Two-dimensional (2D) semiconductors have been explored as potential channel materials in future nanoscale field-effect transistors (FETs). However, searching for suitable gate dielectric materials interfaced with 2D semiconductor channels and controlling their quality to guarantee efficient gate role are critical and challenging in the fabrication of high-performance nanoscale FETs. In the present article, we adopt first-principles calculations to explore the binding energies, band structures, and electronic properties of heterojunctions between monolayer blue phosphorene (BlueP) semiconductor and dielectrics, including BlueP-BN, BlueP-HfO2, BlueP-TiO2, and BlueP-CaF2. For the first time, we deeply investigate the electronic properties of BlueP-dielectric heterojunctions under perpendicular external electric fields. Our calculated results indicate that HfO2 thin layer and monolayer CaF2 dielectrics are appropriate as gate dielectrics for BlueP-based FETs, and furthermore, monolayer CaF2 is superior to HfO2. We also investigate the electronic properties of BlueP-HfO2 with interfacial O-vacancy and BlueP-CaF2 with interfacial F-vacancy, as well as hydrogen passivation to the F-vacancy of BlueP-CaF2. Our results indicate that the interfacial atomic vacancies of dielectric layer greatly deteriorate its dielectric properties and have great impacts on the electrical properties of the whole heterojunction. Fortunately, hydrogen passivation to F-vacancy of BlueP-CaF2 can effectively protect the semiconductor properties of BlueP and the dielectric properties of CaF2. This implies that hydrogen passivation strategy can improve the performance of 2D semiconductor-based nanoelectronic devices with CaF2 as a gate dielectric, thus providing guidance for the design and optimization of future nanoscale FETs.

2D semiconductors with atomic size thickness, no hanging bonds, and excellent electronic properties (such as high carrier mobility and high on/off ratio) have been explored as potential channel materials to suppress short-channel effects of future nanoscale channel field-effect transistors (FETs) to extending Moore’s law.1–4 To date, there have been extensive experimental and theoretical efforts to research 2D semiconductors as promising candidates for channel materials of short-channel FETs, including black phosphorene, transition metal dichalcogenides (TMDCs), GaAs, InSe, and WSi2P2As2.5–9 Compared with the above-mentioned 2D semiconductors, monolayer blue phosphorene (BlueP) has attracted great attention and research interest due to its indirect bandgap of about 2 eV and unique anisotropic in-plane properties, since its successful preparation in the laboratory in 2016.10 Moreover, BlueP has both a higher carrier mobility (over 1000 cm2 V−1 s−1) than TMDCs and a higher on/off ratio than crystalline Si, resulting in a better balance between the carrier mobility and on/off ratio of FETs,11 making BlueP a very competitive channel material for short-channel FETs.12,13

It is well known that the performance of practical FETs strongly depends not only on the properties of channel material but also on the quality of gate dielectric layer and the properties of interface between dielectric and channel. However, one of the main challenges in achieving practical applications of 2D semiconductor-based FETs is to explore suitable gate dielectric materials interfaced with 2D semiconductors to guarantee efficient gate control,14,15 and to scale the thickness of the gate dielectric layer and to control its quality. An excellent gate dielectric for 2D semiconductors requires a large dielectric constant, which can effectively screen the charge scattering16,17 and can significantly reduce the device supply gate voltage and power consumption.18 Moreover, a desirable gate dielectric also demands low equivalent thickness (EOT) and a large bandgap,19 which can prevent the carrier transport to gate electrodes.20 Typical oxides used as gate dielectrics are SiO2, Al2O3, HfO2, and ZrO2 insulators.21–24 However, it is difficult to achieve high-quality interfaces between these dielectrics and 2D semiconductor channels due to high interface state density,25,26 which leads to the deterioration of device performance. Therefore, considerable efforts have been made to search for new suitable dielectrics. So far, it has been reported that 2D insulators such as h-BN, CaF2, and Bi2SeO5 layers can achieve high-quality interfaces on 2D TMDC semiconductors by van der Waals interactions27–31 and are expected to be used as gate dielectrics for future 2D FETs. Unfortunately, many problems and obstacles remain to be solved. For example, recent studies have shown that there exists a large leakage current in 2D devices using h-BN as a dielectric.32 Furthermore, whether these 2D dielectrics are also suitable for other 2D semiconductor-based FETs, such as BlueP-based FETs? In addition, the integration of 2D dielectric into 2D semiconductors at high quality for practical applications remains challenging. Although various efforts have been made, finding satisfied dielectrics and integrating them into the channels at high quality still remain a major obstacle to the further development of next-generation FETs for practical applications.

In particular, when a FET works under a gate voltage for practical applications, the dielectric-channel heterostructure is actually placed under a vertical external electric field. It has been reported that applied electric fields can significantly control the band alignment and Schottky barrier height of a heterojunction.33,34 Thus, as an excellent dielectric-channel heterostructure in a FET device, it is crucial that the dielectric and channel can exhibit their own good insulation and semiconductor properties in the absence of an electric field and can continue to maintain their respective desired properties under a vertical external electric field. However, until now, the effects of electric field on the electronic properties of dielectric and channel in their dielectric-channel heterostructure have been little studied.

Since HfO2 has a large dielectric constant (∼25) and excellent thermal stability with Si substrates,35 TiO2 has a very large dielectric constant (∼80) and easy synthesis and stability;36 BN has a dielectric constant about 5, very high thermal and chemical stability and easy stacking on 2D materials;37 monolayer CaF2 has a large dielectric constant (∼8.43), a wide bandgap (∼9.49), and dynamic stability;38 HfO2, TiO2, BN, and CaF2 insulator layers are selected as gate dielectrics of BlueP-based FET in the present work. We focus on investigating the structures and electronic properties of BlueP-dielectric heterojunctions, including BlueP-BN, BlueP-HfO2, BlueP-TiO2, and BlueP-CaF2. For the first time, an intensive insight is gained into the electronic properties of the heterojunctions between BlueP and these dielectrics via applying perpendicular external electric fields. The calculated results suggest that HfO2 and CaF2 dielectrics are suitable as gate dielectrics for BlueP-based FETs; furthermore, CaF2 is superior to HfO2. Then, we investigate the effects of interfacial intrinsic atomic vacancy on the structure and electronic properties of BlueP-HfO2 and BlueP-CaF2 heterostructures. In addition, we use hydrogen passivation to the F-vacancy of BlueP-CaF2 heterostructure to study its effective protection of BlueP channel and CaF2 dielectric, which will help improve the performance of 2D semiconductor-based nanoelectronic devices with CaF2 as a gate dielectric.

