1. Introduction
With the depletion of traditional fossil fuels and the escalating global energy crisis, it is imperative and urgent to explore green and renewable energy sources. The use of semiconductor materials in applications such as photocatalysis or solar cells to convert abundant solar energy into clean power holds significant promise [
1]. For instance, Fujishima and Honda were pioneers in demonstrating that TiO
2 could serve as a photocatalyst for water splitting [
2]. Nevertheless, the efficiency of TiO
2 in converting solar energy to hydrogen is hindered by its wide band gap and high rate of carrier recombination. Chapin et al. were the first to create a solar cell using single-crystal silicon as the primary material. However, the photoelectric conversion efficiency (PCE) was disappointingly low, measuring only 6% [
3]. As a result, the quest for suitable materials for photocatalysis and photovoltaics has been a prominent research area for a considerable period of time.
The discovery of graphene has sparked researchers’ interest in two-dimensional (2D) materials [
4,
5]. The 2D materials demonstrate amazing properties, including high carrier mobility, a semiconducting band gap, prominent catalytic activities, and abundant active sites. Therefore, they can be utilized in the fields of photocatalytic water splitting and photovoltaics. At present, many 2D materials have been synthesized experimentally or theoretically, such as transition metal carbides/nitrides (MXenes) [
6], transition metal dichalcogenides (TMDCs) [
7], hexagonal boron nitride (h-BN) [
8], black phosphorus (BP) [
9], and silicene [
10]. However, 2D materials have a large band gap, poor light absorption capacity, and a high carrier recombination rate, thereby leading to low efficiency. Therefore, various strategic techniques such as doping [
11], metal loading [
12], and constructing heterostructures have been proposed. Among these strategies, constructing van der Waals (vdW) heterostructures with type-II band alignment has promising applications in the fields of photocatalytic water splitting and solar cells due to the lower exciton binding energy and enhanced optical absorbance compared to monolayers [
13]. In type-II heterostructures, the photogenerated electron–hole pairs are separated onto different monolayers, which significantly reduces the carrier recombination rate. With the deepening of research, direct Z-scheme heterostructures can be designed by selecting two appropriate monolayer materials. In the Z-scheme heterostructure, photogenerated electrons and holes accumulate on the surfaces of distinct monolayers. The Z-scheme heterostructure not only possesses a strong redox ability to drive photocatalytic reactions but also provides active sites for spatially separated oxidation and reduction processes [
14]. This mechanism significantly enhances the efficiency of water splitting in the heterostructure. According to previous research, the narrow band gap of the direct Z-scheme heterostructures can achieve a broader range of solar energy harvesting [
15]. The Z-scheme heterostructures show great promise in photocatalytic water splitting, photocatalytic reduction of carbon dioxide, and environmental remediation [
16,
17]. In recent years, more and more Z-scheme heterostructures have been discovered and studied. Indeed, examples such as the WO
3/Bi
2MoO
6 heterostructure [
18], β-SnSe/HfS
2 heterostructure [
19], GaSe/ZrS
2 heterostructure [
20], MoSTe/g-GeC heterostructure [
21], GeC/BSe heterostructure [
22], and SnC/PtS
2 heterostructure [
23] all represent direct Z-scheme heterostructures.
