Xinli
Wang
ab,
Juping
Xu
ab,
Peng-Fei
Liu
ab,
Bao-Tian
Wang
ab and
Wen
Yin
*abc
aInstitute of High Energy Physics, Chinese Academy of Sciences (CAS), Beijing 100049, People's Republic of China
bSpallation Neutron Source Science Center, Dongguan 523803, People's Republic of China
cUniversity of Chinese Academy of Sciences, Beijing 100049, People's Republic of China. E-mail: yinwen@ihep.ac.cn
First published on 20th April 2023
It is crucial to understand the electronic properties of two-dimensional (2D) semiconductor heterostructures for better application in photocatalyst and nano-electronic devices. In this work, using first-principles calculations, taking quintuple-layer (QL) Al2O3 and a Janus MoSO monolayer with out-of-plane polarization as an example, we systematically study the electronic properties of QL-Al2O3/MoSO heterostructures. By changing the different polarization direction arrangements of QL-Al2O3/MoSO heterostructures, we find that the evolution of band alignment, spatial charge distribution and interface charge transfer is synergetic. Three parts are included: first, the tuning of band alignment corresponds to the tuning of the surface charge distribution of heterostructures. Second, the charge redistribution of two monolayers corresponds to the interface charge transfer of heterostructures. The charge transfer process contains the interlayer charge transfer and inner-layer charge transfer between two monolayers. Third, the unidirectional charge transfer process through the interface is driven by the inner polarization electric field. Our work not only clarifies the interface charge transfer mechanism that can be applicable to other 2D non-polar and polar heterostructures, but also provides a theoretical basis for the application of heterostructures.
It is necessary to understand the electronic properties (band structures, charge distribution, and interlayer charge transfer) of these 2D semiconductor heterostructures for their application in catalysts and electric devices. The band structures of 2D heterostructures, containing the component of each layer, bring out novel electric characteristics because of the different band alignments. According to the band alignment, the 2D semiconductor heterostructures are usually divided into type-I, type-II and type-III types.23 At present, most research studies focus on tuning the electronic properties and application design based on the band structures of 2D heterostructures as shown in Fig. 1. Understanding the basic charge distribution and redistribution (under extra stimulation) of 2D semiconductor heterostructures is favorable for their application, since the different band alignments correspond to different charge distributions of 2D heterostructures.
Taking 2D semiconductor heterostructures with type-II band alignment as an example, it is interesting to find that the band alignment of type-II and direct Z-scheme heterostructure photocatalysts is the same, while the interfacial charge (photo-excited) transfer direction is opposite as shown in Fig. 1(c) and (d). Actually, the interface charge transfer of 2D heterostructures, which is accompanied by the charge redistribution of two single layers, contains two stages during and after the formation of a heterojunction: the first stage of interface charge transfer occurs in the process of two layers connecting together. The second stage occurs when heterostructures receive stimulations, such as light-irradiation, an applied electric field and stress. Understanding the driving force of interface charge transfer is the key to the interface charge transfer process.
In the process of heterostructure formation, the driving force of interface charge transfer (first stage) is the difference in the work function. The electron is transferred from the A layer with a lower work function to the B layer with a higher work function, which forms the spatially separated charge distribution of the heterojunction in Fig. 1(a). The type-II band alignment means the interface charge transfer process (first stage) is finished. The second stage of the interface charge transfer process of 2D heterostructures corresponds to the charge redistribution of two monolayers under external stimulation (light irradiation, extra electric field and stress). In general, the interface built-in electric field in Z-scheme heterostructures photocatalysts (not in type-II heterostructures photocatalysts), drives the recombination of photo-generated carriers (electrons and holes) by transferring the interface.65–86
However, the charge density difference results, for analyzing interface charge transfer (in the first stage) of 2D type-II or Z-scheme heterostructures, show that both the monolayers gain and lose electrons from the interface, respectively. Usually, the situation is that one layer gets electrons and another layer loses electrons. Previous research studies24–86 on 2D vdWs semiconductor heterostructures with type-II band alignment are listed in Table 1. Some of them have been proved theoretically to be Z-scheme photocatalysts for water-splitting. To fully understand interface charge transfer, the following questions should be reconsidered: (1) what is the driving force of interface charge transfer? (2) What is the direction of interface charge transfer? (3) What is the number of interfacial charge transfers? (4) What is the subsequent behavior of the charge transferred at the interface?
