Next Article in Journal
Nucleation of L12-Al3M (M = Sc, Er, Y, Zr) Nanophases in Aluminum Alloys: A First-Principles ThermodynamicsStudy
Previous Article in Journal
Zinc(II) and Copper(II) Complexes of 4-Styrylpyridine and 1-Adamantanecarboxylic Acid: Syntheses, Crystal Structures, and Photopolymerization
Previous Article in Special Issue
Excitation Density Effects in the Luminescence Yield and Kinetics of MAPbBr3 Single Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen-like Impurity States in β-Ga2O3/(AlxGa1−x)2O3 Core/Shell Nanostructures: Comparison between Nanorods and Nanotubes

1
College of Sciences, Inner Mongolia University of Technology, Hohhot 010051, China
2
School of Physical Science and Technology, Inner Mongolia University, Hohhot 010021, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(8), 1227; https://doi.org/10.3390/cryst13081227
Submission received: 28 July 2023 / Revised: 4 August 2023 / Accepted: 5 August 2023 / Published: 9 August 2023
(This article belongs to the Special Issue Advances in Crystals for Optoelectronics)

Abstract

:
The binding energy of an off-center hydrogen-like impurity in an ultra-wide band gap β-Ga2O3/(AlxGa1−x)2O3 core/shell nanostructure is studied using a variational method combined with a finite-difference algorithm. Four impurity states with the radial and axial quantum numbers being 0 or 1 in two kinds of core/shell nanostructures, including nanorods and double-walled nanotubes, are taken into account in the numerical calculations. The variation trends in binding energy corresponding to the four impurity states as functions of structural dimension and Al composition differ in nanorods and nanotubes when the impurity moves toward the interface between the Ga2O3 and (AlxGa1−x)2O3 layers. The quantum confinement due to the structural geometry has a considerable influence on the probability density of the impurity states as well as the impurity binding energy. The numerical results will pave the way toward theoretical simulation of the electron states in rapidly developing β-Ga2O3 low-dimensional material systems for optoelectronic device applications.

1. Introduction

Core–shell nanostructured semiconductors, such as nanorods, nanotubes, and nanodots, are regarded as an intriguing class of materials in electronics and optoelectronics due to their quantum size effect and novel physicochemical features compared with their bulk materials and other low-dimensional structured counterparts [1]. It is known that impurities play a crucial role in semiconductors by adjusting their electric conductivity and optical spectral characteristics [2,3,4]. To determine the impurity-bound states in core–shell nanostructures, the variational approach [5,6,7,8,9,10,11,12,13] with different forms of trial wavefunctions containing one [5,6,11], two [6,7,9,12,13], or three [8] variational parameters has been generally adopted for a coupled electron-impurity system with confinement potential. In addition to the variational approach, the finite-difference [14,15] and finite-element [16] numerical methods have also been used to directly solve the Schrödinger equation after the Coulombic potential was decoupled effectively as one. These works show that the binding energy [6,7,8,9,10,11,12,13,15,16,17,18], oscillator strength [8], and impurity-induced nonlinear optical properties like the optical absorption coefficient [8,14], refractive index change [8,14], photoionization cross section [9,13,16], and magnetic susceptibility [5,6,7,10] have great dependence on impurity position, structure dimension, composition, temperature, pressure, as well as external field (electric field, magnetic field, terahertz field, etc.). However, most authors have focused their attention on the impurity states in core–shell nanostructures composed of III-V compound semiconductors, and there still lacks adequate information related to the rapidly developing fourth-generation semiconductor material systems such as Ga2O3 and its sesquioxides.
Ga2O3 is a promising ultrawide band gap semiconductor with Eg ~4.9 eV for high frequency, high temperature, and high voltage applications in device domains such as power electronics [19,20], solar-blind deep-ultraviolet photodetectors [21,22], gas sensors [23,24], and so on. Among all of the polymorphs, β-Ga2O3 is the most stable modification. The band alignment between Ga2O3 and its ternary mixed crystal, (AlxGa1−x)2O3 or (InxGa1−x)2O3, is II type with nearly no valence band discontinuity and the conduction band offset can be tuned up to 1.7 eV through alloy composition. The Al composition of the as-grown pure phase (AlxGa1−x)2O3 can reach x < 0.71 for a monoclinic structure and x > 0.71 for a corundum structure, in spite of the solid phase miscibility gap between β-Ga2O3 and α-Al2O3 [25]. To date, some Ga2O3/(AlxGa1−x)2O3 low-dimensional structures like heterojunction [26,27], quantum well [28,29], and core–shell nanowire [30] have been fabricated using various vacuum techniques. Lyman and Krishnamoorthy [31] recently performed a theoretical investigation of optical intersubband transitions of electrons in β-Ga2O3/(AlxGa1−x)2O3 quantum well structures without considering the impurity effect. It was found that the electronic transition wavelength can be tuned from shortwave infrared (1–3 μm) to far infrared (>30 μm) wavebands in this kind of low-dimensional structure. Their work gave us the impetus to further study the hydrogen-like shallow impurity states in β-Ga2O3/(AlxGa1−x)2O3 core/shell nanostructures, where the electrons are confined in both the axial and radial directions. The binding energy as functions of structural dimension and aluminum composition when the impurity is located at different positions will be numerically computed using our previously developed algorithm that combines finite-difference approximation with a variational approach [11,12]. For comparison, both the hollow nanotube and solid nanorod core–shell structures, where the quantum confinement of electronic movement is differentiated in the core or shell layer, will be discussed in detail.