All calculations for BlueP-dielectric heterojunctions, including BN-BlueP-BN, HfO2-BlueP-HfO2, TiO2-BlueP-TiO2, and CaF2-BlueP-CaF2, were performed by first-principles calculations based on density-functional theory (DFT),39 as implemented in the Vienna Ab initio Simulation Package (VASP). The projector augmented wave (PAW) was used, and the Perdew–Burke–Ernzerhof (PBE) method of generalized gradient approximation (GGA)40–43 was employed to deal with the exchange–correlation energy. The energy cutoff of the plane wave basis was 520 eV, and the Monkhorst–Pack k-mesh density with ∼0.03 Å−1 was sampled to ensure numerical accuracy. Geometry optimizations were carried out until the Hellmann–Feynman force on each atom was less than 0.02 eV/Å and the total energy converged to 10−6 eV. In the supercell models for BlueP-dielectric heterojunctions, a vacuum separation of 20 Å was employed to avert the artificial coupling role between layers in two adjacent periodic images. The optimized PBE van der Waals (optB88-vdW) was used to consider the van der Waals interactions between layers in the heterojunctions.

Figures 2(a)2(e) show that our calculated lattice constants for monolayer BlueP, bilayer BN, monolayer CaF2, bulk HfO2, and bulk TiO2 were 3.28, 2.51, 3.59, 5.06, and 4.66 Å, respectively, which are in agreement with other theoretical calculations.38,44–47 Figures 1(a)1(f) show the energy band structures of monolayer BlueP and the dielectric BN, TiO2, HfO2, HfO2, and CaF2 thin-layers. Monolayer BlueP exhibits an indirect bandgap of 1.93 eV, which is consistent with our preceding reported result and those of other studies.12,48,49 Bilayer BN exhibits an indirect bandgap of 4.44 eV, which is close to the result of 4.37 eV of other research,50 and TiO2(101) thin-layer has a direct bandgap of 1.89 eV. The bandgaps for HfO2(111) thin-layers with thicknesses of three Hf-O atomic layers (labeled as HfO2-3) and four Hf-O atomic layers (labeled as HfO2-4) are very close, with direct bandgaps of 4.25 and 4.12 eV, respectively. Monolayer CaF2 exhibits an indirect bandgap of 7.26 eV, which agrees well with the previous result of 7.17 eV.38 

FIG. 1.

Band structures of (a) monolayer BlueP, (b) bilayer BN, (c) TiO2(101) thin-layer, (d) HfO2(111)-3 thin-layer, (e) HfO2(111)-4 thin-layer, and (f) monolayer CaF2.

FIG. 1.

Band structures of (a) monolayer BlueP, (b) bilayer BN, (c) TiO2(101) thin-layer, (d) HfO2(111)-3 thin-layer, (e) HfO2(111)-4 thin-layer, and (f) monolayer CaF2.

Close modal
Considering lattice matching and all strains being within 5%, lattice constants of BlueP are fixed during the construction of BlueP-BN, BlueP-TiO2, and BlueP-CaF2 heterojunctions, and the lattice constants of 2D BN, TiO2(101), and CaF2 were strained to adapt to them. However, for the formation of BlueP-HfO2 heterojunctions, the HfO2(111) stretches 2% and BlueP compresses 4%. In the supercell models, the 23 × 2 BlueP matches 7 × 21 BN, the 3 × 3 BlueP matches 1 × 2 TiO2(101), the 3×2 BlueP matches 3/2×1 HfO2(111), and the 3 × 1 BlueP matches 3 × 1 CaF2, as shown in Figs. 2(f)2(j). The stability of the heterojunction is described by the binding energy Eb, which can be defined as51 
(1)
where EBlueP-dielectric, EBlueP, and Edielectric are the total energies of the heterojunction, monolayer BlueP, and the dielectric thin-layer (HfO2-3, HfO2-4, BN, CaF2, and TiO2 layers), respectively. n is the number of P atoms for BlueP in the heterojunction. Our calculated results show that the binding energies of BlueP-BN, BlueP-TiO2, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 heterojunctions are −0.20, −0.23, −0.20, −0.18, and −0.11 eV/Atom, respectively. Negative binding energies indicate that heterojunction structures can be realized exothermically in the experiment. The binding energies show that all five heterojunctions are relatively easily available energetically.
FIG. 2.

Top and side views of the (a) monolayer BlueP, (b) bilayer BN, (c) TiO2(101) thin-layer, (d) HfO2(111) thin-layer, (e) monolayer CaF2, (f) BlueP-BN, (g) BlueP-TiO2, (h) BlueP-HfO2-3, (i) BlueP-HfO2-4, and (j) BlueP-CaF2 interfacial structures.

FIG. 2.

Top and side views of the (a) monolayer BlueP, (b) bilayer BN, (c) TiO2(101) thin-layer, (d) HfO2(111) thin-layer, (e) monolayer CaF2, (f) BlueP-BN, (g) BlueP-TiO2, (h) BlueP-HfO2-3, (i) BlueP-HfO2-4, and (j) BlueP-CaF2 interfacial structures.

Close modal

There is no obvious deformation at the interfaces when five 2D layers are combined to form heterojunctions. The interlayer spacings d1 and d2 are the same for BlueP-BN, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 heterojunctions, being 3.53, 3.23, 3.47, and 3.42 Å, respectively, except that the d1 and d2 of BlueP-TiO2 heterojunction are slightly different, which are 2.29 and 2.34 Å, respectively. Here, the marks of d1 and d2 are labeled in Fig. 2. Large interlayer spacings between BlueP and dielectric layers suggest weak interactions between the components of heterojunctions. As shown in Figs. 3(a)3(e), both BlueP and dielectrics basically remain their original semiconductor and dielectric properties. However, the band alignments are different in five heterogeneous structures, as shown in Figs. 3(f)3(h). In these figures, CBM and VBM are conduction band minimum and valence band maximum, respectively. In particular, for BlueP-HfO2 (or BlueP-CaF2) heterojunction, the band alignment is as follows: VBM-HfO2 (or VBM-CaF2) < VBM-BlueP < CBM-BlueP < CBM-HfO2 (or CBM-CaF2), called type-I. In type-I, the bandgap of HfO2 (or CaF2) is greater than that of BlueP. The CBM of HfO2 (or CaF2) is higher than that of BlueP, while the VBM of HfO2 (or CaF2) is lower than that of BlueP. This type-I enables effective carrier confinement within the BlueP layer, significantly suppressing leakage currents through spatial charge separation.59 For BlueP-BN heterojunction, VBM-BlueP < VBM-BN < CBM-BlueP < CBM-BN, called type-II. This type-II promotes efficient electron–hole pair separation across the interface, thereby minimizing recombination losses.60 For BlueP-TiO2, VBM-TiO2 < VBM-BlueP < CBM-TiO2 < CBM-BlueP, called type-III. As a result of weak coupling among components and band alignment, the VBM and CBM of entire heterojunctions are significantly changed compared to isolated BlueP and dielectrics. The bandgaps of BlueP-BN, BlueP-TiO2, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 are 1.73, 1.51, 1.68, 1.73, and 1.99 eV, respectively.