On the other hand, MXenes have been widely explored in applications such as photocatalysts, solar cells, heavy-metal removal, battery anodes, and electromagnetic interference shielding. MXenes are produced from their corresponding MAX phases, where M represents an early transition metal, A represents a group of IIIA or IVA elements, and X represents a C or N atom [
24]. MXenes have attracted increasing attention due to their excellent stability and large specific surface area. In the field of photocatalysis, heterostructures based on MXenes, such as Cs
2AgBiBr
6/Ti
3C
2T
x [
25], Hf
2CO
2/WS
2 [
26], AsP/Sc
2CO
2 [
27], and Sc
2CF
2/MoSSe [
28], exhibit superior electronic properties. For the application of solar cells, Wen et al. demonstrated that the PCE of Hf
2CO
2/MoS
2 and Zr
2CO
2/MoS
2 heterostructures in solar cell applications was 19.75% and 17.13%, respectively [
29]. The PCE of Ti
2CO
2/Zr
2CO
2 and Ti
2CO
2/Hf
2CO
2 heterostructures reaches 22.74% and 19.56%, respectively [
30]. This indicates that MXenes have promising potential for applications as photovoltaic materials. Pure Sc
2C exhibits metallic properties; however, after functionalization by F, Cl, and Br atoms, Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 exhibit semiconductor characteristics with band gaps of 1.85 eV, 1.70 eV, and 1.54 eV, respectively [
31]. As a member of MXenes, Sc
2CX
2 (X = F, Cl, Br) exhibits kinetic and thermal stabilities, which have potential applications in photocatalytic water splitting and solar cells [
32]. However, the Sc
2CF
2 monolayer cannot facilitate the oxygen evolution reaction (OER) because its valence band maximum (VBM) is higher than that of E
O2/H2O. For Sc
2CCl
2 and Sc
2CBr
2 monolayers, the conduction band minimum (CBM) is lower than E
H+/H2, which renders them unable to meet the requirements for the HER. The construction of heterostructures using Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 not only addresses the mentioned deficiency of materials but also shows significant potential for photocatalytic and optoelectronic applications. Zhang et al. investigated the electrical and optical properties of Sc
2CF
2/WSSe heterostructures and found that they have the potential for water splitting [
33]. In addition, Sun et al. revealed that the PCE of the Sc
2CCl
2/SiS
2 heterostructure can reach 23.20%, indicating promising prospects for application in the field of solar cells [
34]. It is noteworthy that the VBM and CBM of the Sc
2CF
2 monolayer are higher than those of the Sc
2CCl
2 (or Sc
2CBr
2) monolayer, and the VBM and CBM of the Sc
2CBr
2 monolayer are higher than those of the Sc
2CCl
2 monolayer. This indicates that the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures may have a type-II band alignment. In addition, the CBM in Sc
2CCl
2 (or Sc
2CBr
2) and the VBM in Sc
2CF
2 are very close. This suggests that photogenerated carrier transfer in the Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures may follow the Z-scheme pathway. Therefore, it is worthwhile to study the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures. Their potential applications in photocatalytic water splitting and solar cells show great promise.
In this paper, three types of monolayers, namely Sc2CF2, Sc2CCl2, and Sc2CBr2, were successfully vertically stacked to create Sc2CF2/Sc2CCl2, Sc2CF2/Sc2CBr2, and Sc2CCl2/Sc2CBr2 heterostructures. The stacking geometries, electronic, and optical properties of the heterostructures have been systematically studied based on first-principles calculations. According to band edge alignment and charge carrier transfer processes, the Sc2CF2/Sc2CCl2 and Sc2CF2/Sc2CBr2 heterostructures were found to have a direct Z-scheme band alignment, making them promising for photocatalytic water splitting applications. On the other hand, the Sc2CCl2/Sc2CBr2 heterostructure showed potential for use in solar cells, with a notable PCE of 20.78%. The present findings indicate that Sc2CX2/Sc2CY2 (X, Y = F, Cl, Br) heterostructures have the potential for application in solar energy conversion.
3. Results and Discussion
The structural parameters and electronic properties of Sc
2CX
2 (X = F, Cl, Br) were initially studied. The atomic structures of optimized Sc
2CX
2 (X = F, Cl, Br) monolayers are displayed in
Figure 1a. The lattice constant of the Sc
2CF
2 monolayer was determined to be 3.235 Å, which closely matches the theoretical value of 3.26 Å, as reported by Khang et al. [
42]. The corresponding result of 3.422/3.499 Å for the Sc
2CCl
2/Sc
2CBr
2 monolayer is close to the previous theoretical value of 3.42/3.507 Å [
31,
43]. When the surface groups change from F to Br, the lattice parameters increase slightly due to the increase in the halogen atomic radius [
31]. In addition, the band structures of the Sc
2CX
2 (X = F, Cl, Br) monolayers were calculated using the HSE06 method, as displayed in
Figure 1b–d. It can be distinctly observed that the band shapes are fundamentally the same, despite the differences in band gap values. Moreover, we can observe that Sc
2CX
2 (X = F, Cl, Br) monolayers are all indirect band gap semiconductors. The CBM and VBM of the Sc
2CX
2 (X = F, Cl, Br) monolayers are located at the M point and Γ point, with corresponding band gaps of 1.80 eV, 1.70 eV, and 1.55 eV, respectively. All band gap values are in good agreement with the earlier reports, with percentage differences of less than 2% [
32,
44,
45]. The results verify the rationality of our approach and parameterization.
Then, the structural properties of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures were researched in pursuit of the most stable configuration. There are three typical stacking configurations for all three heterostructures, i.e., A, B, and C, as illustrated in
Figure 2. The structure coordinate information (POSCAR) is provided in
Table S1.