2. Theoretical Model

We studied a cylindrical core–shell nanostructure consisting of a Ga2O3 core layer and an (AlxGa1−x)2O3 shell layer. The schematic of the core/shell nanostructure is given in Figure 1. Two types of this three-dimensional confined nanostructure, named nanorod and nanotube, were considered in our calculation. The radii of the core and shell layer of the nanorod are defined as d2 and d3, respectively. The nanostructure comprised a hollow structure to form a double-walled nanotube, and we defined the radius of the hollow region as d1. The length of the nanorod or nanotube is defined as L. The z-axis is assumed to be along the nanostructure. Within the framework of effective mass approximation, the Hamiltonian of a conduction electron bound to a hydrogenic donor impurity shown in Figure 1b,c, bearing a charge e located at (ρ0, θ0, z0) in the Ga2O3/(AlxGa1−x)2O3 core/shell nanostructure, is written in a cylindrical coordinate system as follows:
H = 2 2 m * ( 2 ρ 2 + 1 ρ ρ + 1 ρ 2 2 θ 2 ) 2 2 m * 2 z 2 + V 1 ( ρ ) + V 2 ( z ) e 2 ε 0 ε r ρ 2 + ρ 0 2 2 ρ ρ 0 cos ( θ θ 0 ) + ( z z 0 ) 2
where m* is the electronic effective mass and εr is the static dielectric constant.
For a nanorod, V1(ρ) is the radial confinement potential given by:
V 1 ( ρ ) = { 0 , 0 < ρ < d 2 V 0 , d 2 ρ d 3 , ρ > d 3 ,
where V0 is the conduction band offset ∆Ec between Ga2O3 and (AlxGa1−x)2O3. For a nanotube, V1(ρ) is written as
V 1 ( ρ ) = { , 0 < ρ < d 1 0 , d 1 ρ d 2 V 0 , d 2 < ρ d 3 , ρ > d 3 .
V2(z) is the axial confinement potential given by:
V 2 ( z ) = { 0 , 0 z L , z > L .
In order to obtain the ground state energy of the electron-impurity bound system, the variational approach is used to solve the coupled Schrödinger equation:
H ψ ( ρ , z , θ ) = E ψ ( ρ , z , θ ) .
The two-parameter variational wave function is chosen as:
ψ ( ρ , z , θ ) = C e i m θ ϕ l ( ρ ) φ n ( z ) e α ρ 2 + ρ 0 2 2 ρ ρ 0 cos ( θ θ 0 ) e β ( z z 0 ) ,
in which C is the normalization constant of the wavefunction, and α and β are the variational parameters that account for both in-plane and z-axial correlations between the electron and the impurity. l and n are the quantum numbers related to the radial and z-axial relative motion of an electron, respectively. The angular moment quantum number m is taken as zero. The unbound electron states in the absence of impurities can be calculated using the method of separation of variables in the adiabatic approximation if the size difference of the core–shell nanostructure between the radial and axial directions is large. The boundary conditions at the interface and surface are determined by the Dirichlet boundary condition and continuity requirement. The radial wavefunction ϕl(ρ) and the z-axial wavefunction φn(z) have the exact forms based on Bessel and trigonometric functions (e.g., Refs. [5,6,7]). In this work, we utilized our previously developed algorithm based on the finite difference method [14,15] to deal with both the radial and z-axial Schrödinger equations to obtain the wavefunctions and energy levels. Computation time can be saved without solving the transcendental equations. However, the numerical error mainly caused by boundary truncation of wavefunctions and differential segmentation is somewhat larger than that obtained by the algorithm using the exact solutions.
The radial Schrödinger equation is given as follows:
[ 2 2 m * ( 2 ρ 2 + 1 ρ ρ ) + V 1 ( ρ ) ] ϕ l ( ρ ) = E l ϕ l ( ρ ) .
The radial Schrödinger equation in Equation (7) can be numerically solved using a finite-difference algorithm. First, the interval of ρ in the radial direction is divided into j parts, and thus the algebraic equations on the j + 1 nodes are solved simultaneously. The first-order central difference and second-order central difference formulas of the k-th node are given as:
d ϕ l , k ( ρ ) d ρ = 1 2 h ( ϕ l , k + 1 ϕ l , k 1 ) ,
and
d 2 ϕ l , k ( ρ ) d ρ 2 = 1 h 2 ( ϕ l , k + 1 2 ϕ l , k + ϕ l , k 1 ) .
In Equations (8) and (9), h is the step size. The position of the k-th node can be written as ρ = hk and V1(ρ) can be expressed as V1,k.
Therefore, the radial Schrödinger equation can be rewritten by substituting Equations (8) and (9) into Equation (7) in the finite-difference form, given as:
2 2 m * [ ( 1 h 2 1 2 k h 2 ) ϕ l , k 1 2 h 2 ϕ l , k + ( 1 h 2 + 1 2 k h 2 ) ϕ l , k + 1 ] + V 1 , k ϕ l , k = E l ϕ l , k .
The algebraic equations from the first to the j-th node can be written as:
2 2 m * [ 1 2 h 2 ϕ l , 0 2 h 2 ϕ l , 1 + 3 2 h 2 ϕ l , 2 ] + V 1 , 1 ϕ l , 1 = E l ϕ l , 1 2 2 m * [ 3 4 h 2 ϕ l , 1 2 h 2 ϕ l , 2 + 5 4 h 2 ϕ l , 3 ] + V 1 , 2 ϕ l , 2 = E l ϕ l , 2      