FIG. 3.

Band structures of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, (e) BlueP-CaF2 heterojunctions, and (f)–(h) three typical band alignment structures. The dashed lines represent Fermi levels. CBM and VBM denote the conduction band minimum and valence band maximum, respectively.

FIG. 3.

Band structures of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, (e) BlueP-CaF2 heterojunctions, and (f)–(h) three typical band alignment structures. The dashed lines represent Fermi levels. CBM and VBM denote the conduction band minimum and valence band maximum, respectively.

Close modal

In practical semiconductor nanotransistor applications, to minimize the leakage current through gate electrodes, the conduction band offset (CBO) and valence band offset (VBO) between gate dielectric and semiconductor channel should be larger than 1 eV.16 Our estimated CBO and VBO for BlueP-BN, BlueP-TiO2, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 are about 2.4 and 0.3, 0.3 and 0.7, 1.0 and 1.7, 1.0 and 1.6, and 2.4 and 4.0 eV, respectively, which can be obtained from Figs. 3(a)3(e). Obviously, only BlueP-HfO2 and BlueP-CaF2 can match the desired CBO and VBO well. Furthermore, CaF2 is superior to HfO2 as a dielectric for BlueP-based FETs.

When 2D layers combine to form a heterojunction, charge transfer may occur at the interface even if the interaction is weak, thus changing the band structure. The plane-averaged electron density differences of BlueP-BN, BlueP-TiO2, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 heterojunctions are shown in Figs. 4(a)4(e). The electron density difference was calculated as
(2)
where ρBlueP-dielectric, ρBlueP and ρdielectric are electron density of BlueP-dielectric heterojunction, isolated monolayer BlueP and isolated dielectric thin-layers, including BN, TiO2, HfO2-3, HfO2-4 and CaF2, respectively. As shown in Fig. 4, positive values represent electron accumulation, and negative values indicate electron depletion. The electron accumulation and depletion are mainly concentrated in interfacial region for all heterojunctions. The interactions between BlueP and dielectric in BlueP-BN and BlueP-TiO2 are slightly stronger than those in BlueP-HfO2 and BlueP-CaF2. These stronger interactions lead to a slightly larger amount of electron transfer in their interface region, which is consistent with the results of binding energy above. However, in general, the interactions in all five heterojunctions are weak. The amount of electron transfer at their interfaces is much smaller than that in BlueP/C2N heterojunction,52 especially in BlueP-HfO2-4 and BlueP-CaF2 heterojunctions.
FIG. 4.

Plane-averaged electron density differences of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, and (e) BlueP-CaF2 heterojunctions.

FIG. 4.

Plane-averaged electron density differences of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, and (e) BlueP-CaF2 heterojunctions.

Close modal

It is well-known that an excellent gate dielectric layer in 2D semiconductor-based nanoscale FETs requires a high potential barrier for electrons or holes so that the dielectric can perform its main screening role, which could effectively suppress current flow from the gate to the channel.53,54 When a gate voltage is applied to the FET, a vertical electric field is essentially applied to the dielectric-channel heterojunction. In this case, the dielectric still requires a high barrier for electrons or holes to perform its screening role. However, it is well known that applying an electric field can effectively modulate the electronic structure of 2D materials and allows significant control of band alignment and Schottky barrier height of the heterojunction.33,34 So, when a gate voltage is applied to the FET, whether the dielectric layer can preserve its good screening role is very important. Next, we mainly explore the electronic properties of heterostructures under applied electric fields and by analyzing the screening role of dielectric by valence and conduction band offsets (VBO and CBO). The electric fields are applied perpendicular to the interface of heterojunction, ranging from 0 to 0.6 V/Å, and the step is 0.1 V/Å. Figures 5(a)5(t) plot the electronic band structures of BlueP-BN, BlueP-TiO2, BlueP-HfO2-3, BlueP-HfO2-4, and BlueP-CaF2 heterojunctions under different applied electric fields (E). Figure 6(a) plots the bandgap evolution of BlueP and dielectric in five heterojunctions under the influence of an external electric field (E). For all five heterojunctions, as E increases, the bandgap of BlueP changes little, while the bandgap of dielectric decreases significantly. Careful observation from Figs. 5(a)5(t) shows that the band alignment for BlueP and dielectric changes greatly. Then, we further analyze the effects of an external electric field on the band structures of dielectric layers located on both sides of BlueP. Figures 7(a)7(d) show the band structures of layer-I, layer-II, layer-III, and layer-IV dielectric layers (labeled in Fig. 2) in BlueP-HfO2-4 and BlueP-CaF2 heterojunctions without and under an external electric field of 0.6 V/Å. Similar behavior can also be observed in BlueP-BN, BlueP-TiO2, and BlueP-HfO2-3 heterojunctions. It can be found that under E, the band structures of layer-I and layer-II dielectric layers move downward, and the band structures of layer-III and layer-IV move upward. The conduction bands of the whole dielectric are mainly contributed by layer-I and layer-II, and the valence bands are mainly contributed by layer-III and layer-IV. Consequently, all the bandgaps of the whole dielectric become small, as shown in Fig. 6(a). Observed clearly from the relative movement of valence bands, as shown in Fig. 7, it is found that the response of outer dielectric layers (layer-I and layer-IV) to an external electric field is greater than that of inner layers (layer-II and layer-III). As E increases, the band alignment types for BlueP-BN, BlueP-TiO2, and BlueP-CaF2 heterojunctions roughly remain their original types until E = 0.6 V/Å. However, the band alignment types for BlueP-HfO2-3 and BlueP-HfO2-4 heterojunctions have changed at about E = 0.4 V/Å.

FIG. 5.

Electronic band structures of [(a)–(d)] BlueP-BN, [(e)–(h)] BlueP-TiO2, [(i)–(l)] BlueP-HfO2-3, [(m)–(p)] BlueP-HfO2-4, and [(q)–(t)] BlueP-CaF2 heterojunctions under external electric fields E.

FIG. 5.

