less than the combined energies of its individual components. Such energetic difference is called cohesive energy Ecoh, which can be calculated for the Ga2Ge2X3Y3 monolayers as follows:
![]() | (1) |
The mechanical stability, stiffness and deformation behavior of 2D materials, such as graphene, transition metal dichalcogenides, and MXenes, etc.,33–35 can be analyzed using Born's stability criteria, Young's modulus (Y) and Poisson's ratio (ν), respectively. These properties are determined from the elastic constants C11, C12, and C66, which are derived by calculating the second partial derivatives of the total energy of the system with respect to the corresponding components of strain, thereby linking the material's atomic-level interactions to its macroscopic mechanical properties.36
For 2D hexagonal materials such Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers, the Born's mechanical stability criteria are simplified due to their symmetry and two-dimensional nature as follows: C11 > 0, C66 > 0, and C11 − C12 > 0.37,38 As reported in Table 1, the calculated C11, C12 and C66 moduli of Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers comply these criteria. These findings are crucial, as the Born criteria ensure the strength of materials under small deformations and indicate the absence of lattice instabilities. Furthermore, the Young's modulus and Poisson's ratio as angular functions were calculated in the xy plane. As plotted in Fig. 2(c) and (d) for Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers, respectively, the graphs of these functions are nearly perfect circles. This angular dependence directly correlates with the structural isotropy and uniformity of these monolayers, underscoring their potential for applications where uniform mechanical properties are essential.39,40 The calculated Young's modulus values are 65.31 N m−1 for Ga2Ge2S3Se3 and 50.02 N m−1 for Ga2Ge2Se3Te3, placing them among 2D materials with moderate stiffness such as GaGeSe3, GaGeS3, Al2Ge2S3Se3, Al2Ge2Se3Te3, tinselenidene and bismuth telluride nanosheets.22,32,41,42 Meanwhile, Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers are more flexible than phosphorene, graphene, black phosphorus (BP), hexagonal boron nitride and MoS2 monolayer.43–45 The reduced modulus compared to graphene can be attributed to the buckled structure and the chemical composition of these monolayers, which result in lower in-plane bonding strength. This intermediate stiffness makes Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 suitable for applications requiring flexibility without significant loss of mechanical stability. In addition, the Poisson's ratios of Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 range from 0.39 to 0.41, exceeding those observed in many ternary and quaternary monolayers of similar atomic structures, including GaGeS3, GaGeSe3, and Al2Ge2Se3Te3.22,32 A high Poisson's ratio indicates the material's ability to undergo lateral expansion when stretched, which is a desirable trait for enhancing toughness and ductility. For comparison, the Poisson's ratios of typical 2D transition metal dichalcogenides MoS2 and WS2, tin chalcogenides, boron nitride (h-BN)46–48 hover around 0.3, making Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 superior in terms of toughness and deformation resilience.
This combination of moderate stiffness and high Poisson's ratio suggests that Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers can endure significant deformation without losing their structural integrity. The demonstration of FeClF, MnSeTe, and MnSTe monolayers as promising materials for a wide range of electronic applications, particularly due to their favorable elastic properties,49,50 provides strong support for considering Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 as potential candidates for flexible electronic devices, such as foldable displays, wearable sensors, and stretchable photovoltaics, where both mechanical resilience and flexibility are required.51–53 Furthermore, their structural and mechanical characteristics could enable their integration into composite materials to enhance performance under mechanical stress.