Table 1 presents various parameters associated with different stackings. For each heterostructure, the lattice constants of the three configurations closely match the lattice constants of the corresponding monolayer. In order to assess the stability of the heterostructures and determine the most stable configurations, the binding energy (
Eb) values of all configurations are computed as follows:
where
represents the energy of the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, respectively. Here,
represents the interface area, while
and
represent the energy of the Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers, respectively. From
Table 1, we can see that the minus
Eb values for all stacking configurations manifest that the interface formation is exothermic, which is favorable for their preparation [
46]. Clearly, for Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, stacking-B exhibits the smallest
Eb of −35.67 meV∙Å
−2, −28.53 meV∙Å
−2, and −19.96 meV∙Å
−2, indicating that stacking-B is the most stable among the three stacking configurations. In addition, this value is smaller than the previously reported C
2N/ZnSe heterostructure (−12.1 meV∙Å
−2) [
47] and BiTeCl/GeSe heterostructure (−11.07 meV∙Å
−2) [
48], revealing that Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 are vdW heterostructures. Thus, only the stacking-B heterostructure was taken into consideration in all the following calculations. Indispensably, AIMD simulations are performed to validate the thermodynamic stability of the heterostructure. As depicted in
Figure S1, the geometrical structures of the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures remained stable during the 6 ps simulation at a temperature of 300 K. No bonds were broken, and the energy fluctuation was minimal, indicating that each heterostructure is sufficiently stable at room temperature. Furthermore, to verify the dynamical stability of the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, we calculated their phonon spectrum with a 3 × 3 × 1 supercell and implemented them in the PHONOPY code with the density functional perturbation theory (DFPT), as shown in
Figure S2. It can be seen that there are some insignificant imaginary frequencies near the G-point. This phenomenon also exists in the phonon spectra of some experimentally prepared 2D materials, but the imaginary frequency near the G-point can be ignored [
49,
50,
51]. This phenomenon may be attributed to inadequate computational accuracy, which can be eliminated by creating a larger supercell or setting a higher parameter accuracy. Thus, the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures are dynamically stable.
The projected band structures of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures were calculated based on the HSE06 hybrid functional, as depicted in
Figure 3a–c. It can be found that the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures all show the characteristics of semiconductors with indirect band structures. The VBM and CBM are located at the M point and Γ point, with band gaps of 0.58 eV, 0.78 eV, and 1.35 eV, respectively. Compared with the band gaps of monolayers, the significantly reduced band gaps of heterostructures are due to the interaction of vdW forces, which lead to a change in the band structure upon contact [
21]. It should be noted that the smaller band gap of Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures can lead to improved optical absorption performance during the photocatalytic reaction process. In addition, we can clearly see that the VBM and CBM of the three heterostructures are each occupied by two monolayers, demonstrating an inherent type-II heterostructure. Among them, the VBM of Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures is mainly attributed to the Sc
2CF
2 monolayer, while the CBM mainly comes from the Sc
2CCl
2 (or Sc
2CBr
2) monolayer. Hence, electrons mainly occupy Sc
2CCl
2 (or Sc
2CBr
2), while holes mainly occupy Sc
2CF
2. Similarly, the VBM of the Sc
2CCl
2/Sc
2CBr
2 heterostructure is mainly contributed by the Sc
2CBr
2 layer, whereas the CBM is entirely dominated by the Sc
2CCl
2 layer. It is certain that the type-II band structures can separate the photoexcited electrons and holes into different monolayers, which is conducive to reducing the carrier recombination rate. This separation can improve the utilization of photogenerated carriers and extend their lifetime [
44].