Electronic band structures of [(a)–(d)] BlueP-BN, [(e)–(h)] BlueP-TiO2, [(i)–(l)] BlueP-HfO2-3, [(m)–(p)] BlueP-HfO2-4, and [(q)–(t)] BlueP-CaF2 heterojunctions under external electric fields E.

Close modal
FIG. 6.

(a) Bandgaps of BlueP and dielectric in heterojunctions as a function of electric fields. (b) Band alignment of BlueP-BN, BlueP-TiO2, BlueP-HfO2, and BlueP-CaF2 heterojunctions under E.

FIG. 6.

(a) Bandgaps of BlueP and dielectric in heterojunctions as a function of electric fields. (b) Band alignment of BlueP-BN, BlueP-TiO2, BlueP-HfO2, and BlueP-CaF2 heterojunctions under E.

Close modal
FIG. 7.

Energy band structures of layer-I, layer-II, layer-III, and layer-IV shown in Fig. 2, for (a) BlueP-HfO2-4 and (c) BlueP-CaF2 heterojunctions without external electric fields E, for (b) BlueP-HfO2-4 and (d) BlueP-CaF2 under E = 0.6 V/Å.

FIG. 7.

Energy band structures of layer-I, layer-II, layer-III, and layer-IV shown in Fig. 2, for (a) BlueP-HfO2-4 and (c) BlueP-CaF2 heterojunctions without external electric fields E, for (b) BlueP-HfO2-4 and (d) BlueP-CaF2 under E = 0.6 V/Å.

Close modal

Next, we explore why the energy bands of heterojunctions vary with external electric fields. When E is exerted, the different junctions (atom layers) of the heterojunction have their own different Fermi energies (EF), shown in Fig. 6(b), and the EF of the whole heterojunction is changed with E. As a consequence, the variation of the bandgap of the whole heterojunction under E is attributed to the shifts of EF for the BlueP layer and two dielectric layers on both sides under E, which promotes the transfer of electrons between BlueP and dielectric layers. Notably, when E reaches 0.6 V/Å, the BlueP-BN heterojunction abruptly becomes metallic due to the breakdown and charge tunneling of dielectric layers. Similar behavior can also be observed in BlueP-TiO2 and BlueP-HfO2-3 heterojunctions, as shown in Figs. 5(d), 5(h), and 5(l), resulting in large vertical leakage currents and poor dielectric properties of dielectric layers. If this happens in a FET device, it will cause a large gate leakage current and reduce the performance of the device. Fortunately, the BlueP-HfO2-4 and BlueP-CaF2 heterojunctions remain semiconductor, as shown in Figs. 5(p) and 5(t). However, as E increases, only the CaF2 dielectric layer can still satisfy the requirements that CBO and VBO between CaF2 and BlueP are larger than 1 eV, which can play a protective role as gate dielectrics. In particular, its bandgap still mainly depends on BlueP, which is less affected by an external electric field, as shown in Figs. 5(q)5(t). This implies that the resistance of CaF2 dielectric layer to an external electric field is the strongest among the five dielectric layers. Furthermore, monolayer CaF2 dielectric is significantly thinner than the BlueP-HfO2-4 layer, which is an urgent requirement for reducing dielectric layer thickness in nanoscale FETs. These indicate that monolayer CaF2 may be a good gate dielectric for BlueP-based FETs.

Under the external electric field, the charge transfer between the dielectrics and the BlueP will be affected evidently. The plane-averaged charge difference of the heterojunctions under E is determined as52 
(3)
where ρTE0(x,y,z)dxdy and ρTE=0(x,y,z)dxdy are the charge densities of the heterojunction with and without the electric field E, respectively. The positive and negative values represent charge accumulation and charge depletion at the position, respectively. Figures 8(a)8(e) plot the averaged charge difference of five heterojunctions along the z-direction (perpendicular to the heterojunction interface). On the whole, for all heterojunctions, the amplitude of transferring electrons depends on the strength of E. What makes a difference is that the charge transfer between BN layers in BlueP-BN heterojunction is much larger than that of the other four heterojunctions under the circumstance of the E field. Obviously, the inner HfO2, TiO2, or CaF2 layers in BlueP-HfO2, BlueP-TiO2, or BlueP-CaF2 heterojunctions have a smaller response to an external electric field than their outer and interfacial layers. The charge transfer of their inner layers is small, which also reflects their better dielectric properties. Furthermore, it is worth noting that when E < 0.6 V/Å, for all five heterojunctions, the amounts of charge accumulation and depletion of the surface layer on both sides of the heterojunction are roughly equal, as shown in Figs. 8(a)8(e). While E reaches 0.6 V/Å, for BlueP-BN, BlueP-TiO2, and BlueP-HfO2-3 heterojunctions, the amount of charge accumulation (on the left surface layer) is significantly smaller than that of charge depletion (on the right surface layer). It suggests that, responding to an external electric field, the electrons move from the left side to the right side of the dielectric layer and accumulate on the right surface. But the amounts of charge accumulation and depletion of surface layers on both sides for BlueP-HfO2-4 and BlueP-CaF2 heterojunctions are still roughly equal, which again indicates that HfO2-4 and CaF2 has better dielectric properties and good screening role to resist external electric field. They make it difficult for electrons to move freely from the left surface to the right surface of heterojunction under an electric field. These results coincide with those derived from the electronic band structures.
FIG. 8.

Plane-averaged electron density differences of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, and (e) BlueP-CaF2 heterojunctions under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

FIG. 8.

Plane-averaged electron density differences of (a) BlueP-BN, (b) BlueP-TiO2, (c) BlueP-HfO2-3, (d) BlueP-HfO2-4, and (e) BlueP-CaF2 heterojunctions under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

Close modal

In addition, an ab initio molecular dynamics (AIMD) simulation was performed to further investigate the thermodynamic stabilities of BlueP-HfO2-4 and BlueP-CaF2, which are very important in practical applications for FETs. Canonical ensemble (NVT) was used in AIMD simulations, the simulation time was 10 ps, and the time step was 2 fs. As shown in Fig. 9, our AIMD simulations show only minimal variation in the energy and no significant reconstruction in the structure throughout the calculation at 300 K. This reveals the excellent thermal stabilities of BlueP-HfO2-4 and BlueP-CaF2 heterojunctions at room temperature.

FIG. 9.

Evolution of the total energy for (a) BlueP-HfO2-4 and (b) BlueP-CaF2 at 300 K using the AIMD simulation. The insets represent the snapshot of structures at the end of the simulations.

FIG. 9.

Evolution of the total energy for (a) BlueP-HfO2-4 and (b) BlueP-CaF2 at 300 K using the AIMD simulation. The insets represent the snapshot of structures at the end of the simulations.

Close modal

Atomic vacancy is easy to generate in the preparation process of 2D materials, which can have significant impacts on the structure and electronic properties of materials, especially for monolayer structures. Next, we focus on the effects of interfacial intrinsic atomic vacancies on the structure and electronic properties of BlueP-HfO2 and BlueP-CaF2 heterostructures, including O-vacancy in an interfacial HfO2 layer and F-vacancy in CaF2. O-vacancy defects are more likely to appear on the HfO2 surface in the interfacial region,55–57 which may affect the dielectric properties of HfO2. Therefore, the performance of HfO2 as a gate dielectric for 2D BlueP-based FETs with O-vacancy defects in the interfacial region was further investigated in this paper. For simplicity, we consider the case in which O-vacancy defects are only in one side of the interfacial region of BlueP-HfO2-4, as shown in Fig. 10(a), called BlueP-HfO2-4 (VO). The calculated interfacial binding energy Eb is −0.82 eV/Atom. Compared with the Eb of −0.18 eV/Atom for BlueP-HfO2-4 without O-vacancy defects, O-vacancy defects have resulted in an obvious increase in Eb, which indicates that the O-vacancy is easy to exist at the interface of the BlueP-HfO2-4 heterojunction. After geometrical optimization, the O-vacancy makes the HfO2 layer, where the vacancy is located close to the BlueP layer, and the interlayer distance d1 is 1.26 Å, which is about 2.21 Å shorter than that of BlueP-HfO2-4. Another interlayer distance d2 is 2.84 Å, which is 0.63 Å shorter than that of BlueP-HfO2-4. This indicates that the O-vacancy induces large deformations, leading to the reconstruction of the structure.

FIG. 10.

(a) Top and side views of the structure for BlueP-HfO2-4 with O vacancy. Electronic band structures of the BlueP-HfO2-4 (VO) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

FIG. 10.

(a) Top and side views of the structure for BlueP-HfO2-4 with O vacancy. Electronic band structures of the BlueP-HfO2-4 (VO) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

Close modal

Figure 10(b) illustrates that the type-I energy band alignment is preserved in the BlueP-HfO2-4 (VO) heterojunction without an external electric field. However, the band offsets vary compared with those of BlueP-HfO2-4 without O-vacancy. When E is exerted, the different junctions of heterojunction have similar behavior as they do in the BlueP-HfO2-4 heterojunction without O-vacancy, as shown in Fig. 5. The bandgap of heterojunction is reduced as E increases. The difference is that when E reaches 0.6 V/Å, BlueP-HfO2-4 (VO) becomes metallic. It implies that the HfO2-4 dielectric with O-vacancy is less resistant to an external electric field than HfO2-4 without O-vacancy, indicating that O-vacancy at the interface will reduce the dielectric properties of the gate dielectric. Moreover, comparing Fig. 8(d) with Fig. 11, when E reaches 0.6 V/Å, the amount of charge accumulation on the left surface layer (the red dashed box on the left of Fig. 11) is slightly smaller than the amount of charge depletion on the right surface layer (the red dashed box on the right of Fig. 11), suggesting that the electrons move to the right side of the dielectric layer and accumulate on this surface to respond to an external electric field. It coincides with the result of the electronic band structure in Fig. 10(e). The interfacial O-vacancy not only has a strong effect on the structure of the BlueP-HfO2-4 heterostructure but also has a serious effect on its electrical properties, greatly reducing the screening role of the HfO2 dielectric layer. If such O-vacancy occurs in 2D semiconductor-based FETs with HfO2 as a gate dielectric, leakage current through gate will greatly increase and the electron transport along channel will be greatly reduced, resulting in the deterioration of device performance.

FIG. 11.

Plane-averaged difference in the electron density for the BlueP-HfO2-4 (VO) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

FIG. 11.

Plane-averaged difference in the electron density for the BlueP-HfO2-4 (VO) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

Close modal

Figure 12(a) shows the structure of BlueP-CaF2 with F-vacancy defects only in one side of the interface region, called BlueP-CaF2 (VF). Our calculated interfacial binding energy Eb of BlueP-CaF2 (VF) is about −0.43 eV/Atom, indicating that F-vacancy defects also have resulted in an obvious increase in Eb compared with that of BlueP-CaF2 without F-vacancy defects (Eb = −0.11 eV/Atom). It indicates that the F-vacancy can exist at the interface of the BlueP-CaF2 (VF) heterojunction. However, F-vacancy causes significant structural deformation, as shown in Fig. 12(a). F-vacancy makes the CaF2 layer closer to the BlueP layer, and the interlayer distances d1 and d2 are 2.74 and 2.69 Å, respectively, which are about 0.68 and 0.72 Å shorter than that of BlueP-CaF2. Similar to the BlueP-HfO2-4 (VO) heterojunction, F-vacancy does not change the type-I band alignment, as shown in Fig. 12(b). However, as E increases, the bandgap of BlueP is almost constant and the CaF2 bandgap becomes slightly smaller, as can be clearly seen in Figs. 12(c)12(e). Moreover, BlueP-CaF2 (VF) exhibits metallic. It implies that F-vacancy at the interface will seriously damage the dielectric properties of CaF2. Comparing Fig. 13 with Fig. 8(e), when E is exerted, the charge transfer amount of BlueP-CaF2 (VF) at the left interface is distinctly smaller than that at the right interface where F-vacancy is located, which is different from BlueP-CaF2 without F-vacancy. It indicates that the gain and loss of electrons near the vacancy increase under E. The above-mentioned results show that F-vacancy seriously affects the structure of the BlueP-CaF2 heterostructure and its electrical properties, leading to great deterioration of the screening role of the CaF2 dielectric layer. If such F-vacancy occurs in 2D semiconductor-based FETs with CaF2 as a gate dielectric, the performance of the device will also be greatly reduced. The vacancy both seriously reduces the dielectric properties of BlueP-CaF2 and BlueP-HfO2. However, compared with BlueP-HfO2, the F-vacancy deteriorates the dielectric properties of BlueP-CaF2 more severely. BlueP-CaF2 exhibits metallic even without an external electric field, as shown in Fig. 12(b).

FIG. 12.

(a) Top and side views of the structure for BlueP-CaF2 with F-vacancy. Electronic band structures of the BlueP-CaF2 (VF) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

FIG. 12.

(a) Top and side views of the structure for BlueP-CaF2 with F-vacancy. Electronic band structures of the BlueP-CaF2 (VF) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

Close modal
FIG. 13.