2 2 m * [ 2 j 3 ( 2 j 2 ) h 2 ϕ l , j 1 2 h 2 ϕ l , j + 2 j 1 ( 2 j 2 ) h 2 ϕ l , j + 1 ] + V 1 , j ϕ l , j = E l ϕ l , j .
The derivative of the wavefunction at j + 1 equals 0, and the wavefunction reaches a maximum at the center of the core–shell nanostructure. So, the central boundary condition can be treated using the Newton interpolation method. It is obtained as follows:
3 2 h ϕ l , 0 + 2 h ϕ l , 1 1 2 h ϕ l , 2 = 0 .
Equation (11) can then be written in matrix form as:
2 2 m * ( 4 h 2 4 h 2 1 2 h 2 2 h 2 3 2 h 2 3 4 h 2 2 h 2 5 4 h 2 2 j 3 ( 2 j 2 ) h 2 2 h 2 2 j 1 ( 2 j 2 ) h 2 2 j 1 ( 2 j 2 ) h 2 2 h 2 ) ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j )
+ ( V 1 , 1 V 1 , 2 V 1 , 3 V 1 , j - 1 V 1 , j ) ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) = E l ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) .
By using the matrix transformation D−1CD = T, in which:
D 1 = ( 1 2 2 4 2 6 2 2 j 1 ) ,
and
D = ( 1 1 / ( 2 2 ) 1 / 4 1 / 2 6 1 / ( 2 2 j 1 ) ) .
Equation (13) can be transformed as a j × j symmetric tridiagonal matrix
2 2 m * ( 4 h 2 2 h 2 2 h 2 2 h 2 3 2 4 h 2 3 2 4 h 2 2 h 2 5 3 6 h 2 2 j 3 2 ( j 3 ) 2 ( j 1 ) h 2 2 h 2 2 j 1 2 ( j 1 ) 2 j h 2 2 j 1 2 ( j 1 ) 2 j h 2 2 h 2 ) ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j )
+ ( V 1 , 1 V 1 , 2 V 1 , 3 V 1 , j - 1 V 1 , j ) ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) = E l ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) .
Finally, the energy level El can be obtained by solving the minimum eigenvalues and eigenvectors of the j × j symmetric tridiagonal matrix, and the radial wavefuntion ϕl(ρ) can be obtained by another matrix transformation:
( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) = ( 1 1 / ( 2 2 ) 1 / 4 1 / ( 2 6 ) 1 / ( 2 2 j 1 ) ) ( ϕ l , 1 ϕ l , 2 ϕ l , 3 ϕ l , j 1 ϕ l , j ) .
The z-axial Schrödinger equation is written as follows:
[ 2 2 m * 2 z 2 + V 2 ( z ) ] φ n ( z ) = E n φ n ( z ) .
It can be solved using another finite-difference algorithm, which is somewhat different from that used to solve the radial Schrödinger equation. The interval of z in the axial direction is divided into i parts, and thus the algebraic equations on the i + 1 nodes are solved simultaneously. The first-order central difference and second-order central difference formulas of the k-th node are given as:
d φ n , k ( z ) d z = 1 2 h ( φ n , k + 1 φ n , k 1 ) ,
d 2 φ n , k ( z ) d z 2 = 1 h 2 ( φ n , k + 1 2 φ n , k + φ n , k 1 ) .
Therefore, the axial Schrödinger equation can be rewritten by substituting Equations (19) and (20) into Equation (18) in the finite-difference form, given as:
2 2 m * [ 1 h 2 φ n , k 1 2 h 2 φ n , k + 1 h 2 φ n , k + 1 ] + V 2 , k φ n , k = E n φ n , k ,
in which z = hk and V2(z) can be expressed as V2,k. The boundary condition is very different from the radial Schrödinger equation. Since the outmost region of a core–shell nanostructure is assumed to be a vacuum, the z-axial wavefunction reaches 0 at the boundary, that is, φn,1 = 0 and φn,i+1 = 0.
Thus, the algebraic equations from the second to the i-th node can be written as:
2 2 m * [ 2 h 2 φ n , 2 + 1 h 2 φ n , 3 ] + V 2 , 2 φ n , 2 = E n φ n , 2 2 2 m * [ 1 h 2 φ n , 2 2 h 2 φ n , 3 + 1 h 2 φ n , 4 ] + V 2 , 3 φ n , 3 = E n φ n , 3 2 2 m * [ 1 h 2 φ n , 3 2 h 2 φ n , 4 + 1 h 2 φ n , 5 ] + V 2 , 4 φ n , 4 = E n φ n , 4      
2 2 m * [ 1 h 2 φ n , i 1 2 h 2 φ n , i ] + V 2 , i φ n , i = E n φ n , i .
Equation (22) can then be written in matrix form as:
2 2 m * ( 2 h 2 1 h 2 1 h 2 2 h 2 1 h 2 1 h 2 2 h 2 1 h 2 1 h 2 2 h 2 1 h 2 1 h 2 2 h 2 ) ( φ n , 2 φ n , 3 φ n , 4 φ n , i 1 φ n , i )
+ ( V 2 , 2 V 2 , 3 V 2 , 4 V 2 , i - 1 V 2 , i ) ( φ n , 2 φ n , 3 φ n , 4 φ n , i 1 φ n , i ) = E n ( φ n , 2 φ n , 3 φ n , 4 φ n , i 1 φ n , i ) .
Because the matrix in the above equation itself is an (i − 1) × (i − 1) symmetric tridiagonal matrix, the energy level En and z-axial wavefuntion φn(z) can be directly obtained by solving the minimum eigenvalues and eigenvectors of the (i − 1) × (i − 1) symmetric tridiagonal matrix without a matrix transformation.
Finally, the binding energy of an impurity can be solved by:
E b = E l + E n E D ,
where the expectation energy ED for a certain impurity state (l, n, m) can be derived by the energy minimization:
E D = min α , β ψ ( ρ , z , θ ) | H | ψ ( ρ , z , θ ) .
In the literature, most authors only calculated the binding energy of the impurity ground state with l = 0, n = 0. However, there are many states related to the quantum numbers l and n; we mainly take the lowest two states, i.e., l = 0 or 1 and n = 0 or 1, into account for simplicity. In other words, if we assume m = 0, four impurity states denoted as Ψ00, Ψ01, Ψ10, and Ψ11 are considered in our computation.