Compared to the HSE06 band gaps of GaGeS3 (2.51 eV) and GaGeSe3 (1.91 eV) monolayers,22 replacing the S sublayer with a Se sublayer or the Se sublayer with a Te sublayer reduces the band gaps of the original monolayers. These results highlight that substituting heavier element provides an efficient strategy for narrowing the band gaps of monolayers. This effect results from the increased spin–orbit coupling (SOC) induced by heavier atoms, which brings the conduction and valence band edges closer together.54,55 In two-dimensional materials, the impact of SOC is further amplified by quantum confinement,56,57 leading to more pronounced band gap reductions and tunable electronic properties. The band gap of Ga2Ge2Se3Te3 is below 1.23 eV, rendering it unsuitable as a photocatalyst for water splitting due to insufficient energy for driving the reaction.58 In contrast, the Ga2Ge2S3Se3 monolayer is well-suited for this application. Additionally, it represents an advancement over the GaGeS3 monolayer,22 as its narrower band gap enables greater absorption of visible light, which makes up a significant portion of the solar energy spectrum. In the Ga2Ge2S3Se3 monolayer, the PDOS reveals that p-orbitals from Ga, Ge, S, and Se dominate both the valence and conduction bands across a wide energy range. Strong hybridization is observed, particularly between the s- and p-orbitals of Ga and Ge. The direct bandgap is formed by the valence band maximum (VBM) dominated by Se-p orbitals and the conduction band minimum (CBM) dominated by Ga-s orbitals. A similar PDOS profile is observed in the Ga2Ge2Se3Te3 monolayer, with p-orbitals from the constituent atoms dominating at all energy levels. The VBM is primarily composed of Te-p orbitals, while the CBM is dominated by Se-s, p orbitals and p-orbitals from the other atoms.
To gain deeper insights into the orbital contributions of each constituent element to the electronic band structure, we analyzed the projected band structures of Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers, as depicted in Fig. 4. In these plots, the s, px, py, and pz orbitals are depicted in blue, pink, red, and green, and their relative contributions are visualized by the size of the corresponding spheres. By examining these projected band structures, we can identify the specific orbitals of each element that dominate the formation of the valence band maximum and conduction band minimum, providing valuable information about the electronic and hole transport properties of these materials. It is obvious that the highest valence bands are mainly occupied by Se-p orbitals in Ga2Ge2S3Se3 monolayer and by Te-p orbitals in Ga2Ge2Se3Te3 monolayer. Consequently, the width of the band gap is partially determined by the substituted Se or Te element. It is worth noting that the lowest conduction bands, on the other hand, are mainly constructed by the Ga-s orbitals, making them another important factor in creating the band gap.
In Al2Ge2S3Se3 and Al2Ge2Se3Te3 monolayers, the bandgaps range from 1.07 to 2.38 eV.32 These values are smaller than those observed in Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers. This reduction in bandgap can be attributed to the lighter atomic mass of Al compared to Ga, resulting in weaker SOC effects in the Al-containing compounds. When analyzing the bandgaps of In2Ge2S3Se3, In2Ge2S3Te3, Al2Ge2S3Se3, and Al2Ge2Se3Te3 monolayers32,59 within the A2B2X3Y3 family, it is observed that the Y atoms (Se or Te) play a significant role in determining the bandgap width. Specifically, heavier Y atoms result in narrower bandgaps. These findings confirm the mechanism discussed in the previous section, which emphasizes the significant influence of SOC on bandgap values. At lower valence bands, ranging from −2 to −1 eV, orbital hybridization is observed to happen between pz orbitals from Ge, S, and Se elements in Ga2Ge2S3Se3 monolayer or Ge, Se, and Te elements in Ga2Ge2Se3Te3 monolayer. The hybridization of other orbitals happens in the deeper energy levels. Meanwhile, the py orbitals from Se or Te atoms participate in the hybridization with other orbitals at almost all valence energy bands, showing the important role of Se and Te elements in the formation of covalent bonds in Al2Ge2S3Se3 and Al2Ge2Se3Te3 monolayers.
Strain engineering emerges as a powerful tool for enhancing the photocatalytic activity of two-dimensional materials, particularly Janus structures like Ga2Ge2S3Se3. By applying strain, precise control over the bandgap and band alignment becomes possible. This allows for optimization of light absorption and efficient charge carrier transport, crucial for driving photocatalytic reactions.63–66 Furthermore, strain can effectively align the band edges with the redox potentials of water splitting reactions. This alignment facilitates efficient charge transfer and improves the overall catalytic efficiency.67,68 Most of 2D materials can withstand substantial strains, typically up to around 10%,69 before significant deformation or failure. This high flexibility stems from their unique atomic structure. However, it's crucial to recognize that this value is not universal and depends strongly on the specific material and the nature of the applied strain. For example, reported strain limits for 2D GeP range from 12.9% to 26.2%,70 and MoS2 monolayers have been shown to maintain stability above 10% strain.71 The Ga2Ge2S3Se3 and Ga2Ge2Se3Te3 monolayers, with their moderate Young's moduli (50.02–65.31 N m−1) and high Poisson's ratios (0.39–0.41), are predicted to be relatively flexible. Therefore, the compressive and tensile strains ranging from −8% to 8% were applied on these two monolayers. Our investigation delves into the impact of strain on the bandgap and band edge alignment of the Ga2Ge2S3Se3 monolayer (Fig. 5(b)). Compressive strains ranging from −2% to −4% lead to a bandgap widening of approximately 2.0 eV. Beyond this range, both further compression and tensile strains result in a bandgap reduction. Notably, the band edges of the Ga2Ge2S3Se3 monolayer remain suitable for water redox reactions when subjected to compressive strains between −6% and −8% or tensile strains below 4%.