Plane-averaged difference in the electron density for the BlueP-CaF2 (VF) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

FIG. 13.

Plane-averaged difference in the electron density for the BlueP-CaF2 (VF) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

Close modal

Yang et al. have found that hydrogen passivation can effectively terminate the dangling bonds at the surface or interface.58 Next, we study the structure and electrical properties of the BlueP-CaF2 (VF) heterojunction after hydrogen passivation. Our calculated adsorption energy for BlueP-CaF2 (VF) with hydrogen passivation on the F vacancies is −0.11 eV/Atom. The negative adsorption energy suggests that hydrogen adsorption is energetically favorable and easy to achieve. Comparing the structures of heterojunctions with/without hydrogen passivation on the F vacancies, as shown in Figs. 12(a) and 14(a), hydrogen adsorption apparently makes CaF2 remain its original structure and nearly intact, thus greatly improving the interface quality between BlueP and CaF2. More importantly, hydrogen adsorption on the F vacancies of CaF2 can remarkably improve the dielectric properties of CaF2 and protect the semiconductor properties of BlueP. It can be clearly seen from Figs. 14(b)14(e) that the bandgap of hydrogen passivated BlueP-CaF2 (VF) is nearly identical to that of BlueP-CaF2 without F-vacancy [Figs. 5(q)5(t)] except for adding the electron states of hydrogen. Expectedly, the better dielectric properties of CaF2 are reached after hydrogen passivation. As shown in Fig. 15, hydrogen passivation well suppressed the movement of Fermi level to the conduction band edge of BlueP-CaF2 (VF). More excitingly, when the external electric field is exerted, hydrogen passivation well inhibits the electrons from moving from the left to the right of BlueP-CaF2 (VF). It implies that the hydrogenation strategy can very effectively protect the CaF2 dielectric and BlueP semiconductor. These results will help improve the performance of 2D semiconductor-based nanoelectronic devices with CaF2 as a gate dielectric.

FIG. 14.

(a) Top and side views of the structure for hydrogen passivated BlueP-CaF2 (VF). Electronic band structures of the hydrogen passivated BlueP-CaF2 (VF) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

FIG. 14.

(a) Top and side views of the structure for hydrogen passivated BlueP-CaF2 (VF). Electronic band structures of the hydrogen passivated BlueP-CaF2 (VF) heterojunction (b) without electric field E and (c)–(e) under E = 0.0, 0.2, 0.4, and 0.6 V/Å.

Close modal
FIG. 15.

Plane-averaged difference in the electron density for the hydrogen passivated BlueP-CaF2 (VF) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

FIG. 15.

Plane-averaged difference in the electron density for the hydrogen passivated BlueP-CaF2 (VF) heterojunction under E = 0.2, 0.4, 0.5, and 0.6 V/Å.

Close modal

In summary, the binding energies, band structures, and electronic properties of BlueP-dielectric heterojunctions, including BlueP-BN, BlueP-HfO2, BlueP-TiO2, and BlueP-CaF2 were studied by first-principles calculations. Moreover, the electronic properties of all heterojunctions under perpendicular external electric fields were studied. The calculated results show that all heterojunctions are relatively easily available energetically. The results of band structures show that HfO2 layer and monolayer CaF2 dielectrics are appropriate as gate dielectrics for BlueP-based FETs, and monolayer CaF2 is superior to HfO2 dielectric. AIMD calculations reveal the excellent thermal stability of BlueP-HfO2-4 and BlueP-CaF2 heterojunctions at room temperature. In addition, the effects of interfacial intrinsic atomic vacancy on the structure and electronic properties of BlueP-HfO2 and BlueP-CaF2 heterostructures with perpendicular external electric fields and hydrogen passivation to the F-vacancy of BlueP-CaF2 (VF) were also studied. The results show that the interfacial intrinsic atomic vacancy seriously affects the electrical properties of the whole heterojunction, greatly reducing the screening role of the dielectric layer. Fortunately, we found that hydrogen passivation to the F-vacancy of BlueP-CaF2 (VF) can effectively protect the semiconductor properties of BlueP and dielectric properties of CaF2. It suggests that the hydrogen passivation strategy may help improve the performance of 2D semiconductor-based nanoelectronic devices with CaF2 as a gate dielectric.

This work was supported by the National Natural Science Foundation of China (Grant Nos. 62174032 and 62474041) and the Natural Science Foundation of Fujian Province of China (Grant Nos. 2021J01189 and 2021J01188). This work was carried out at the National Supercomputer Center in Tianjin, and the calculations were performed on Tianhe-1(A).

The authors have no conflicts to disclose.

Xian Lin: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Visualization (equal); Writing – original draft (equal). Jian-Min Zhang: Data curation (equal); Methodology (equal); Validation (equal). Guigui Xu: Data curation (equal); Methodology (equal); Validation (equal). Kehua Zhong: Funding acquisition (lead); Supervision (equal); Writing – review & editing (equal). Zhigao Huang: Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article.

1.
J.
Miao
,
X.
Liu
,
K.
Jo
,
K.
He
,
R.
Saxena
,
B.
Song
,
H.
Zhang
,
J.
He
,
M.-G.
Han
,
W.
Hu
, and
D.
Jariwala
,
Nano Lett.
20
,
2907
(
2020
).
2.
Y.
Liu
,
X.
Duan
,
H.-J.
Shin
,
S.
Park
,
Y.
Huang
, and
X.
Duan
,
Nature
591
,
43
(
2021
).
3.
B.
Radisavljevic
,
A.
Radenovic
,
J.
Brivio
,
V.
Giacometti
, and
A.
Kis
,
Nat. Nanotechnol.
6
,
147
(
2011
).
4.
M.
Chhowalla
,
D.
Jena
, and
H.
Zhang
,
Nat. Rev. Mater.
1
,
16052
(
2016
).
5.
R.
Quhe
,
Q.
Li
,
Q.
Zhang
,
Y.
Wang
,
H.
Zhang
,
J.
Li
,
X.
Zhang
,
D.
Chen
,
K.
Liu
,
Y.
Ye
,
L.
Dai
,
F.
Pan
,
M.
Lei
, and
J.
Lu
,
Phys. Rev. Appl.
10
,
024022
(
2018
).
6.
J. Y.
Lim
,
M.
Kim
,
Y.
Jeong
,
K. R.
Ko
,
S.
Yu
,
H. G.
Shin
,
J. Y.
Moon
,
Y. J.
Choi
,
Y.
Yi
,
T.
Kim
, and
S.
Im
,
npj 2D Mater. Appl.
2
,
37
(
2018
).
7.
Q.
Li
,
S.
Fang
,
S.
Liu
,
L.
Xu
,
L.
Xu
,
C.
Yang
,
J.
Yang
,
B.
Shi
,
J.
Ma
,
J.
Yang
,
R.
Quhe
, and
J.
Lu
,
ACS Appl. Mater. Interfaces
14
,
23597
(
2022
).
8.
Y.-T.
Huang
,
Y.-H.
Chen
,
Y.-J.
Ho
,
S.-W.
Huang
,
Y.-R.
Chang
,
K.
Watanabe
,
T.
Taniguchi
,
H.-C.
Chiu
,
C.-T.
Liang
,
R.
Sankar
,
F.-C.
Chou
,
C.-W.
Chen
, and
W.-H.
Wang
,
ACS Appl. Mater. Interfaces
10
,
33450
(
2018
).
9.
C.-x.
Xia
,
J.
Du
,
X.-w.
Huang
,
W.-b.
Xiao
,
W.-q.
Xiong
,
T.-x.
Wang
,
Z.-m.
Wei
,
Y.
Jia
,
J.-j.
Shi
, and
J.-b.
Li
,
Phys. Rev. B
97
,
115416
(
2018
).
10.
Y.
Guo
,
F. Q.
Wang
, and
Q.
Wang
,
Appl. Phys. Lett.
111
,
073503
(
2017
).
11.
Y.
Liu
,
X.
Duan
,
Y.
Huang
, and
X.
Duan
,
Chem. Soc. Rev.
47
,
6388
(
2018
).
12.
J.
Xiao
,
M.
Long
,
C.-S.
Deng
,
J.
He
,
L.-L.
Cui
, and
H.
Xu
,
J. Phys. Chem. C
120
,
4638
(
2016
).
13.
J.
Xiao
,
M.
Long
,
X.
Zhang
,
J.
Ouyang
,
H.
Xu
, and
Y.
Gao
,
Sci. Rep.
5
,
9961
(
2015
).
14.
S.
Das
,
A.
Sebastian
,
E.
Pop
,
C. J.
McClellan
,
A. D.
Franklin
,
T.
Grasser
,
T.
Knobloch
,
Y.
Illarionov
,
A. V.
Penumatcha
,
J.
Appenzeller
,
Z.
Chen
,
W.
Zhu
,
I.
Asselberghs
,
L.-J.
Li
,
U. E.
Avci
,
N.
Bhat
,
T. D.
Anthopoulos
, and
R.
Singh
,
Nat. Electron.
4
,
786
(
2021
).
15.
Y. C.
Lin
,
C. M.
Lin
,
H. Y.
Chen
,
S.
Vaziri
,
X.
Bao
,
W. Y.
Woon
,
H.
Wang
, and
S. S.
Liao
,
IEEE Trans. Electron Devices
70
,
1454
(
2023
).
16.
17.
D.
Jena
and
A.
Konar
,
Phys. Rev. Lett.
98
,
136805
(
2007
).
18.
S.
Yang
,
K.
Liu
,
Y.
Xu
,
L.
Liu
,
H.
Li
, and
T.
Zhai
,
Adv. Mater.
35
,
2207901
(
2023
).
19.
B.
Wang
,
W.
Huang
,
L.
Chi
,
M.
Al-Hashimi
,
T. J.
Marks
, and
A.
Facchetti
,
Chem. Rev.
118
,
5690
(
2018
).
20.
Y.
Xu
,
H.
Sun
,
A.
Liu
,
H.
Zhu
,
W.
Li
et al,
Adv. Mater.
30
,
1706364
(
2018
).
21.
S. H.
Kim
,
K.
Hong
,
W.
Xie
,
K. H.
Lee
,
S.
Zhang
,
T. P.
Lodge
, and
C. D.
Frisbie
,
Adv. Mater.
25
,
1822
(
2013
).
22.
J.
Zhang
,
M.
Jia
,
M. G.
Sales
,
Y.
Zhao
,
G.
Lin
,
P.
Cui
,
C.
Santiwipharat
,
C.
Ni
,
S.
McDonnell
, and
Y.
Zeng
,
ACS Appl. Electron. Mater.
3
,
5483
(
2021
).
23.
G.
He
,
Z.
Sun
,
G.
Li
, and
L.
Zhang
,
Crit. Rev. Solid State Mater. Sci.
37
,
131
(
2012
).
24.
C.
Henkel
,
S.
Abermann
,
O.
Bethge
,
G.
Pozzovivo
,
P.
Klang
,
M.
Reiche
, and
E.
Bertagnolli
,
IEEE Trans. Electron Devices
57
,
3295
(
2010
).
25.
Y. Y.
Illarionov
,
T.
Knobloch
,
M.
Jech
,
M.
Lanza
,
D.
Akinwande
,
M. I.
Vexler
,
T.
Mueller
,
M. C.
Lemme
,
G.
Fiori
,
F.
Schwierz
, and
T.
Grasser
,
Nat. Commun.
11
,
3385
(
2020
).
26.
X.
Zou
,
J.
Wang
,
C.-H.
Chiu
,
Y.
Wu
,
X.
Xiao
,
C.
Jiang
,
W.-W.
Wu
,
L.
Mai
,
T.
Chen
,
J.
Li
,
J. C.
Ho
, and
L.
Liao
,
Adv. Mater.
26
,
6255
(
2014
).
27.
S.
Fukamachi
,
P.
Solís-Fernández
,
K.
Kawahara
,
D.
Tanaka
,
T.
Otake
,
Y.-C.
Lin
,
K.
Suenaga
, and
H.
Ago
,
Nat. Electron.
6
,
126
(
2023
).
28.
Y. Y.
Illarionov
,
A. G.
Banshchikov
,
D. K.
Polyushkin
,
S.
Wachter
,
T.
Knobloch
,
M.
Thesberg
,
L.
Mennel
,
M.
Paur
,
M.
Stöger-Pollach
,
A.
Steiger-Thirsfeld
,
M. I.
Vexler
,
M.
Waltl
,