3. Numerical Results and Discussion

The related parameters of β-Ga2O3 and β-(AlxGa1−x)2O3 used in the computation are listed in Table 1. According to the references, the calculation was performed by only considering the difference in the effective mass between the inner layer Ga2O3 and the barrier material (AlxGa1−x)2O3. The conduction band offset between Ga2O3 and (AlxGa1−x)2O3 were taken as 2.15x + 0.94x2 eV for x < 0.5 and 0.96–0.24x + 1.95x2 eV for x > 0.5. The dielectric constants of Ga2O3 and (AlxGa1−x)2O3 were assumed to be 10. Without loss of generality, the length L of both the nanorod and nanotube was fixed at L = 30 nm and the radius of the outer shell layer was d3 = 20 nm. To obtain the binding energy of the impurity states, we first calculated the wavefunctions of the impurity-bound electron states.
Figure 2 illustrates the probability density in the ρ-z plane of the impurity states corresponding to Ψ00, Ψ01, Ψ10, and Ψ11 in a β-Ga2O3/(AlxGa1−x)2O3 core/shell nanorod and in a β-Ga2O3/(AlxGa1−x)2O3 core/shell nanotube. The impurity was assumed to be located at the position (d2/2, L/2) in the ρ-z plane in the nanorod and at the position (d2/2 + d1/2, L/2) in the nanotube, respectively. It is obvious that the distribution of the probability density features symmetry regarding z = L/2 (15 nm) for all of the impurity states in the two nanostructures. It also implies that the probability density has a circular symmetry for the four impurity states in the two kinds of nanostructures due to the structural geometry. As for the nanorod structure, the probability density related to Ψ00 is mainly distributed around the impurity and spreads over the core Ga2O3 layer. There are two or more distribution regions of the probability density for the excited states. The two parts of the probability density related to Ψ01 are distributed close to the core center, while the main probability density related to Ψ10 goes toward the interface, but this impurity state has a small part distributed near the core center. Most of the probability density of Ψ11 is distributed in the core region, but some is near the interface. As for the nanotube structure, the situation is quite different. The probability density concerning Ψ00 is mainly distributed around the impurity and closer to the interface between Ga2O3 and (AlxGa1−x)2O3. The probability density related to the three excited impurity states moves much closer to the interface, except Ψ01. The probability density related to Ψ01 is located more uniformly in the inner Ga2O3 layer. It is concluded that if the quantum number l > n, the probability densities of the impurity states change more prominently since the finite confinement potential at the ρ-axis is weaker than the confinement potential at the z-axis, which can be found in Equations (3) and (4). We also mention that tunneling behavior hardly occurs in the two nanostructures due to the large width of the confined core region as well as the impurity Coulombic interaction.
Figure 3 shows the change in impurity binding energy Eb with increasing core radius d2 for a core/shell nanorod and increasing hollow tube radius d1 for a core/shell nanotube. The order in terms of the magnitude of Eb for different impurity states is Ψ00 < Ψ01 < Ψ10 < Ψ11 in both nanostructures. It can be seen from Figure 3a that Eb decreases as the core radius d2 increases from 2 to 7 nm, which can be attributed to the weaker quantum confinement. This variation behavior is very similar to that in Ref. [12]. The influence of the impurity position (at the interface or in the core region) on Eb does not seem to be very prominent. It has a stronger impact on the binding energy corresponding to the ground state, Ψ00. The binding energy when the impurity is at the interface is lower than that when the impurity appears in the core region. If the impurity is not at the center ρ = 0, the trend in the binding energy of the ground state goes toward zero. If the impurity is at the center, the binding energy will reach the bulk value. In the contrast, one can see from Figure 3b that Eb