Strain engineering offers a versatile approach to fine-tune the electronic and optical properties of 2D materials, including their bandgap, band edge positions, and light absorption characteristics. These tunable properties hold significant potential for enhancing photocatalytic performance in applications such as water splitting.72–74 Given the promising photocatalytic potential of Ga2Ge2S3Se3 monolayer, we investigated the effects of strain ranging from −8% to 8%. The resulting changes in band structure and light absorption rate are illustrated in Fig. 6, respectively.
![]() | ||
Fig. 6 (a) Impact of strain on the band structure and (b) the light absorption rate of the Ga2Ge2S3Se3 monolayer. |
Fig. 6(a) presents the band structures with the valence band maximum (VBM) aligned to the Fermi level, allowing us to focus on the shifts in the conduction band minimum (CBM). Previous studies on black phosphorus, transition metal dichalcogenides, and ternary Janus monolayers have demonstrated that strain can induce significant band structure modifications, such as direct-to-indirect bandgap transitions, bandgap tuning, and changes in band curvature.75,76 In contrast, the Ga2Ge2S3Se3 monolayer maintains a direct bandgap configuration under both compressive and tensile strain, with the VBM and CBM remaining at the Γ – point. Notably, the curvature of the conduction bands increases significantly under strain compared to the unstrained monolayer (Fig. 3). This increased curvature suggests a reduced effective mass of electrons, potentially leading to enhanced electron mobility.
The variation in electronic structure, particularly the changes in bandgap discussed above, significantly influences the optical properties of Ga2Ge2S3Se3 monolayers. Fig. 6(b) presents the calculated absorption rates, α(ω), for the monolayer under strains ranging from −8% to 8%. The unstrained Ga2Ge2S3Se3 monolayer exhibits a relatively high absorption rate (black curve), exceeding 105 cm−1 in the infrared region and increasing further in the visible region. In the ultraviolet region, the absorption rate surpasses 106 cm−1. This strong absorption across a broad spectral range, including the infrared and visible regions, makes Ga2Ge2S3Se3 a promising candidate for solar energy applications, as these regions contain the majority of solar energy reaching Earth. Compressive strains (−4% and −8%) lead to a widening of the bandgap, resulting in a reduced absorption of infrared and visible light. However, these larger bandgaps enable the absorption of higher-energy ultraviolet photons. Conversely, tensile strains (4% and 8%) narrow the bandgap, enhancing the absorption of infrared and visible light.
A critical step in water splitting involves the adsorption of hydrogen atoms onto the catalyst surface, facilitating electron transfer. This adsorption process induces a change in the system's total energy, which can be analyzed through Gibbs free energy calculations. To calculate the change in Gibbs free energy for the Ga2Ge2S3Se3 monolayer, it is necessary to determine the total energies of hydrogen-adsorbed system EGa2Ge2S3Se3+H, pristine Ga2Ge2S3Se3 monolayer EGa2Ge2S3Se3, and isolated hydrogen molecule EH2. Then the change in energy ΔEH is calculated as ΔEH = EGa2Ge2S3Se3+H − EGa2Ge2S3Se3 − 0.5EH2, and the Gibbs free energy ΔGH is defined as ΔGH = ΔEH + 0.24 eV.77–79 Fig. 7(a) and (b) depicts the Gibbs free energy diagram of the Ga2Ge2S3Se3 monolayer, wherein the bottom and top surfaces are designated as the S-side and Se-side, respectively. The calculated ΔGH values at the Ga, Ge, and Se sites exhibit negative values, ranging from −0.95 to −2.01 eV. The negative ΔGH suggests that the adsorption of hydrogen atoms at these specific sites is thermodynamically favorable.