N. S.
Sokolov
,
T.
Mueller
, and
T.
Grasser
,
Nat. Electron.
2
,
230
(
2019
).
29.
J.-K.
Huang
,
Y.
Wan
,
J.
Shi
,
J.
Zhang
,
Z.
Wang
,
W.
Wang
,
N.
Yang
,
Y.
Liu
,
C.-H.
Lin
,
X.
Guan
,
L.
Hu
,
Z.-L.
Yang
,
B.-C.
Huang
,
Y.-P.
Chiu
,
J.
Yang
,
V.
Tung
,
D.
Wang
,
K.
Kalantar-Zadeh
,
T.
Wu
,
X.
Zu
,
L.
Qiao
,
L.-J.
Li
, and
S.
Li
,
Nature
605
,
262
(
2022
).
30.
G.-H.
Lee
,
Y.-J.
Yu
,
X.
Cui
,
N.
Petrone
,
C.-H.
Lee
,
M. S.
Choi
,
D.-Y.
Lee
,
C.
Lee
,
W. J.
Yoo
,
K.
Watanabe
,
T.
Taniguchi
,
C.
Nuckolls
,
P.
Kim
, and
J.
Hone
,
ACS Nano
7
,
7931
(
2013
).
31.
X.
Cui
,
G.-H.
Lee
,
Y. D.
Kim
,
G.
Arefe
,
P. Y.
Huang
,
C.-H.
Lee
,
D. A.
Chenet
,
X.
Zhang
,
L.
Wang
,
F.
Ye
,
F.
Pizzocchero
,
B. S.
Jessen
,
K.
Watanabe
,
T.
Taniguchi
,
D. A.
Muller
,
T.
Low
,
P.
Kim
, and
J.
Hone
,
Nat. Nanotechnol.
10
,
534
(
2015
).
32.
T.
Knobloch
,
Y. Y.
Illarionov
,
F.
Ducry
,
C.
Schleich
,
S.
Wachter
,
K.
Watanabe
,
T.
Taniguchi
,
T.
Mueller
,
M.
Waltl
,
M.
Lanza
,
M. I.
Vexler
,
M.
Luisier
, and
T.
Grasser
,
Nat. Electron.
4
,
98
(
2021
).
33.
C.
Wang
,
L.
Peng
,
Q.
Qian
,
J.
Du
,
S.
Wang
, and
Y.
Huang
,
Small
14
,
1703536
(
2018
).
34.
J. E.
Padilha
,
R. H.
Miwa
,
A. J. R.
da Silva
, and
A.
Fazzio
,
Phys. Rev. B
95
,
195143
(
2017
).
35.
S.
Kol
and
A.
Oral
,
Acta Phys. Pol. A
136
,
873
(
2019
).
36.
P.
Dwivedi
,
N.
Chauhan
,
V.
Dhyani
,
D. S.
Kumar
, and
S.
Dhanekar
, “
Design, fabrication, characterization and packaging of bottom gate and nano-porous TiO2 based FET
,” in
2017 IEEE 17th International Conference on Nanotechnology (IEEE-NANO)
(
IEEE
,
2017
), p.
946
.
37.
G.
Cassabois
,
P.
Valvin
, and
B.
Gil
,
Nat. Photonics
10
,
262
(
2016
).
38.
M.
Baskurt
,
M.
Yagmurcukardes
,
F. M.
Peeters
, and
H.
Sahin
,
J. Chem. Phys.
152
,
164116
(
2020
).
39.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
,
1758
(
1999
).
40.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
41.
G.
Kresse
and
J.
Hafner
,
Phys. Rev. B
48
,
13115
(
1993
).
42.
43.
G.
Kresse
and
J.
Furthmüller
,
Comput. Mater. Sci.
6
,
15
(
1996
).
44.
J. L.
Zhang
,
S.
Zhao
,
C.
Han
,
Z.
Wang
,
S.
Zhong
,
S.
Sun
,
R.
Guo
,
X.
Zhou
,
C. D.
Gu
,
K. D.
Yuan
,
Z.
Li
, and
W.
Chen
,
Nano Lett.
16
,
4903
(
2016
).
45.
Z.
Yang
and
J.
Ni
,
J. Appl. Phys.
107
,
104301
(
2010
).
46.
G. H.
Chen
,
Z. F.
Hou
, and
X. G.
Gong
,
Comput. Mater. Sci.
44
,
46
(
2008
).
47.
A.
Grünebohm
,
M.
Siewert
,
C.
Ederer
, and
P.
Entel
,
Ferroelectrics
429
,
31
(
2012
).
48.
W.
Chen
,
X.
Lin
,
G.
Xu
,
K.
Zhong
,
J.-M.
Zhang
, and
Z.
Huang
,
J. Phys.: Condens. Matter
36
,
125303
(
2024
).
49.
K.
Xu
,
Y.
Wang
,
Y.
Zhao
, and
Y.
Chai
,
J. Mater. Chem. C
5
,
376
(
2017
).
50.
Y.
Fujimoto
and
S.
Saito
,
J. Ceram. Soc. Jpn.
124
,
584
(
2016
).
51.
K.
Yang
,
Z.
Ding
,
Q.
Hu
,
J.
Sun
, and
Q.
Li
,
Crystals
13
,
1077
(
2023
).
52.
Z.
Zhang
,
Z.
Xie
,
J.
Liu
,
Y.
Tian
,
Y.
Zhang
,
X.
Wei
,
T.
Guo
,
L.
Ni
,
J.
Fan
,
Y.
Weng
et al,
Phys. Chem. Chem. Phys.
22
,
5873
(
2020
).
53.
L.
Huang
,
H.
Liu
, and
W.
Cui
,
ACS Appl. Electron. Mater.
5
,
5082
(
2023
).
54.
M. R.
Osanloo
,
M. L.
Van de Put
,
A.
Saadat
, and
W. G.
Vandenberghe
,
Nat. Commun.
12
,
5051
(
2021
).
55.
J.
Kang
,
D. Y.
Kim
, and
K.-J.
Chang
,
J. Korean Phys. Soc.
50
,
552
(
2007
).
56.
G.
Groeseneken
,
R.
Degraeve
,
T.
Nigam
,
G.
Van den bosch
, and
H.
Maes
,
Microelectron. Eng.
49
,
27
(
1999
).
57.
R.
Degraeve
,
B.
Kaczer
, and
G.
Groeseneken
,
Microelectron. Reliab.
39
,
1445
(
1999
).
58.
Y.
Yang
,
T.
Yang
,
T.
Song
,
J.
Zhou
,
J.
Chai
,
L. M.
Wong
,
H.
Zhang
,
W.
Zhu
,
S.
Wang
, and
M.
Yang
,
Nano Res.
15
,
4646
(
2022
).
59.
V. O.
Özçelik
,
J. G.
Azadani
,
C.
Yang
,
S. J.
Koester
, and
T.
Low
,
Phys. Rev. B
94
,
035125
(
2016
).
60.
J.
Robertson
,
J. Vac. Sci. Technol. A
31
,
050821
(
2013
).