increases as the hollow tube radius d1 of the core/shell nanotube increases. In other words, the inner layer of the double walls of the nanotube become narrower, thus enhancing the quantum confinement of the impurity states. The impurity position has a more obvious influence on the binding energy, especially related to the excited impurity states, in a nanotube than in a nanorod. The difference in Eb between two impurity positions is enhanced when d1 is larger. In the extreme case, the ground state binding energy will reach zero if d1 goes to zero. It needs to be pointed out that the binding energy corresponding to the two excited states, Ψ10 and Ψ11, decreases when the impurity moves from the inner layer to the interface between the two walls of the nanotube.
Next, we turn to discuss the influence of Al composition on the binding energy of the impurity located at different positions in the two kinds of nanostructures. It can be observed from Figure 4 that an inflection appears especially for the excited impurity states, which can be attributed to the conduction band offset between Ga2O3 and (AlxGa1−x)2O3 at x = 0.5 [22]. As the Al composition increases, the binding energy also increases because the probability density becomes more located in the inner layer, which is confined by a higher potential barrier. This increment is less significant for the Ψ00 and Ψ01 states. Moreover, the binding energy decreases if the impurity moves towards the interface. The binding energy for the Ψ10 and Ψ11 states in a nanotube structure is somewhat different, which can be also seen in Figure 3b, possibly because of the quantum tunneling towards the (AlxGa1−x)2O3 barrier layer. If the radial dimension is smaller than the electronic mean free path, the quantum tunneling at the tubular interface will make a greater contribution to the binding energy of the off-center hydrogen-like impurity in a core/shell nanostructure. However, the phase transition from β to α phase and the strain effect, which may both alter energy band diagrams at high compositions approaching x = 0.7, were not taken into account in this paper. These will be carefully considered in future work. We point out that degenerate states with different radial and axial quantum numbers, which were always neglected in most of the literature, should be considered in some quantum confined structures like rods, tubes, rings, ribbons, disks, and so forth.

4. Conclusions

In conclusion, we have made a comparison of the impurity states in a coaxial core/shell nanorod and coaxial core/shell nanotube composed of the β-Ga2O3/(AlxGa1−x)2O3 material system. The probability density and binding energy of an off-center hydrogenic donor impurity with regards to four impurity states, namely Ψ00, Ψ01, Ψ10, and Ψ11, were numerically calculated using the finite-difference method combined with a variational approach in the framework of effective mass approximation and single electron approximation. Due to the synergic mechanism of nanostructured quantum confinement and Coulombic interaction from the impurity, the quantum localization maps of the four impurity-bound electron states were quite different for core/shell nanorods and nanotubes. As a consequence, varying the Al composition and geometrical parameters makes it possible to modify the density of the electron states and the binding energy of impurities in these kinds of core/shell nanostructures. Comparatively, the impurity binding energies corresponding to the excited states were larger than the ground state binding energy and more sensitive to the structural dimension and Al composition. If the radial dimension is very small (smaller than the electronic mean free path), the interface impurity plays a more important role in determining the electronic properties due to the quantum tunneling at the tubular interface, especially in a nanotube structure. It is believed that our results will help to measure the electrical transport and electroluminescence properties of rapidly developing Ga2O3/(AlxGa1−x)2O3 low-dimensional nanostructures for device applications.