Type-II | Z-Scheme | ||||
---|---|---|---|---|---|
Non-polar/non-polar | InSe/As24 | InSe/InSb29 | InSe/MoSe2 (ref. 33) | PtS2/As65 | MoSe2/SnS2 (ref. 69) |
InSe/Sb25,26 | InS/GaN30 | InSe/MoS2 (ref. 34) | HfS2/As66 | ||
InSe/AlN27 | BSe/AlN31 | InSe/GaTe35 | ZrS2/As67 | MoSe2/PtO2 (ref. 70) | |
InSe/SiC28 | GaS/As32 | Sb/AlAs36 | HfS2/β-GeSe68 | MoTe2/CrS2 (ref. 71) | |
MoS2/As37 | MoSe2/PtS2 (ref. 40) | PtSe2/GaN43 | WSe2/Ti2CO2 (ref. 72) | BSe/GeC75 | |
MoSe2/Blue P38 | MoS2/WS2 (ref. 41) | WTe2/InSe73 | C3B/C3N76 | ||
MoS2/Black P39 | MoSe2/WSe2 (ref. 42) | WSe2/BP44 | MoTe2/BAs74 | BCN/C2N77 | |
Non-polar/polar | In2Se3/Bi2Se3 (ref. 45) | GaN/MoS2 (ref. 50) | MoSSe/Blue P54 | GeC/SnSSe78 | SnS2/CdS82 |
In2Se3/InTe46 | GaN/MoSe2 (ref. 51) | MoSSe/Te55 | CdO/PtSSe79 | InSe/CdS83 | |
In2Se3/MoS2 (ref. 47 and 48) | GaN/MgI2 (ref. 52) | In2STe/InSe56 | HfS2/MoSSe80 | ||
Al2Se3/Mo2CF2 (ref. 49) | Ga2O3/MoS2 (ref. 53) | InS/SeGa2Te57 | BCN/In2Se3 (ref. 81) | C7N6/GaSnPS84 | |
Polar/polar | In2Se3/In2Se3 (58 and 59) | In2SSe/In2SSe61 | PtSeTe/LiGaS2 (ref. 63) | MoSSe/WSeTe85 | TeIn2S/TeIn2Se86 |
Al2O3/Al2O3 (ref. 60) | InGaSTe/InGaSTe62 | WSSe/In2Se3 (ref. 64) | In2Te3/Ga2Te3 (ref. 86) |
Actually, the charge redistribution or interface charge transfer (in the second stage) of 2D heterostructures under extra electric field and stress has been researched.24–29 Under increasing extra electric field, more electrons flow from one layer to another layer. In addition, by changing the direction of the extra electric field, the direction of interface transfer charge also reverses. Apart from the applied extra electric field and interface built-in electric field, there is another polarization electric field, widely present in 2D polar materials, which could drive charge transfer through the interface.60 According to the intrinsic out-of-plane polarization of 2D monolayers, the 2D vdWs semiconductor heterostructures can also be classified into three kinds: non-polar/non-polar heterostructure, non-polar/polar heterostructure, and polar/polar heterostructure (Table 1). For example, previous results reveal that the intrinsic ferroelectricity polarization in α-In2Se3 can dramatically tune the electronic properties of In2Se3/MoS2 heterostructures,47 while the effect of polarization on interface charge transfer (in the first stage) is not clearly explained.
In this work, we take two representative 2D polar materials of monolayer MoSO and quintuple layer (QL) Al2O3 as examples, which are promising piezoelectric materials87 and ferroelectric (FE) tunnel barriers.88 By analyzing the connection of band alignment, surface charge distribution and interface charge transfer in QL-Al2O3/MoSO heterostructures with different polarization directions, we reveal the effect of the polarization electric field on interface charge transfer and charge distribution of polarized 2D heterostructures. These unique characteristics can be suitable for other 2D polarized materials.