The Ga2Ge2S3Se3 monolayer's capacity for photocatalytic water splitting can be assessed by examining the thermodynamic feasibility of both the Hydrogen Evolution Reaction (HER) and the Oxygen Evolution Reaction (OER).59,80 The Gibbs free energy changes (ΔG) associated with each of these half-reactions at pH = 0 are shown in Fig. 7(c) and (d). The OER half-reaction on the Ga2Ge2S3Se3 monolayer involves a series of steps. Initially, energy is required to break the O–H bond in water to form an adsorbed hydroxyl ion (OH*). Subsequently, this hydroxyl ion undergoes further oxidation to generate an adsorbed oxygen atom (O*). The adsorbed oxygen atom then reacts with another water molecule to form a hydroperoxide intermediate (OOH*), which finally decomposes to release molecular oxygen (O2). In the absence of light irradiation (U = 0 eV), the calculated ΔG values for these four steps are 1.38 eV, 3.71 eV, 4.64 eV, and 4.92 eV, respectively. The highest energy barrier, and thus the potential-determining step, is the transformation from OOH* to O2. As mentioned in previous sections, the electronic characteristics of Ga2Ge2S3Se3 monolayer is favorable for the generation of photoexcited electron–hole pairs. The photogenerated holes in the valence bands can effectively interact with the intermediate species formed during the OER half-reactions, thereby facilitating the subsequent oxidation steps. Upon illumination, the energy barriers for the OER steps are modified. The calculation method of energy barriers of the illuminated material is presented in ESI.† As shown in Fig. 7(d), these effective barriers are reduced to negative or near-zero values, indicating that the OER steps become spontaneous or require minimal activation energy. These results strongly suggest that light irradiation can effectively trigger the OER half reactions on the Ga2Ge2S3Se3 monolayer spontaneously.
In water splitting, the OER oxidizes water to produce oxygen O2, protons H+, and electrons. These products are subsequently utilized by the HER to generate hydrogen gas H2. However, the thermodynamic feasibility of this process is governed by the change in Gibbs free energy ΔG. Fig. 7(c) illustrates the ΔG for the HER process on the Ga2Ge2S3Se3 monolayer. Without light irradiation, the formation of adsorbed hydrogen atoms H* requires an energy input of 1.36 eV. Under light irradiation, photogenerated electrons contribute an external potential of U = 0.47 eV, effectively reducing the HER reaction barrier on the Janus Ga2Ge2S3Se3 to 0.89 eV. Many promising photocatalysts exhibit positive HER reaction barriers, which can be mitigated through various strategies such as adsorbing alkali and transition metals,81 constructing heterojunctions,82 and introducing point defects like vacancies.83 In practice, applying an external bias voltage can significantly enhance the separation of photogenerated electron–hole pairs, thereby improving the overall photocatalytic efficiency. For example, studies have demonstrated that combining a TiO2 photoanode with an organo-photocathode enables stoichiometric water decomposition at bias voltages below the theoretical minimum of 1.23 V. This exemplifies the effectiveness of external voltage in lowering the energy barrier for water splitting reactions and boosting photocatalytic activity. However, the wide bandgap of TiO2 (approximately 3.2 eV) restricts its activity to the ultraviolet region of the solar spectrum, which constitutes a relatively small portion of solar energy.84 Therefore, the Ga2Ge2S3Se3 monolayer, with its moderate bandgap and low HER energy barrier, emerges as a promising candidate for efficient hydrogen generation through photocatalytic water splitting.
To calculate the charge carrier mobility, it is necessary to establish the dependence of the system total energy Etotal and band edges Eedge (the positions of VBM and CBM) as functions of the uniaxial strains εuni. As shown in Fig. 8(a), the Etotal values are calculated at 5 discrete values of the εuni, which ranges from −1% to 1%. The Eedge values, as shown in Fig. 8(b), are also determined at these discrete values of uniaxial strains. Therefore, an interpolation model was applied to construct the continuous functions Etotal and Eedge, making it possible to the determine the elastic modulus C2D by taking the second derivative of the total energy function and divided by the area of the unit-cell. Meanwhile, the deformation energy Edef is derived from the Laplacian of Eedge the edge energy function. The charge carrier mobility in 2D material, μ2D, is then defined as μ2D = eℏ3C2D/kBTm**Edef2, where e, ℏ, kB, and T represent charge of a free electron, the temperature in kelvin, the reduced Planck constant, and the Boltzman constant, respectively. The symbol
* is the average value of the effective masses m*, which is defined based on the energy function of k – point E(k) as m* = (ℏ2∂2Ek/∂2k)−1.
![]() | ||
Fig. 8 (a) Total energy Etotal and (b) band edge energy Eedge as functions of uniaxial strains εuni applied on the Ga2Ge2S3Se3 monolayer. |
The Ga2Ge2S3Se3 monolayer exhibits nearly isotropic electron mobility in the range of 430.82–461.50 cm2 V−1 s−1, enabling efficient charge transport in all directions. This isotropic mobility minimizes anisotropic losses that can hinder device performance in complex architectures.85,86 Moreover, the Ga2Ge2S3Se3's electron mobility is comparable to well-established 2D photocatalysts like MoS2, PdS2, PdSe2, SnP2S6, and As2S3,87–91 facilitating rapid electron transport to the reduction site for reactions such as hydrogen evolution. The Ga2Ge2S3Se3 monolayer exhibit significantly higher electron mobility compared to their hole mobility (19.55–33.47 cm2 V−1 s−1). This disparity can suppress electron–hole recombination, a major loss mechanism in photocatalytic processes.
The previous sections have established Ga2Ge2S3Se3 monolayer as a promising material for photocatalytic applications. Its 1.83 eV bandgap aligns well with the solar spectrum, making it an ideal candidate for sunlight-driven water splitting. To further explore its potential, additional calculations were performed to assess its performance in this process. The solar-to-hydrogen efficiency ηSTH of Ga2Ge2S3Se3 monolayer, a measure for its ability to convert solar energy into hydrogen, is determined by two primary factors: light absorption efficiency ηabs and carrier utilization efficiency ηcu. As outlined in equations ηSTH = ηabs × ηcu,92 ηabs, which quantifies the photocatalyst's ability to absorb sunlight, is primarily governed by the semiconductor's bandgap Eg. The Ga2Ge2S3Se3 monolayer exhibits a high light absorption rate α(ω) of 105–106 cm−1 across a broad spectral range from infrared to ultraviolet, leading to the ηabs of 32.18%. Carrier utilization efficiency ηcu, on the other hand, reflects the efficiency of converting absorbed sunlight into usable charge carriers. As shown in the supplementary, this parameter is influenced the bandgap Eg. The calculated value of ηcu is 47.54%. Based on the values of ηabs and ηcu, the ηSTH of Ga2Ge2S3Se3 monolayer is predicted to be 15.30%. However, this value must be corrected92 because in the Janus monolayers with difference in vacuum level (ΔEV), the intrinsic electric field can do positive work Win in separating the electron–hole pair. As defined in the discussion regarding to Fig. 5(a), the ΔEV equals to the difference in work function ΔΦ, which is 0.27 eV. With P(ℏω) is the AM1.5G solar energy flux, the work Win is determined as . Therefore, the total work for photocatalytic process becomes
. The actual efficiency
is
. Applying this correction, the efficiency of Ga2Ge2S3Se3 monolayer becomes 14.80%, making this monolayer a promising candidate for photon-driven water splitting photocatalysis. As mentioned above, a water splitting photocatalyst must meet many strict requirements. Therefore, to achieve the
in the range of 3.33–16%, a monolayer is usually modified by elemental substitution, phase transitions, or the application of strain.22,59,62,93,94 The Ga2Ge2S3Se3 monolayer offers several advantages over its parent GaGeS3 monolayer.22 Notably, its solar-to-hydrogen conversion efficiency (14.80%) surpasses that of GaGeS3 (11.33%). The transition from an indirect to a direct bandgap in Ga2Ge2S3Se3 enhances light absorption and facilitates faster charge separation. While direct bandgap materials are often susceptible to rapid electron–hole recombination, this effect is mitigated in Ga2Ge2S3Se3 due to an intrinsic electric field that promotes charge separation. Consequently, the substitution of Se for S results in a monolayer with demonstrably improved photocatalytic properties.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra00812c |
This journal is © The Royal Society of Chemistry 2025 |