Author Contributions

S.H.: Methodology, Software, Data curation, Visualization, Validation, Writing—review & editing. J.Z.: Conceptualization, Writing-original draft, Supervision, Investigation, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 12164031) and the Science Foundation of Inner Mongolia Autonomous Region (Grant No. 2020MS06007 and 2023LHMS01004). The APC was funded by the National Natural Science Foundation of China (Grant No. 12164031).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zou, H.; Luo, Z.; Yang, X.; Xie, Q.; Zhou, Y. Toward emerging applications using core-shell nanostructured materials: A review. J. Mater. Sci. 2022, 57, 10912. [Google Scholar] [CrossRef]
  2. Dakhlaoui, H.; Nefzi, M. Simultaneous effect of impurities, hydrostatic pressure, and applied potential on the optical absorptions in a GaAs field-effect transistor. Res. Phy. 2019, 15, 102618. [Google Scholar] [CrossRef]
  3. Wang, W.; Xu, L.; Wei, X.; Zhang, S.; Yao, Z. The effects of hydrostatic pressure and temperature on the nonlinear optical properties of shallow-donor impurities in semiconductors in a magnetic field. J. Appl. Phys. 2020, 127, 195903. [Google Scholar] [CrossRef]
  4. Wang, W.; Xu, L.; Wei, X.; Zhang, S. Intense-terahertz-laser modulated photoionization cross section of shallow-donor impurity in semiconductors in a magnetic field. Res. Phy. 2021, 20, 103692. [Google Scholar] [CrossRef]
  5. Portacio, A.A.; Rodríguez, B.A.; Villamil, P. Influence of the position of a donor impurity on the second-order nonlinear optical susceptibility in a cylindrical quantum dot. Superlatt. Microstruct. 2018, 113, 550. [Google Scholar] [CrossRef]
  6. Sherly, I.J.; Lidiya, A.E.; Nithiananthi, P. Tailoring the electronic properties of concentric double quantum rings in the presence of a donor impurity. J. Appl. Phys. 2019, 125, 165707. [Google Scholar] [CrossRef]
  7. El-Bakkari, K.; Sali, A.; Iqraoun, E.; Ezzarfi, A. Polaron and conduction band non-parabolicity effects on the binding energy, diamagnetic susceptibility and polarizability of an impurity in quantum rings. Superlatt. Microstruct. 2020, 148, 106729. [Google Scholar] [CrossRef]
  8. Ghukasyan, T.K.; Vartanian, A.L. Polaronic effects on the impurity-related linear and nonlinear optical properties in a nanowire with magnetic field. Superlatt. Microstruct. 2020, 137, 106339. [Google Scholar] [CrossRef]
  9. Shi, L.; Yan, Z.W.; Meng, M.W. Binding energy and photoionization cross section of hydrogenic impurities in elliptic cylindrical core/shell quantum dots under a non-axial electric field. Superlatt. Microstruct. 2021, 150, 106818. [Google Scholar] [CrossRef]
  10. Maouhoubi, I.; En-nadir, R.; Zorkani, I.; Hassani, A.O.T.; Jorio, A. The effects of the dielectric screening, temperature, magnetic field, and the structure dimension on the diamagnetic susceptibility and the binding energy of a donor-impurity in quantum disk. Phys. B 2022, 646, 414371. [Google Scholar] [CrossRef]
  11. Bilekkaya, A. The electronic properties of coaxial triangular quantum well wires. Phys. Lett. A 2022, 450, 128389. [Google Scholar] [CrossRef]
  12. Hbibi, M.; Mommadi, O.; Chouef, S.; Boussetta, R.; Belamkadem, L.; El Moussaouy, A.; Falyouni, F.; Duque, C.M.; Vinasco, J.A.; Duque, C.A. Finite confinement potentials, core and shell size effects on excitonic and electron-atom properties in cylindrical core/shell/shell quantum dots. Sci. Rep. 2022, 12, 14854. [Google Scholar] [CrossRef] [PubMed]
  13. El-Yadri, M.; El Hamdaoui, J.; Aghoutane, N.; Pérez, L.M.; Baskoutas, S.; Laroze, D.; Díaz, P.; Feddi, E.M. Optoelectronic properties of a cylindrical core/shell nanowire: Effect of quantum confinement and magnetic field. Nanomaterials 2023, 13, 1334. [Google Scholar] [CrossRef] [PubMed]
  14. Ha, S.H.; Zhu, J. Hydrogenic impurity-related optical properties in a piezoelectric core-shell nanowire. J. Appl. Phys. 2020, 128, 034305. [Google Scholar] [CrossRef]
  15. Ha, S.H.; Zhu, J. Temperature effect on shallow impurity states in a wurtzite GaN/AlxGa1−xN core-shell nanowire. Phys. E 2020, 122, 114179. [Google Scholar] [CrossRef]
  16. Fakkahi, A.; Jaouane, M.; Limame, K.; Sali, A.; Kirak, M.; Arraoui, R.; Ed-Dahmouny, A.; El-bakkari, K.; Azmi, H. The impact of the electric field on the photoionization cross section, polarizability, and donor impurity binding energy in multilayered spherical quantum dot. Appl. Phys. A 2023, 129, 188. [Google Scholar] [CrossRef]