In addition,
Figure 3d–f shows the projected density of states (PDOS) of the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, respectively. From
Figure 3d, it can be seen that in the Sc
2CF
2/Sc
2CCl
2 heterostructure, the peak with the highest energy below the Fermi level mainly originates from the Sc and C atoms in Sc
2CF
2, while the peak with the lowest energy above the Fermi level is mainly contributed by the Sc atom in Sc
2CCl
2. This shows that the VBM of the Sc
2CF
2/Sc
2CCl
2 heterostructure is contributed by the Sc
2CF
2, while the CBM is contributed by the Sc
2CCl
2. As shown in
Figure 3e, the VBM of the Sc
2CF
2/Sc
2CBr
2 heterostructure is mainly contributed by the Sc and C atoms of Sc
2CF
2, while the CBM mainly comes from the Sc atom of Sc
2CBr
2. This indicates that the VBM of the Sc
2CF
2/Sc
2CBr
2 heterostructure comes from the electronic states of Sc
2CF
2, while the CBM comes from the electronic states of Sc
2CBr
2. In
Figure 3f, we can clearly observe that the CBM of the Sc
2CCl
2/Sc
2CBr
2 configuration is contributed by the Sc atom of Sc
2CCl
2. However, the VBM is not only contributed by the Sc and C atoms but also by the Br atom. This shows that the VBM of the Sc
2CCl
2/Sc
2CBr
2 heterostructure originates from the Sc
2CBr
2 monolayer, while the CBM comes from the Sc
2CCl
2 monolayer. In addition, the orbitals of the C atom and Sc atom are completely hybridized. This PDOS result further confirms that the CBM and VBM of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures are located on different monolayers.
In
Figure 3g–i, we displayed the band decomposed charge densities of the VBM and CBM in Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, respectively. In Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures, it can be observed that the VBM is located in Sc
2CF
2, while the CBM is located in Sc
2CCl
2 (or Sc
2CBr
2). Consistent with the above analysis, the VBM and CBM of the Sc
2CCl
2/Sc
2CBr
2 heterostructure are located on the lower layer (Sc
2CBr
2) and upper layer (Sc
2CCl
2), respectively. There is no charge density overlap between the VBM and CBM, indicating that heterostructures like Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 can effectively separate electrons and holes [
52].
The above analysis shows that the Sc2CF2/Sc2CCl2, Sc2CF2/Sc2CBr2, and Sc2CCl2/Sc2CBr2 heterostructures exhibit staggered type-II band alignment. This structure can promote the effective separation of holes and electrons, reduce the carrier recombination rate, and play an important role in photocatalytic water splitting and optoelectronic devices.
The difference in work functions between two semiconductors can lead to charge redistribution and the formation of an electric field at the interface. This electric field will determine the transfer process of photogenerated charges. Thus, the work functions of the Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers, as well as the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, are calculated using the following formula:
in which
and
represent the vacuum level and Fermi level, respectively. As shown in
Figure S3a–c, Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers exhibit a fixed work function of 5.02 eV, 5.86 eV, and 5.48 eV, respectively, due to their highly symmetrical crystal structure [
28]. Compared to Sc
2CCl
2 and Sc
2CBr
2 monolayers, the Sc
2CF
2 monolayer exhibits a smaller work function and a higher Fermi level. Thus, in the Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures, free electrons can migrate from Sc
2CF
2 to Sc
2CCl
2 (or Sc
2CBr
2) until their Fermi levels reach equilibrium. As shown in
Figure 4a,b, the work functions of the Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures are 5.19 eV and 4.99 eV, respectively. At the same time, there are potential drops of 5.43 eV and 3.25 eV at the Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures, indicating the presence of a built-in electric field at the interface of the heterostructures [
52]. It also indicates that electrons are inclined to flow to Sc
2CCl
2 (or Sc
2CBr
2) monolayers. The built-in electric field will create a driving force to promote the combination of photogenerated electron–hole pairs between the electrons in the CBM of Sc
2CCl
2 (or Sc
2CBr
2) and the holes in the VBM of Sc
2CF
2. As displayed in
Figure 4c, the difference in monolayer work function leads to the transfer of electrons from Sc
2CBr
2 to Sc
2CCl
2, causing a decline in the Fermi level in Sc
2CCl
2 and Sc
2CBr
2. The work function of the heterostructure in the final equilibrium state is 5.34 eV. Moreover, a potential drop of 2.12 eV is found across the interface. This is proof of a built-in electric field at the interface of the Sc
2CCl
2/Sc
2CBr
2 heterostructure.