The Janus transition metal dichalcogenide (TMD) monolayers, MXY (M = Mo, W; X, Y= O, S, Se, Te, X ≠ Y), have been proposed as efficient photocatalysts for water splitting.103,104 Different from the pristine MX2 monolayer, there is an intrinsic dipole in Janus MXY due to the different electronegativities of X and Y atoms which will separate the photo-generated electrons and holes.87,103–109 At present, ultrafast charge transfer in Janus MoSSe/MoS2 heterostructures has been observed.110 However, studying the native mechanism of interlayer charge transfer in 2D polarized materials is still rare. It is crucial for designing semiconductor devices and catalysts based on 2D vdWs heterostructures. It is worth mentioning that the α-Al2O3 (0001) surface is an ideal substrate for growing single crystal MoS2 layers because of the matching lattices.111 In addition, Miao et al. successfully synthesized a MoS2−xOx layer by oxidizing MoS2 layers.112 Thus, in this work, we take QL-Al2O3/MoSO heterostructures as an example to reveal the intrinsic mechanism of interlayer charge transfer. The QL-Al2O3/MoSO heterostructures are divided into two categories depending on the octahedral Al termination (Aloct) or tetrahedral Al (Altet) termination of QL-Al2O3 connected with the MoSO monolayer, as shown in Fig. 2. In each category, there are two cases, where the S or O atom of MoSO is connected with the QL-Al2O3 monolayer, respectively.
The first principles calculations for calculating the geometric structures, stability, and electronic structures of QL-Al2O3 and MoSO monolayers are performed using VASP (Vienna ab initio simulation package) software package113 based on density functional theory.114 The projector-augmented wave approach115 was used for describing the interaction of the electron and atom core. The exchange correlation potential is the PBE potential in the generalized gradient approximation (GGA).116 The electronic structures are also calculated using the HSE06 (ref. 117) functional, with the mixing parameter for the Hartree–Fock potential set to 0.25. A vacuum layer approximately 15 Å was applied to cancel the interaction between layers. The cutoff energy is set as 500 eV. The convergence criteria for the Hellmann–Feynman force and energy were less than 0.001 eV Å−1 and 10−5 eV, respectively. A 5 × 5 × 1 gamma centered k-point mesh was used for the geometry relaxations and 10 × 10 × 1 gamma centered k-point mesh was used for electronic property calculations.
Phonon spectra were calculated on the basis of the density functional perturbation theory method by using the Phonopy program.118 The thermodynamic stability of free-standing QL-Al2S3 and Al2S(O)3 structures is judged based on the ab initio molecular dynamic simulation (AIMD) method119 at room temperature for 10 ps. The electronic properties were analyzed with the VASPKIT120 package. The atomic configurations and charge density difference were visualized by using the VESTA121 package.
Monolayer | a = b (Å) | d1 (Å) | d2 (Å) | d3 (Å) | d4 (Å) | Thickness (Å) |
---|---|---|---|---|---|---|
Al2O3 | 2.930 | 1.886 | 1.701 | 2.063 | 1.897 | 4.476 |
MoSO | 3.030 | 2.380 | 2.090 | \ | \ | 2.791 |
Structurally, for the MoSO monolayer, the bond length between Mo–O (dMo–O) is shorter than that of the Mo–S (dMo–S) bond. The shorter bond length of Mo–O than that of Mo–S corresponds to the stronger binding energy of Mo–O than that of Mo–S and the higher electronegativity of O atoms than that of S atoms.87,122
The phonon spectra and corresponding phonon density of states (PHDOS) are calculated using the finite displacement method with a supercell to ensure the dynamical stability of QL-Al2O3 and Janus MoSO monolayers. For QL-Al2O3 and MoSO monolayers, no negative frequency phonon can be observed, thus confirming the dynamical stability of the QL-Al2O3 and MoSO monolayers as shown in Fig. 3(a) and (c). The thermal stability of QL-Al2O3 and MoSO is explored via AIMD simulations. The simulations lasted for 10 ps in the canonical ensemble controlled by a Nose–Hoover thermostat (NVT)123 with a time step of 2 fs at 300 K. After 10 ps simulation, no distinct structural destruction is observed, as shown in Fig. 3(b) and (d), and the total energy fluctuation is small. These findings indicate that the QL-Al2O3 and MoSO monolayers are thermally stable at 300 K.
For exploring the atomic contributions to the bands, we study the weighted band-structure and layer-resolved projected density of states (PDOS) of QL-Al2O3 and Janus MoSO monolayers, as shown in Fig. 5. Following this, we only show the DFT results by PBE calculations; the results by HSE06 calculations show the same trend and are not shown here.
The weighted band-structure of QL-Al2O3 has shown that the valence bands are greatly contributed by the O atoms (electron acceptor, below the Fermi level (EF)), while the conduction bands are greatly contributed by the Al atoms (electron donor, above EF). The layer-resolved PDOS of QL-Al2O3 has shown that the CBM is mainly contributed by the O–Altet atoms from the tetrahedrally terminated O–Al bilayer in Fig. 5(b), while the VBM is mainly contributed by the O–Aloct atoms from the octahedrally terminated O–Al–O trilayer. Similarly, the weighted band-structure and layer-resolved PDOS of the Janus MoSO monolayer have shown that the CBM is greatly contributed by the Mo–S atoms, while the VBM is greatly contributed by the Mo–O atoms.
The planar-averaged electrostatic potential distribution of QL-Al2O3 and MoSO monolayers is shown in Fig. 5(c) and (f). The electrostatic potential of Aloct termination is higher than that of Altet termination in QL-Al2O3, and the potential difference (Δϕ) across the monolayer is 3.978 eV. The electrostatic potential of O termination is 1.951 eV higher than that of S termination in Janus MoSO monolayer, as shown in Fig. 5(f).
The electrostatic potential difference (Δϕ) across the QL-Al2O3 and MoSO monolayers results from the structural asymmetry in QL-Al2O3 and the electronegativity difference of O and S atoms in the MoSO monolayer,87,99 respectively. It is necessary to point out that the higher electron potential termination is equivalent to the center of negative effective charge. Similarity the lower potential termination is equivalent to the center of positive effective charge. The schematic diagram inset in Fig. 5(c) and (f) illustrates the separated surface charge distribution in QL-Al2O3 and MoSO monolayers. The surface electrons and holes (green and yellow) reflect the CBM and VBM, which come from two separate surfaces in the layer-resolved PDOS. The Δϕ corresponds to the intrinsic polarization (bound) charge on the two separated sides.
When the S atom of MoSO connects with Altet termination of QL-Al2O3, the CBM and VBM of the QL-Al2O3/MoSO heterostructure are located at the K-point. The band-gap of the QL-Al2O3/MoSO heterostructure is calculated to be 2.615 (4.179) eV at the PBE (HSE06), respectively, as shown in Fig. 6(c). When the O atom of MoSO connects with Altet termination of QL-Al2O3, both the CBM and VBM (near the K-point) cross the EF by PBE calculations as shown in Fig. 6(d). As a comparison, the Altet–O heterostructure is the most energy-favorable configuration, as shown in Fig. 2. The energy of Altet–S, Aloct–O, and Aloct–S heterostructure configurations is higher by 0.053, 0.065, and 0.061 eV per unit than that of the Altet–O heterostructure. Although the QL-Al2O3 and Janus MoSO monolayers are indirect bandgap semiconductors, it is interesting to find that QL-Al2O3/MoSO heterostructures are a direct band gap semiconductor or metal depending on the stacking method. Why do the band-structures in QL-Al2O3/MoSO heterostructures change with different stacking models? Next, the detailed electric properties of four heterostructures are analyzed for explaining the reason based on the PBE level.
When the Aloct termination of QL-Al2O3 connects with the MoSO surface (S or O) atom, the weighted band-structure and layer-resolved PDOS of two kinds of QL-Al2O3/MoSO heterostructures are compared with those of single QL-Al2O3 and MoSO monolayers, and the CBM from QL-Al2O3 in two QL-Al2O3/MoSO heterostructures comes near the EF, even crossing the EF when the Aloct termination connects with O and S atoms of the MoSO surface, as shown in Fig. 7(a), (b), (d) and (e), respectively. From band alignment, the CBM in QL-Al2O3 means that it got electrons from the MoSO monolayer when a heterostructure was formed. On comparing the different energy level positions of the CBM (VBM) in two QL-Al2O3/MoSO heterostructures, we can assume that there are more electrons (holes) transferred from MoSO to QL-Al2O3 when QL-Al2O3 connects with S rather than O atoms of the MoSO surface.
The interface charge transfer in two QL-Al2O3/MoSO heterostructures in Fig. 7(g) and (i) is obviously different: (1) the interfacial transferred charge is larger in the second heterostructure (Fig. 7(i)) than in the first heterostructure (Fig. 7(g)). (2) The electrons are transferred unidirectionally from MoSO to the QL-Al2O3 layer through the interface just in the second heterostructure, while two sets of interface charge transfer are present in the first heterostructure. (3) The subsequent behavior of transferred charge at the interface is different. The transferred charge spreads over the whole heterostructure and is not just located at the interface in the second heterostructure. In the first heterostructure, interface transferred charge just stays on S termination of MoSO and O termination of QL-Al2O3 at the interface. (4) From the above three differences, the driving force of interface charge transfer is different in two heterostructures. Next, we focus on the driving force of interface charge transfer.
The polarization direction arrangement of two QL-Al2O3/MoSO heterostructure configurations is different. When the O atom of MoSO connected with the Aloct termination, the intrinsic polarization direction inside QL-Al2O3 and MoSO is opposite. When the S atom of MoSO connected with the Aloct termination, the intrinsic polarization direction inside QL-Al2O3 and MoSO is the same. Compared with the polarization arrangement, the total electrostatic potential distribution shows two different situations. From the leftmost to the rightmost surface, the total potential across the QL-Al2O3/MoSO heterostructure increases first and then decreases, as shown in Fig. 7(c), while the total potential across the QL-Al2O3/MoSO heterostructure decreases step-like gradually, as shown in Fig. 7(f).
The surface charge distribution of the heterostructure is verified from the layer-resolved PDOS (induced charge) and the electrostatic potential distribution (bound charge). When the S atom of MoSO connects with the Aloct termination, the electrons and the holes are located at the rightmost and leftmost surfaces. In this configuration, the negative and positive polarization (bound) charges are accompanied by the holes and electrons as shown in Fig. 7(j). How is the surface charge distribution formed? The existence of a potential difference (Δϕ) represents the polarization strength or the number of polarization charges, which provides the driving force of interface charge transfer between QL-Al2O3 and MoSO layers. In addition, the interfacial transferred electron (or hole) will move to the outside surfaces driven by the inner polarization electric field in this heterostructure configuration. The electrons flow unidirectionally from a higher electrostatic potential to a lower electrostatic potential through the whole QL-Al2O3/MoSO heterostructure. When the O atom of MoSO connects with the Aloct termination, since the inner polarization electric fields in QL-Al2O3 and MoSO are opposite, both monolayers want to get electrons (holes) from another monolayer, so there is little electron (hole) flow through the interface. Since the highest electrostatic potential is located at the interface, there are double-directional electron (or holes) transfer processes through the interface.
The polarization electric field across two monolayers in the second heterostructure (Fig. 7(i) and (j)) provides the driving force for unidirectional charge transfer, while in the first heterostructure (Fig. 7(g)and (h)), there is no polarization electric field across the two monolayers. The uneven distribution of free charge (hole) number or polarization (bound) charge on both sides of the interface leads to the existence of an interface electric field. Corresponding, the direction of interface electric field is from MoSO to QL-Al2O3, since the total polarization electric field point from QL-Al2O3 to MoSO. The interface electric field in the polarization same-direction arrangement heterostructure is eliminated by the inner polarization electric field.
The changing band alignment in two heterostructures means different surface charge distributions and interface charge transfers. As shown in Fig. 8(c) and (g), the interface charge transfer in two heterostructures is surely different: (1) the amount of interfacial transferred charge is lower in the first heterostructure (Fig. 8(g)) than in the second heterostructure (Fig. 8(i)). (2) The electrons are transferred unidirectionally from QL-Al2O3 to MoSO through the interface in the second heterostructure, and not in the first heterostructure. (3) The transferred charge spreads over the whole heterostructure and is not just located at the interface S termination of MoSO and O termination of QL-Al2O3 in the second heterostructure. The interfacial transferred charge is just located at the interface in the first heterostructure.
The polarization direction arrangement and the total electrostatic potential distribution of two QL-Al2O3/MoSO heterostructure configurations are different. From the leftmost to the rightmost surface, the total potential across the QL-Al2O3/MoSO heterostructure decreases first and then increases as shown in Fig. 8(c), while the total potential across the QL-Al2O3/MoSO heterostructure decreases gradually as shown in Fig. 8(f).
When the O atom of MoSO connects with the Altet termination, the electron (hole) and positive (negative) bound charge are located at the rightmost (leftmost) surface as shown in Fig. 8(i) and (j). The existence of Δϕ represents the polarization electric field, which provides the driving force of unidirectional charge transfer. When the S atom of MoSO connects with the Altet termination, the highest and lowest electrostatic potential distribution is located on the leftmost surface of QL-Al2O3 and the interface between QL-Al2O3 and MoSO monolayers. The electrons flow from a high electrostatic potential to a low electrostatic potential in this heterostructure, while the transferred charge through the interface is still low, because of the existence of an interface electric field. Corresponding, the direction of interface electric field is from QL-Al2O3 to MoSO, since the total polarization electric field point from MoSO to QL-Al2O3. The interface electric field in the polarization same-direction arrangement heterostructure is eliminated by the inner polarization electric field.
In summary, there are three types of interfacial charge transfer processes with different polarization arrangements in QL-Al2O3/MoSO heterostructures. The polarization direction arrangement and electrostatic potential distribution decide the direction of charge transfer, which will induce different surface charge distributions and band-structures of heterostructures. For the same-direction polarization heterojunction, the potential distribution is monotonically decreased, and the polarization electric field drives unidirectional transferred charges to accumulate at the surface atoms, which result in energy level shifting and the presence of metallicity in the QL-Al2O3/MoSO heterostructure. For the opposite-direction polarization heterojunction, the potential distribution is not monotonically decreased, and the interface transferred charge is low and trapped at the interface, since there is no total electric field through the heterostructure. Correspondingly, the charge distribution of two monolayer surfaces changes a little, and the electric properties keep the semiconductor in this QL-Al2O3/MoSO heterostructure.
The Aloct termination of QL-Al2O3 has connected with the MoSO surface (S or O) atom. The band structures and layer-resolved PDOS of two QLs-Al2O3/2MoSO heterostructures have shown metallicity. When the O atom of MoSO connects with QL-Al2O3, the layer-resolved PDOS clearly shows that both the VBM (from the middle MoSO layer) and CBM (from the rightmost MoSO layer) cross the EF as shown in Fig. 9(b). When the S atom of MoSO connects with QL-Al2O3, the layer-resolved PDOS clearly shows that both the VBM (from the leftmost MoSO layer) and CBM (from the rightmost QL-Al2O3) cross the EF as shown in Fig. 9(e).
The different surface charge distributions of two QLs-Al2O3/2MoSO heterostructures correspond to different interface charge transfers between QLs-Al2O3 and two MoSO monolayers. For verifying the interfacial charge transfer process, two kinds of definitions of charge density difference of QL-Al2O3/2MoSO heterostructures are shown in Fig. 9(g), (h) and (i), (j). First, two MoSO layers are defined as one unit.60 Second, two MoSO monolayers are defined as separated monolayers.
The interface charge density differences of two QL-Al2O3/2MoSO heterostructures are different: (1) the amount of interfacial transferred charge is low in the first heterostructure (Fig. 9(g)) compared to in the second heterostructure (Fig. 9(i)). (2) The electrons (or holes) are transferred unidirectionally from left MoSO to QL-Al2O3 in the second heterostructure, while not in the first heterostructure. (3) The interface transferred charge spreads over the whole heterostructure and is not just located at the interface atoms in the second heterostructure. Here, since two MoSO monolayers are one unit, we can clearly see the unidirectional transfer of electrons from the left interface (left-MoSO/middle-MoSO) to the right interface (middle-MoSO/QL-Al2O3) by comparing two kinds of charge density differences.
The polarization direction and electrostatic potential distribution across two QLs-Al2O3/2MoSO heterostructures in Fig. 9(c) and (f) are consistent with that of the two QLs-Al2O3/2MoSO heterostructures in Fig. 7(c) and (f). For the same-direction polarization arrangement QLs-Al2O3/2MoSO heterostructure, the total potential across the QL-Al2O3/2MoSO heterostructure decreases gradually from the leftmost MoSO to the rightmost QL-Al2O3 surface as shown in Fig. 9(c). The electrons flow unidirectionally from the high potential (left-MoSO) layer to the low potential (QL-Al2O3) layer. Correspondingly, metallicity in this QL-Al2O3/2MoSO heterostructure originates from the increased surface charge induced by unidirectional charge transfer. For opposite-direction polarization QLs-Al2O3/2MoSO heterostructures, the highest electron potential is located at the interface between QLs-Al2O3 and the middle-MoSO monolayer. The charge transfer through the QLs-Al2O3/middle-MoSO interface is low and double-directional, while the potential distribution across the MoSO bilayer plays a major role in determining the metallicity of this QL-Al2O3/2MoSO as shown in Fig. 9(c). The electrons flow unidirectionally from a higher electrostatic potential (middle-MoSO) to a lower electrostatic potential (left-MoSO) as shown in Fig. 9(h), resulting in the shifting of the CBM (VBM) to EF on the left (middle)-MoSO layer.
The Altet termination of QL-Al2O3 has connected with the MoSO surface (S or O) atom. The band-structure of the two QLs-Al2O3/2MoSO has shown metallicity as shown in Fig. 10(a) and (d). When the S atom of MoSO connects with QL-Al2O3, both the CBM (from the middle-MoSO layer) and VBM (from the right-MoSO layer) cross the EF as shown in Fig. 10(b). When the O atom of MoSO connects with QL-Al2O3, both the VBM (from the leftmost QL-Al2O3 layer) and CBM (from the right-MoSO layer) cross the EF as shown in Fig. 10(e).
When the S atom of MoSO connects with the Altet termination, the potential distribution in MoSO bilayers play a major role in determining the electric properties of QL-Al2O3/2MoSO. The metallicity originates from 2MoSO because of the unidirectional charge transfer between two MoSO monolayers as shown in Fig. 10(h). When the O atom of MoSO connects with the Altet termination, the intrinsic polarization direction inside QL-Al2O3 and 2MoSO is the same. From the leftmost QL-Al2O3 to the rightmost MoSO surface, the total potential across the QL-Al2O3/2MoSO heterostructure decreases gradually as shown in Fig. 10(f). The metallicity in this heterostructure originates directly from the increased surface charge, induced by unidirectional charge transfer through the heterostructure. The interface charge transfer process of QL-Al2O3/2MoSO heterostructures is schematically summarized in Fig. 11.
According to the existence of the polarization (bound) charge of monolayers, we classify the 2D vdWs heterostructures into three kinds. Correspondingly, the electrostatic potential distributions across the 2D heterostructures are shown in Fig. 12. For non-polar/non-polar heterostructure, there is no Δϕ across the 2D heterostructures. For a non-polar/polar heterostructure and polar/polar heterostructure, there is Δϕ across the 2D heterostructures. The electrostatic potential distribution across the heterostructure not only depicts the spatial charge distribution, but also the electric field distribution, which decides the interface charge transfer in the first and second stage.
For a non-polar/non-polar heterostructure, the zero Δϕ means the charge distribution of heterostructures is even after two layers are connected. So, there is no built-in electric field driving the second stage interface charge transfer. For a non-polar/polar heterostructure, the Δϕ is not zero. There are two types of interfacial charge transfer processes as shown in Fig. 12(b) and (c). The interfacial charge transfer direction can be judged by the electrostatic potential distribution, the electrons flow from a higher potential to a low potential, and the built-in polarization electric field provides the driving force for the charge transfer through the interface in the first stage.
For a polar/polar heterostructure, there are three types of interfacial charge transfer processes. When the polarization directions of two monolayers are the same as shown in Fig. 12(d), the spatial charge and electric field distribution are similar to that of a non-polar/polar heterostructure. When the polarization directions of the two monolayers are opposite as shown in Fig. 12(e) and (f), both the built-in polarization and interface electric field affect the interfacial charge transfer processes. There are two common rules from the above results. First, the total potential difference across the heterostructure depends on which side (higher or lower potential) of the polar material is connected to the added 2D material. Second, the total potential difference across the heterostructure obeys the polar monolayer with a larger Δϕ (polarization strength). Specifically, the Δϕ across QL-Al2O3 is larger than the Δϕ across the MoSO monolayer or bilayer. So, the electrostatic potential distribution of QL-Al2O3 plays a main role in deciding the interface charge transfer direction.
For type-II and Z-scheme heterostructures, the zero Δϕ means no polarization electric field exists, while the built-in interface electric field (the uneven charge distribution) exists in Z-scheme heterostructures. The difference between type-II and polar-PN heterostructures (same polarization direction arrangement) is the existence of a built-in polarization electric field, since the Δϕ means the spatially separated polarization charge distribution in the heterostructure. Similarly, the difference between Z-scheme and S-scheme heterostructures is the existence of a built-in polarization electric field. The difference between polar-PN and S-scheme heterostructures is the existence of a built-in interface electric field.
We also classify the potential distribution alignment in 2D vdWs heterostructures according to the intrinsic polarization of monolayers. The total potential difference across the polar heterostructures decides the charge transfer direction. Our work reveals the basic connection between the band alignment and charge distribution in 2D polarized heterostructure systems. The understanding of intriguing electric properties of 2D heterostructures promotes their applications in photocatalysts and nano-optoelectronic devices.
This journal is © The Royal Society of Chemistry 2023 |