  17. Solaimani, M. Stark shift of binding energy for on and off-center donor impurities in quantum rings under the influence of charged rods electric fields. Solid State Sci. 2020, 108, 106386. [Google Scholar] [CrossRef]
  18. Belamkadem, L.; Mommadi, O.; Boussetta, R.; Chouef, S.; Chnafi, M.; El Moussaouy, A.; Vinasco, J.A.; Laroze, D.; Duque, C.A.; Kenfack-Sadem, C.; et al. The intensity and direction of the electric field effects on off-center shallow-donor impurity binding energy in wedge-shaped cylindrical quantum dots. Thin Solid Films 2022, 757, 139396. [Google Scholar] [CrossRef]
  19. Qin, Y.; Wang, Z.; Sasaki, K.; Ye, J.; Zhang, Y. Recent progress of Ga2O3 power technology: Large-area devices, packaging and applications. Jpn. J. Appl. Phys. 2023, 62, SF0801. [Google Scholar] [CrossRef]
  20. Prasad, C.V.; Rim, Y.S. Review on interface engineering of low leakage current and on-resistance for high-efficiency Ga2O3-based power devices. Mater. Today Phys. 2022, 27, 100777. [Google Scholar] [CrossRef]
  21. Kaur, D.; Kumar, M. A strategic review on gallium oxide based deep-ultraviolet photodetectors: Recent progress and future prospects. Adv. Opt. Mater. 2021, 9, 2002160. [Google Scholar] [CrossRef]
  22. Hou, X.; Zou, Y.; Ding, M.; Qin, Y.; Zhang, Z.; Ma, X.; Tan, P.; Yu, S.; Zhou, X.; Zhao, X.; et al. Review of polymorphous Ga2O3 materials and their solar-blind photodetector applications. J. Phys. D Appl. Phys. 2021, 54, 043001. [Google Scholar] [CrossRef]
  23. Zhu, J.; Xu, Z.; Ha, S.; Li, D.; Zhang, K.; Zhang, H.; Feng, J. Gallium oxide for gas sensor applications: A comprehensive review. Materials 2022, 15, 7339. [Google Scholar] [CrossRef]
  24. Zhai, H.; Wu, Z.; Fang, Z. Recent progress of Ga2O3-based gas sensors. Ceram. Inter. 2022, 48, 24213. [Google Scholar] [CrossRef]
  25. Li, H.; Wu, Z.; Wu, S.; Tian, P.; Fang, Z. (AlxGa1−x)2O3-based materials: Growth, properties, and device applications. J. Alloys Comp. 2023, 960, 170671. [Google Scholar] [CrossRef]
  26. Ranga, P.; Bhattacharyya, A.; Chmielewski, A.; Roy, S.; Sun, R.; Scarpulla, M.A.; Alem, N.; Krishnamoorthy, S. Growth and characterization of metalorganic vapor-phase epitaxy-grown β-(AlxGa1−x)2O3/β-Ga2O3 heterostructure channels. Appl. Phys. Exp. 2021, 14, 025501. [Google Scholar] [CrossRef]
  27. Saha, C.N.; Vaidya, A.; Singisetti, U. Temperature dependent pulsed IV and RF characterization of β-(AlxGa1−x)2O3/Ga2O3 hetero-structure FET with ex situ passivation. Appl. Phys. Lett. 2022, 120, 172102. [Google Scholar] [CrossRef]
  28. Kneiß, M.; Storm, P.; Hassa, A.; Splith, D.; von Wenckstern, H.; Lorenz, M.; Grundmann, M. Growth, structural and optical properties of coherent κ-(AlxGa1−x)2O3/κ-Ga2O3 quantum well superlattice heterostructures. APL Mater. 2020, 8, 051112. [Google Scholar] [CrossRef]
  29. Lee, H.; Wang, H.; Lee, C. Whole metal oxide-based deep ultraviolet LEDs using Ga2O3-Al2O3: Ga2O3 multiple quantum wells. IEEE Photon. Technol. Lett. 2022, 34, 621. [Google Scholar] [CrossRef]
  30. Jiang, T.; Qiu, Y.; Tao, J.; Xiao, X. Ga2O3@Al2O3 core-shell nanowires for high-performance solar-blind photodetectors. ACS Appl. Nano Mater. 2023, 6, 9849. [Google Scholar] [CrossRef]
  31. Lyman, J.E.; Krishnamoorthy, S. Theoretical investigation of optical intersubband transitions and infrared photodetection in β-(AlxGa1−x)2O3/Ga2O3 quantum well structures. J. Appl. Phys. 2020, 127, 173102. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic representation of a β-Ga2O3/(AlxGa1−x)2O3 core/shell nanostructure. (b) Schematic of the radial confinement potential of a core/shell nanorod. (c) Schematic of the radial confinement potential V1(ρ) of a core/shell nanotube. d1, d2, and d3 denote the hollow tube radius, the core radius, and the shell radius. ∆Ec denotes the conduction band offset between Ga2O3 and (AlxGa1−x)2O3. O is the coordinate origin and ‘⨁’ denotes the impurity located at different positions.
Figure 1. (a) Schematic representation of a β-Ga2O3/(AlxGa1−x)2O3 core/shell nanostructure. (b) Schematic of the radial confinement potential of a core/shell nanorod. (c) Schematic of the radial confinement potential V1(ρ) of a core/shell nanotube. d1, d2, and d3 denote the hollow tube radius, the core radius, and the shell radius. ∆Ec denotes the conduction band offset between Ga2O3 and (AlxGa1−x)2O3. O is the coordinate origin and ‘⨁’ denotes the impurity located at different positions.