During the formation of a heterostructure, the charge near the interface will be redistributed due to the presence of interlayer interactions. In order to explore the charge transfer mechanism of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, the planar averaged charge density difference and 3D differential charge density difference were calculated using the following equation:
where the
stand for the density of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, and the
and
represent the corresponding densities of Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers. As shown in
Figure 4d,e, for Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2, it can be clearly seen that a large number of negative charges are assembled in the side of Sc
2CCl
2 (or Sc
2CBr
2) monolayers, while positive charges cluster on the side of Sc
2CF
2. This leads to the formation of a built-in electric field from Sc
2CF
2 to Sc
2CCl
2 (or Sc
2CBr
2). As shown in
Figure 4f, the electrons at the interface are depleted near the Sc
2CBr
2 monolayer and accumulate at the Sc
2CCl
2 monolayers, forming a built-in electric field from Sc
2CBr
2 to Sc
2CCl
2. In addition, the Bader charges obtained indicate that about 0.0072 |e| (0.0052 |e|) are transferred from the Sc
2CF
2 monolayer to the Sc
2CCl
2 (or Sc
2CBr
2) monolayers in the case of the Sc
2CF
2/Sc
2CCl
2 (Sc
2CF
2/Sc
2CBr
2) heterostructure. Furthermore, around 0.0018 |e| is transferred from Sc
2CBr
2 to Sc
2CCl
2 within the Sc
2CCl
2/Sc
2CBr
2 heterostructure.
In addition to the band gap value, the band edge alignment is also a crucial parameter for evaluating the application of the heterostructure. Therefore, we computed the band alignments of Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers, as well as the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, using the method suggested by Toroker et al. [
53].
Figure 5a reveals that the VBM of the Sc
2CF
2 monolayer exceeds that of E
O2/H2O. The Sc
2CCl
2 and Sc
2CBr
2 monolayers exhibit very similar characteristics in their band edge alignments, with both CBM being lower than the energy level of E
H+/H2. Based on the aforementioned analysis, the band positions of the Sc
2CF
2, Sc
2CCl
2, and Sc
2CBr
2 monolayers are unsuitable for photocatalysis. For the Sc
2CF
2/Sc
2CCl
2 and Sc
2CF
2/Sc
2CBr
2 heterostructures, the VBM and CBM of the Sc
2CF
2 layer are higher than those of the Sc
2CCl
2 (or Sc
2CBr
2) layer, further affirming that the heterostructure exhibits a type-II band alignment. For the Sc
2CCl
2/Sc
2CBr
2 heterostructure, both the VBM and CBM of Sc
2CBr
2 exceed those of the Sc
2CCl
2 layer, indicating a type-II band alignment. The CBM of Sc
2CCl
2 is lower than that of E
H+/H2, making the Sc
2CCl
2/Sc
2CBr
2 heterostructure unsuitable for photocatalytic water splitting reactions.
The photocatalytic water splitting reaction mechanism of Sc
2CF
2/Sc
2CX
2 (X = Cl, Br) is shown in
Figure 5b. In general, three possible processes are considered here: ① The photoexcited holes at the VBM of Sc
2CF
2 recombine with electrons at the CBM of Sc
2CCl
2 (or Sc
2CBr
2), which represents a direct Z-scheme transfer path (indicated by the green line with double-headed arrows). ②–③ Photogenerated electrons at the CBM of Sc
2CF
2 migrate to the CBM of Sc
2CCl
2 (or Sc
2CBr
2), while photogenerated holes at the VBM of Sc
2CCl
2 (or Sc
2CBr
2) migrate to the VBM of Sc
2CF
2. This migration follows a traditional type-II path (indicated by gray lines with arrows). Electronic property analysis shows that the band alignments of the Sc
2CF
2/Sc
2CX
2 (X = Cl, Br) heterostructure are made up of the CBM of the Sc
2CCl
2 (or Sc
2CBr
2) layer and the VBM of the Sc
2CF
2 layer. Compared to the band gap of two monolayers, the heterostructure has a smaller band gap (
Figure 5a), indicating a higher rate of photogenerated electron–hole pair recombination at the interface compared to the rate of intralayer recombination. Meanwhile, due to the built-in electric field from Sc
2CF
2 to Sc
2CCl
2 (or Sc
2CBr
2), the recombination of photogenerated electrons in the CBM of Sc
2CCl
2 (or Sc
2CBr
2) and photogenerated holes in the VBM of Sc
2CF
2 is accelerated, promoting the recombination of path ① carriers. In addition, electrons have varying additional potential energies at different points in the space charge region, a phenomenon known as energy band bending [
54]. The positive charge on the Sc
2CCl
2 (or Sc
2CBr
2) is repelled by the holes on the Sc
2CF
2, causing the energy band to bend downward. Correspondingly, as the electrons move, the energy bands of the Sc
2CF
2 bend upward, forming a potential barrier at the interface. Due to the presence of built-in electric fields and potential barriers, the transfer of electrons from the CBM of Sc
2CF
2 to the CBM of Sc
2CCl
2 (or Sc
2CBr
2), as well as the transfer of holes from the VBM of Sc
2CCl
2 (or Sc
2CBr
2) to the VBM of Sc
2CF
2, are suppressed. Therefore, electron transfer in paths ② and ③ is repressed. After absorbing photon energy, the electrons are excited to the CBM, while the holes remain in the VBM. Due to the obstruction of path ② and path ③, photogenerated electrons gather in the CBM of Sc
2CF
2, while photogenerated holes gather in the VBM of Sc
2CCl
2 (or Sc
2CBr
2), which facilitates the efficient separation of photogenerated carriers and prolongs their lifetime [
23]. Therefore, it is difficult for electrons and holes to transfer following the type-II pathway, and the Sc
2CF
2/Sc
2CX
2 heterostructure should be used as the photocatalyst for the Z-scheme. According to the above analysis, the Sc
2CF
2 layer exhibits a higher reduction ability. Photogenerated electrons and hydrogen ions undergo a reduction reaction on the CBM of the Sc
2CF
2 layer to produce hydrogen. Meanwhile, in the highly oxidizing Sc
2CCl
2 (or Sc
2CBr
2) layer, the photogenerated holes on the VBM react with hydroxyl groups to produce oxygen, thereby improving the photocatalytic performance.
Differently, the Sc
2CCl
2/Sc
2CBr
2 heterostructure is not suitable as a photocatalyst due to the fact that the CBM is lower than the energy level of E
H+/H2 (
Figure 5a). However, they can function as absorption layers for solar cells. As shown in
Figure 5c, the conduction band offset and valence band offset between the Sc
2CCl
2 and Sc
2CBr
2 layers are 0.11 eV and 0.40 eV, respectively. Therefore, under the influence of valence band offset, the photogenerated holes in the Sc
2CCl
2 layer tend to jump to the VBM of the Sc
2CBr
2 layer. Simultaneously, due to the lower CBM energy of Sc
2CCl
2 in the Sc
2CCl
2/Sc
2CBr
2 heterostructure, photogenerated electrons tend to move to the CBM of Sc
2CCl
2, resulting in a type-II band alignment. The very small conduction band offset can improve the energy conversion efficiency of the solar cell, while the large valence band offset limits the electrons in the Sc
2CCl
2 monolayer and the holes in the Sc
2CBr
2 monolayer [
55]. Therefore, the rate of electron hole recombination will decrease, and the lifetime of photogenerated carriers will be extended. This will promote the formation of indirect excitons, which can be utilized in optoelectronic devices.
Considering that the construction of vdW heterostructures is an effective approach to enhance optical absorption and achieve excellent photovoltaic performance, therefore, to analyze the optical properties of the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, we calculated the optical absorption of the Sc
2CF
2 monolayer, Sc
2CCl
2 monolayer, Sc
2CBr
2 monolayer, and the Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures, as shown in
Figure 6a–c. Among them, the optical absorption coefficient is determined by the following equation [
56]:
where
and
represent the real and imaginary parts of the complex dielectric function
, respectively. As illustrated in
Figure 6a, we found that, compared to monolayers Sc
2CF
2 and Sc
2CCl
2, the Sc
2CF
2/Sc
2CCl
2 heterostructure has a wide absorption range from UV light to visible light due to its reduced band gap. It can be seen that the optical absorption coefficient of the Sc
2CF
2/Sc
2CCl
2 heterostructure is much larger than that of Sc
2CF
2 and Sc
2CCl
2 in both the UV and visible light ranges. More importantly, the Sc
2CF
2/Sc
2CCl
2 heterostructure shows a high absorption coefficient in the visible light region, reaching up to 2.53 × 10
5 cm
−1 at a wavelength of 410 nm. The enhancement of the optical absorption coefficient is mainly due to the interlayer coupling between two monolayers of the Sc
2CF
2/Sc
2CCl
2 heterostructure [
57]. Therefore, it is expected that the Sc
2CF
2/Sc
2CCl
2 heterostructure can act as an efficient visible light-harvesting photocatalyst. As can be seen from
Figure 6b, compared with the Sc
2CF
2 monolayer and Sc
2CBr
2 monolayer, the significantly increased optical absorption in the UV and visible light regions of the Sc
2CF
2/Sc
2CBr
2 heterostructure is due to the interlayer coupling [
58]. At the same time, compared with the Sc
2CF
2 monolayer and Sc
2CBr
2 monolayer, the increase in the optical absorption range of the Sc
2CF
2/Sc
2CBr
2 heterostructure is on account of the decrease in the band gap. Therefore, compared to the Sc
2CF
2 and Sc
2CBr
2 monolayers, the Sc
2CF
2/Sc
2CBr
2 heterostructure exhibits superior optical absorption performance, enabling efficient solar energy harvesting.
Light-absorbing materials not only need to have a suitable electronic structure but also need to have the ability to harvest solar light. Therefore, it is of great significance to study the optical properties of the Sc
2CCl
2/Sc
2CBr
2 heterostructure. The calculated absorption spectra of Sc
2CCl
2 and Sc
2CBr
2 monolayers, as well as the Sc
2CCl
2/Sc
2CBr
2 heterostructure, are shown in
Figure 6c. In the UV and visible regions, the absorption intensity of Sc
2CCl
2 and Sc
2CBr
2 monolayers is weak. However, the absorption peak of the Sc
2CCl
2/Sc
2CBr
2 heterostructure in the visible region is nearly 2.33 × 10
5 cm
−1, which is 1.71 times that of the Sc
2CCl
2 monolayer. The enhancement of the optical absorption coefficient is mainly due to the interlayer coupling between two monolayers of the Sc
2CCl
2/Sc
2CBr
2 heterostructure [
58]. Compared to both monolayers, the absorption range of the Sc
2CCl
2/Sc
2CBr
2 heterostructure increases due to its reduced band gap. Therefore, it can be concluded that the Sc
2CCl
2/Sc
2CBr
2 heterostructure would be a promising material for solar cells.
For device applications, in addition to the electronic and optical properties of the Sc
2CCl
2/Sc
2CBr
2 heterostructure analyzed above, such as limited band gaps, strong solar light-harvesting capabilities, and easy separation of electrons and holes with type-II band alignment, the ability to convert photon energy into electricity is also critical for solar cell applications. We use the method developed by Scharber et al. to calculate the PCE of solar cells, and its formula is as follows [
59]:
where 0.65 represents the band fill factor,
stands for the optical band gap of the donor, and
represents the conduction band offset (CBO). The open circuit voltage is
, and
is the 1.5 AM solar energy flux at the photon energy
. As shown in
Figure 6d, the calculated PCE of the Sc
2CCl
2/Sc
2CBr
2 heterostructure is about 20.78% (highlighted by the red star), which surpasses that of many other heterostructures, such as GeSe/AsP (16%) [
13], InS/InSe (13.17%) [
60], Hf
2CO
2/MoS
2 (19.75%) [
29], and MoS
2/BP (20.42%) [
61] heterostructures (highlighted by the green circle). Thus, we conclude that the Sc
2CCl
2/Sc
2CBr
2 heterostructure is more promising and competitive for 2D vdW heterostructure solar cells.
Strain engineering is an effective method to change the structural, electronic, and magnetic properties of 2D materials [
62]. In addition, strain is unavoidable in industrial production, which comes from bending, external loads, and lattice mismatch [
46]. Applying a biaxial strain will alter the band structure of the heterostructure and affect its photocatalytic and photovoltaic performance [
42,
45]. Then, the effects of in-plane biaxial strain on the electronic properties of Sc
2CF
2/Sc
2CCl
2, Sc
2CF
2/Sc
2CBr
2, and Sc
2CCl
2/Sc
2CBr
2 heterostructures are systematically studied. Here, the inner-layer biaxial strain (ε
in) is defined by ε
in = [(
L −
L0)/
L0] × 100%, where
L and
L0 are the lattice constants before and after the strain application, respectively. The applied strains η are −8%, −6%, −4%, −2%, 2%, 4%, 6%, and 8%, respectively. A negative value of η means that compressive strain is applied to the heterostructure. When η is positive, it indicates that tensile strain is applied to the heterostructure.
As shown in
Figure S4, the electronic properties of the Sc
2CF
2/Sc
2CCl
2 heterostructure are significantly changed by applying biaxial strain. Compared with the Sc
2CF
2/Sc
2CCl
2 heterostructure without strain (
Figure 3a), the applied strain changes the band gap of the heterostructure. From
Figure 7a, it can be seen that when the compressive strain is −8%, −6%, −4%, and −2%, the band gap of the heterostructure decreases to 0.36, 0.39 eV, 0.46 eV, and 0.52 eV, respectively. Among them, the positions of the CBM and VBM have not changed and are still located at the high symmetry points M and Γ, as shown in
Figure S4a–d. When the tensile strains are +2%, +4%, +6%, and +8%, respectively, the band gaps of the heterostructure increase to 0.62 eV, 0.66 eV, 0.72 eV, and 0.75 eV, respectively. The positions of the CBM and VBM are still located at the high symmetry points M and Γ, respectively (
Figure S4e–h). With the increase in strain, the CBM of the Sc
2CCl
2 monolayer gradually moves away from the Fermi level, causing an increase in the band gaps. It can be seen from
Figure 7b that the Sc
2CF
2/Sc
2CCl
2 heterostructure maintains a type-II band alignment throughout the strain. As for the band edge, all the heterostructures maintained photocatalytic activity under strain.
The electronic properties of the Sc
2CF
2/Sc
2CBr
2 heterostructure changed significantly when biaxial strain was applied, as shown in
Figure S5. In contrast to the strain-free Sc
2CF
2/Sc
2CBr
2 heterostructure (
Figure 3b), applying strain not only alters the band gaps of the heterostructure but also changes the band alignment of the heterostructure. As can be seen from
Figure 7a, the band gaps of the Sc
2CF
2/Sc
2CBr
2 heterostructure decrease to 0.31 eV, 0.66 eV, and 0.77 eV when the compression strain is −6%, −4%, and −2%, with the CBM and VBM located at highly symmetric points M and Γ (
Figure S5b–d). However, when the compressive strain increases to −8%, the band gap of the Sc
2CF
2/Sc
2CBr
2 heterostructure decreases to 0 eV. This indicates that the heterostructure transitions from an indirect band gap semiconductor to a metal under −8% compressive strain because the CBM (VBM) moves below (above) the Fermi level, as shown in
Figure S5a. When the tensile strain was +2%, +4%, +6%, and +8%, the CBM and VBM were located at highly symmetric points M and Γ, with band gaps increasing to 0.79 eV, 0.80 eV, 0.81 eV, and 0.82 eV, respectively, as shown in
Figure S5e–h. From
Figure 7c, when the compressive strain is between −6% and −4%, the VBM of the Sc
2CBr
2 layer is positioned at a higher energy level than that of the VBM of the Sc
2CF
2 layer. Consequently, the VBM of the Sc
2CF
2/Sc
2CBr
2 heterostructure shifts from the Sc
2CF
2 layer to the Sc
2CBr
2 layer, leading to a transition from type-II to type-I. In addition, when the compression strain is −2%, the VBM of Sc
2CBr
2 is higher than that of E
O2/H2O, which is unfavorable for the photocatalytic reaction. By analyzing the band structure of the Sc
2CF
2/Sc
2CBr
2 heterostructure under strain, it is considered that the strain affects the relative position of atoms as well as the bonding properties and strength of the atoms, leading to a change in the band structure. The band alignment of the Sc
2CF
2/Sc
2CBr
2 heterostructure can be changed from type-I to type-II under different strain conditions.
For the Sc
2CCl
2/Sc
2CBr
2 heterostructure, the applied biaxial strain range is still −8%~8%. As shown in
Figure S6, it is noteworthy that under −8%~6% biaxial strains, the heterostructures consistently maintain type-II banding and retain indirect band gap characteristics. As displayed in
Figure 7a, the band gaps of the Sc
2CCl
2/Sc
2CBr
2 heterostructure decrease to 0.17 eV, 0.55 eV, 0.88 eV, and 1.14 eV when compressive strain is applied. As the tensile strain increases, the band gap also increases, reaching 1.53 eV, 1.67 eV, 1.79 eV, and 1.88 eV, respectively. Under −8%~6% biaxial strains, the CBM and VBM are still contributed by Sc
2CCl
2 and Sc
2CBr
2, located at the M and Γ points, respectively, as depicted in
Figure 7d. Unlike these changes, under the tensile strain of 8%, the CBM of the Sc
2CCl
2 layer becomes higher than the CBM of the Sc
2CBr
2 layer. Thus, the CBM of the Sc
2CCl
2/Sc
2CBr
2 heterostructure shifts from the Sc
2CCl
2 layer to the Sc
2CBr
2 layer, leading to a type-II to type-I transformation. In addition, we calculated the PCE values of the Sc
2CCl
2/Sc
2CBr
2 heterostructure under various biaxial strains, as illustrated in
Figure 6d (highlighted by the black star). From
Figure S7, we can see that a maximum PCE of 20.07% can be achieved under 2% tensile strain.