Crystals 13 01227 g001
Figure 2. Probability densities in the ρ-z plane of the impurity states (a) Ψ00, (c) Ψ01, (e) Ψ10, and (g) Ψ11 in a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanorod with d2 = 8 nm where the impurity is located at the position (d2/2, L/2), and the impurity states (b) Ψ00, (d) Ψ01, (f) Ψ10, and (h) Ψ11 in a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanotube with d1 = 3 nm and d2 = 8 nm where the impurity is located at the position (d2/2 + d1/2, L/2). Note that only a half of the cross-section is plotted because of the symmetry.
Figure 2. Probability densities in the ρ-z plane of the impurity states (a) Ψ00, (c) Ψ01, (e) Ψ10, and (g) Ψ11 in a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanorod with d2 = 8 nm where the impurity is located at the position (d2/2, L/2), and the impurity states (b) Ψ00, (d) Ψ01, (f) Ψ10, and (h) Ψ11 in a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanotube with d1 = 3 nm and d2 = 8 nm where the impurity is located at the position (d2/2 + d1/2, L/2). Note that only a half of the cross-section is plotted because of the symmetry.
Crystals 13 01227 g002
Figure 3. Impurity binding energy Eb as a function of (a) core radius d2 for a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanorod and (b) hollow tube radius d1 for a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanotube. The black, pink, green, and purple lines denote the contributions of the impurity states Ψ00, Ψ01, Ψ10, and Ψ11, respectively. The solid and dashed lines denote the impurity located at the position (d2/2, L/2) and (d2, L/2) for a nanorod and at the position (d2/2 + d1/2, L/2) and (d2, L/2) for a nanotube, respectively.
Figure 3. Impurity binding energy Eb as a function of (a) core radius d2 for a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanorod and (b) hollow tube radius d1 for a β-Ga2O3/(Al0.3Ga0.7)2O3 core/shell nanotube. The black, pink, green, and purple lines denote the contributions of the impurity states Ψ00, Ψ01, Ψ10, and Ψ11, respectively. The solid and dashed lines denote the impurity located at the position (d2/2, L/2) and (d2, L/2) for a nanorod and at the position (d2/2 + d1/2, L/2) and (d2, L/2) for a nanotube, respectively.
Crystals 13 01227 g003
Figure 4. Impurity binding energy Eb as a function of the aluminum composition x for a β-Ga2O3/(AlxGa1−x)2O3 core/shell (a) nanorod with d2 = 8 nm and (b) nanotube with d1 = 3 nm and d2 = 8 nm. The black, pink, green, and purple lines denote the contributions of the impurity states Ψ00, Ψ01, Ψ10, and Ψ11, respectively. The solid and dashed lines denote the impurity located at the position (d2/2, L/2) and (d2, L/2) for a nanorod and at the position (d2/2 + d1/2, L/2) and (d2, L/2) for a nanotube, respectively.
Figure 4. Impurity binding energy Eb as a function of the aluminum composition x for a β-Ga2O3/(AlxGa1−x)2O3 core/shell (a) nanorod with d2 = 8 nm and (b) nanotube with d1 = 3 nm and d2 = 8 nm. The black, pink, green, and purple lines denote the contributions of the impurity states Ψ00, Ψ01, Ψ10, and Ψ11, respectively. The solid and dashed lines denote the impurity located at the position (d2/2, L/2) and (d2, L/2) for a nanorod and at the position (d2/2 + d1/2, L/2) and (d2, L/2) for a nanotube, respectively.
Crystals 13 01227 g004
Table 1. Material parameters used in the computation [31]. Copyright 2020, AIP Publishing.
Table 1. Material parameters used in the computation [31]. Copyright 2020, AIP Publishing.
Material parametersβ-Ga2O3β-(AlxGa1−x)2O3
Effective mass m* (m0)0.280.28 + 0.11x
Band gap Eg (eV)4.694.69 + 1.34x + x2
Dielectric constant εr (ε0)1010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ha, S.; Zhu, J. Hydrogen-like Impurity States in β-Ga2O3/(AlxGa1−x)2O3 Core/Shell Nanostructures: Comparison between Nanorods and Nanotubes. Crystals 2023, 13, 1227. https://doi.org/10.3390/cryst13081227

AMA Style

Ha S, Zhu J. Hydrogen-like Impurity States in β-Ga2O3/(AlxGa1−x)2O3 Core/Shell Nanostructures: Comparison between Nanorods and Nanotubes. Crystals. 2023; 13(8):1227. https://doi.org/10.3390/cryst13081227

Chicago/Turabian Style

Ha, Sihua, and Jun Zhu. 2023. "Hydrogen-like Impurity States in β-Ga2O3/(AlxGa1−x)2O3 Core/Shell Nanostructures: Comparison between Nanorods and Nanotubes" Crystals 13, no. 8: 1227. https://doi.org/10.3390/cryst13081227

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop