Next Article in Journal
Development of TiO2 Nanosheets with High Dye Degradation Performance by Regulating Crystal Growth
Next Article in Special Issue
The Effect of Periodic Duty Cyclings in Metal-Modulated Epitaxy on GaN:Mg Film
Previous Article in Journal
Effect of the Sorption Layer on the Protection Time Provided by Anti-Smog Half-Masks
Previous Article in Special Issue
The Effect of Nitridation on Sputtering AlN on Composited Patterned Sapphire Substrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polarization Doping in a GaN-InN System—Ab Initio Simulation

1
Institute of High Pressure Physics, Polish Academy of Sciences, Sokolowska 29/37, 01-142 Warsaw, Poland
2
Research Institute for Applied Mechanics, Kyushu University, Fukuoka 816-8580, Japan
3
Institute of Applied Mathematics and Mechanics, University of Warsaw, 02-097 Warsaw, Poland
4
Institute of Physics, Polish Academy of Sciences, Aleja Lotnikow 32/46, 02-668 Warsaw, Poland
5
University Grenoble-Alpes, CEA, Grenoble INP, IRIG, PHELIQS, 17 av. des Martyrs, 38000 Grenoble, France
6
Faculty of Mathematics and Natural Sciences, School of Exact Sciences, Cardinal Stefan Wyszynski University, Dewajtis 5, 01-815 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Materials 2023, 16(3), 1227; https://doi.org/10.3390/ma16031227
Submission received: 16 December 2022 / Revised: 25 January 2023 / Accepted: 30 January 2023 / Published: 31 January 2023
(This article belongs to the Special Issue III-V Semiconductor Optoelectronics: Materials and Devices)

Abstract

:
Polarization doping in a GaN-InN system with a graded composition layer was studied using ab initio simulations. The electric charge volume density in the graded concentration part was determined by spatial potential dependence. The emerging graded polarization charge was determined to show that it could be obtained from a polarization difference and the concentration slope. It was shown that the GaN-InN polarization difference is changed by piezoelectric effects. The polarization difference is in agreement with the earlier obtained data despite the relatively narrow bandgap for the simulated system. The hole generation may be applied in the design of blue and green laser and light-emitting diodes.

1. Introduction

Wurtzite crystalline symmetry permits emergence of a macroscopic vectorial quantity known as polarization [1]. Macroscopically, polarization occurs due to relative translation of the center of the negative electrons with respect to the position of positive atomic nucleus, i.e., creation of electric dipole density [1]. This is associated directly with the iconicity of metal–nonmetal bonding. Nitride semiconductors such as GaN, AlN and InN have different fractions of the covalent and ionic components in their bonding. Therefore, the chemically induced charge effect is different, which leads to a difference in their spontaneous polarization. In addition, external strain can downgrade the symmetry of the lattice, which could induce or change the polarization charge [2,3].
Polarization affects physical properties of semiconductor systems by the emergence of the electric fields of various magnitudes and ranges. In large size systems, the large scale electric fields are negligible due to screening which is described in Debye-Hückel or Thomas-Fermi approximation [3,4,5]. A much stronger influence of polarization-related electric fields is observed in nanometer size systems. A glaring, positive example is the localization of electrons in the GaN-based field-effect transistor (FET) [6]. For III-nitride multi-quantum well (MQWs) laser diodes (LDs) and light-emitting diode (LED) performance, these fields are highly detrimental. The electron–hole wavefunction overlap and accordingly the radiative recombination dipole strength is reduced by the so called quantum-confined Stark effect (QCSE) [7,8,9]. The efforts to avoid the fields in the devices using nonpolar MQWs were not successful due to the poor material quality in such structures [10].
In devices containing heterostructures, polarization difference entails a sheet charge [11,12] and a surface dipole layer [12] at the heterointerfaces. The sheet charge is equivalent to the electric field difference across the interface. The phenomenon is well understood; however the magnitude is still under debate [13,14,15]. At the dipole layer region, the electric potential jumps between both sides of the interface. The effect was identified a long time ago, but it is still disputed [16].
Chemically induced electric phenomena are not limited to sharp interfaces. The transition between two chemically different substances may be also more or less diffuse. Predictably, the diffuse interface generates the extended charge density [17,18,19,20,21,22,23]. The nitride alloys concentration change leads to the polarization difference and creation of a bulk charge density, inducing a phenomenon known as polarization doping [18,19,20,21,22,23]. Since the polarization is proportional to metal concentration, linearly graded materials have a uniform bulk charge density. In insulating systems, without screening the potential profile, it can be used to determine the polarization-related charge density via the Poisson equation. The method was successfully used to determine the polarization density constant for the Al-Ga-N system in [23]. In addition, the mechanism for the generation of the mobile charge was described and used to design the UV laser diode (LD).
The development of blue nitride-based lasers was one of the most spectacular achievements, in terms of scientific achievements, development of technology and contribution to society [24,25]. Unfortunately, the extension of the emission spectrum towards UV [26,27] and green [28,29] light emission encounters considerable difficulties. The problem of p-type doping is one of the roadblocks towards high power green UV lasers which have many promising applications. The best possible remedy is to use Mg and polarization doping simultaneously [30]. In this work, we will use the methods developed in [23] for the Ga-In-N system to obtain polarization-doping constants, verify the creation of mobile charge and its application to device design.

2. Methods

Ab initio simulations reported in this paper employed the density functional theory (DFT) package SIESTA. SIESTA uses numeric atomic orbital series to solve Kohn–Sham equations [23,24]. The following atomic orbitals were used: for In—3s: DZ (double zeta), 3p: TZ (triple zeta), 3d: DZ; N—2s: DZ, 2p: TZ, 3d: SZ (single zeta) and for Ga- 4s: DZ (double zeta), 4p: TZ (triple zeta), 3d: DZ. The pseudopotentials approximation was used for all atoms which allowed us to increase the number of atoms and consequently, the size of supercell used. The pseudopotentials were generated by the authors using the ATOM program designed for all-electron calculations. In SIESTA the norm-conserving Troullier–Martins pseudopotentials, in the Kleinmann–Bylander factorized form are used [31,32,33,34]. The calculations were performed within Generalized Gradient Approximation (GGA) with the PBEsol modified version of the original Perdew, Burke and Ernzerh of the exchange–correlation functional [35,36].
The lattice parameters of bulk indium nitride, calculated for a periodic infinite crystal, are a I n N D F T = 3.525   and c I n N D F T = 5.716   . They are in sufficient agreement with the experimental data for wurtzite bulk InN obtained from X-ray diffraction measurements: a I n N e x p = 3.537   and c I n N e x p = 5.703   [37]. Ab initio calculations of GaN gave a G a N D F T = 3.198   and c G a N D F T = 5.199   , again showing agreement with the X-ray data a G a N e x p = 3.1890   and c G a N e x p = 5.1864   [38]. PBEsol approximation provides an erroneous result for band structure in which the gap is closed. Therefore, the band diagrams were plotted from calculations using Ferreira et al. GGA-1/2 approximation [39,40]. It provides effective masses, bandgap (BG) energies, and more generally, the band structures are in agreement with experimental data [40]. The GGA-1/2 approximation calculations were used for the positions of atoms and a periodic cell parameter relaxed to equilibrium in GGA-PBEsol approximation. All atom positions were changed to reduce the single atom forces at a level below 0.005 eV/Å. The bandgap of the InN crystal was: E g D F T ( I n N ) = 0.75   e V . The bandgap of InN was the subject of initial controversy. Finally, the experimental data of several papers established InN gap value as ( E g e x p ( I n N ) = 0.65   e V ) [41,42,43]. The ab initio gap value of GaN was E g D F T ( G a N ) = 3.51   e V . This could be compared to the well-established experimental GaN bandgap E g e x p ( G a N ) = 3.47 e V [44,45]. A convergence criterion for a self-consistent field (SCF) loop termination was that the maximum difference between the output and the input for any element of the density matrix had to be below 10−4.

3. Results

Ab initio calculations were used to simulate the In-Ga-N supercell composed of the three chemically different regions, arranged along the c-axis, i.e., along [0001] direction: (i) uniform GaN layer, (ii) uniform In0.5G0.5N layer, (iii) graded InxGa1−xN layer ( 0.0 x 0.5 ) . This is a direct implementation of the calculation method successfully used in determination of polarization doping in an AlN-GaN system [23]. As in the standard supercell DFT calculations, the periodic boundary conditions are imposed on all boundaries of the supercell. The atomic arrangement in the supercell is (Figure 1):
(i)
Four Ga-N double atomic layers (metal and nonmetal layers);
(ii)
Four In0.5Ga0.5N double atomic layers (metal and nonmetal layers);
(iii)
Sixteen linearly graded Ga-In-N double atomic layers, with Ga content increasing or decreasing along [0001] direction (metal layers composed of Ga and In, nonmetal layer composed of N).
In the ternary alloy section, the metal atoms are distributed randomly within each layer independently. The number of In atoms in the neighboring layers changes every two layers. Thus, the number of the In and Ga atoms is set in the layer so that average concentration is changed along the z-axis linearly (Figure 2). The lattice positions of In and Ga atoms within the layer are generated based on a random number of generators employed in many Monte Carlo simulations. Therefore, the selection enforces uniform sampling in the lattice sites, i.e., the method not prone to systematic errors, generating a relatively high level of numeric noise. The potential profiles presented in Figure 3 confirm that the random factor is not important. This remarkable effect results from a long range of Coulomb interactions that smooth out the potential profile. In the effect, the potential profile does not show any considerable local variations related to InGaN configuration. The selected positions of In and Ga atoms in the neighboring layers are uncorrelated. In addition, several configurations were used to obtain the averaged quantities.
The averaged composition of the atomic layers is presented in Figure 2. The graded region consist of eight unit cells in which the In concentration changed from 0 to 0.5. Therefore, the average concentration change across the whole graded layer, i.e., at the approximate distance of 8 c   = 4.3612   nm was Δ x = 0.5 . The resulting concentration slope is s c = d x d z = 0.115   nm 1 .
The polarization is expected to vary linearly with the In and Ga concentrations. The polarization change should give a bulk charge density ρ ( r ) , which is related to the electrical potential φ ( r ) in accordance with Poisson equation:
( ε b ( r ) φ ( r ) ) = ρ p d ( r ) ε o
where ε b is the static dielectric constant of the semiconductor and ε o dielectric permittivity. The dielectric constants of InN and GaN are different therefore, the Vegard law is used for graded region ε b ( x ) = x   ε b I n N + ( 1 x )   ε b G a N , where ε b I n N = 14.61 and ε b G a N = 10.28 [46]. Using this approximation, the variation of the dielectric constant along z coordinate in the graded region can be represented as ε b ( z ) = q 1 + q 2 z , where these parameters are:
q 1 = ε b G a N ,   q 2 = ( ε b G a N ε b I n N )   s c
q 1 = ( ε b G a N + ε b I n N ) / 2   ,   q 2 = ( ε b G a N ε b I n N )   s c
The electric potential profile was obtained by exact analytical solution of Equations (1) and (2) within the graded region to obtain
φ ( z )   =   a 0 + a 1 log ( q 1 + q 2 z ) ρ p d ε o q 2 z   =   a 0 + a 1 log ( q 1 + q 2 z ) + a 2 z
where constants a0 and a1 depend on the boundary conditions, and a2 depends on polarization charge ρpd, electric constant ε0 and parameter q1. In this case, boundary conditions are not used, as constants a0, a1 and a2 are fitted to the potential obtained from the profile that is averaged in the plane perpendicular to c-axis, derived from ab initio calculations. The plane averaged potential is highly oscillatory due to maxima from atomic core charges. The basic procedure is smoothing by adjacent averaging over the c lattice parameter period. Still, the potential oscillations cannot be completely averaged out due to the change of the lattice constant in the graded region.
On the other hand, in non-graded regions, parameter εb is constant and we obtain a linear potential, as the averaged bulk charge is zero. The obtained potential profiles corresponding to the supercell presented in Figure 1a are shown in Figure 3.
The segments of the profile obtained in the uniform GaN, uniform In0.5Ga0.5N and linearly graded InxGa1−xN have distinctively different space dependence. The sections with uniform concentration are linear, whereas the potential profile within the linearly graded alloy are parabolic, typical for the systems having uniform bulk charge density. In addition, at the In0.5Ga0.5N/GaN interface, a potential jump is observed, due to the emergence of a dipole layer. The two different averaging procedures provided slightly different profiles: more or less oscillatory in the uniform and graded regions and the different potential jump at the In0.5Ga0.5N/GaN interface.
As it was shown, the potential profile in the graded region closely follows the solution (3) confirming the assumption of uniform charge density in this region. The following parameters were obtained: a 0 = 134   V , a 1 = 52.7   V and a 2 = 2.173 V nm . The bulk charge density derived via Equations (2a) and (3) is: ρ pd = 9.55 × 10 6 C m 3 (i.e., 5.96 × 1019 cm−3). From the geometric data, the thickness of the graded region is d grad 4.3612   nm . Thus, the obtained surface sheet charge density is ρ s u r f = 4.165 × 10 2 C m 2 .
In the uniform regions, the potential profiles are linear, confirming the absence of the bulk charge density. From the linear approximation, the electric field can be derived. The following field values were obtained: E G a N ( a ) = φ = 0.287   V / nm = 2.87 × 10 2 V / and E G a I n N ( a ) = φ = 0.487   V / nm = 4.87 × 10 2   V / . The sheet charge density at the interfaces obtained via Gauss law ( n —unit normal vector, perpendicular to the interface):
ρ s = ε o [ ( ε G a N + ε I n N 2 ) E G a I n N ε G a N E G a N ]   n = ( P G a I n N P G a N )   n
is ρ s = 7.59 × 10 2 C / m 2 . This has to be compared with the estimate obtained above by integration of bulk charge over the thickness of the graded region which was ρ s u r f = 4.165 × 10 2 C / m 2 . The difference is due to interfacial charges present on the interfaces between the graded layer and neighboring layers, which do not contribute to the spatial polarization charge.
Similar analysis could be made for the case presented in Figure 1b. The obtained potential profiles, corresponding to the supercell presented in Figure 1b are shown in Figure 4. From the fit to Formula (3), the following parameters were obtained: a 0 = 250   V , a 1 = 103   V / nm and a 2 = 4.73   V / nm 2 . The charge density derived via Equation (3) is: ρ p d = 2.08 × 10 7 C / m 3 (i.e., 1.3 × 1020 cm−3). From the geometric data, the thickness of graded region is d g r a d 4.3612   nm . Thus, the obtained surface sheet charge density is ρ s u r f = 9.065 × 10 2 C m 2 .
In the uniform regions, the potential profiles are linear, which confirms the absence of the bulk charge density. The following field values were obtained from linear approximations: E G a N ( a ) = φ = 0.353 V / nm = 3.53 × 10 2 V / and E G a I n N ( a ) = φ = 0.309 V / nm = 3.09 × 10 2 V / . The surface charge density at the interfaces obtained from Gauss law (Equation (4)) is ρ s = 6.62 × 10 2 C / m 2 . This has to be compared with the estimate obtained above by integration of bulk charge over the thickness of the graded region which was is ρ s u r f = 9.07 × 10 4 C / m 2 . Again, the difference is due to interfacial charges present on the interfaces between the graded layer and neighboring layers.
In summary, the surface density is obtained in case (a) from the bulk density ρ s u r f = 4.165 × 10 2 C m 2 . In the second case, this value is ρ s u r f = 9.065 × 10 2 C m 2 . The discrepancy between case (a) and (b) may be related to the fact that at the Ga0.5In0.5N/GaxIn1−xN system bandgap is small so the mobile band charge may affect the obtained potential profile and the resulting polarization doping bulk charge. Therefore, the polarization doping bulk charge obtained is ρ p d = 9.55 × 10 6 C m 3 and ρ p d = 2.079 × 10 7 C m 3 in case (a) and (b) respectively. This could be translated into the elementary charge density (a) n = 5.97   ×   10 19   cm 3 and (b) p = 1.3   ×   10 20   cm 3 . From these, the values of polarization doping parameters Q p o l could be obtained as:
Q p o l =   ρ p d e   s c = Δ P e
From these calculations, the values of polarization doping parameters are: (a) Q p o l = 0.52     nm 2 and (b) Q p o l = 1.13     nm 2 . Thus these values can be used to calculate the InN-GaN polarization difference as: (a) Δ P G a N I n N =   P G a N   P I n N =   Q p o l   e = 0.083   C m 2 and (b) Δ P G a N I n N =   P G a N   P I n N =   Q p o l   e = 0.181   C m 2 . The critical evaluation of the polarization difference was given in [46]. The obtained polarization difference was calculated using the piezoelectric constants from [13,14]. In the calculations, the elastic constants obtained from Mahata et al. [47] have been used. Invoking the mechanical stability rule ε 33 = 2 C 13 ε 11 / C 33 , the polarization difference may be obtained as a function of a lattice parameter. The data collected in [46] indicate that the polarization difference obtained in [13] changes from Δ P G a N I n N = 0.09 C / m 2   for a = 3.113   nm to Δ P G a N I n N = 0.10 C / m 2   for a = 3.195   nm [13,46]. Ref. [14] provides similar dependence on the lattice parameters but the values are systematically lower Δ P G a N I n N = 0.07 C / m 2   for a = 3.113   nm to Δ P G a N I n N = 0.09 C / m 2   for a = 3.113   nm [14,46]. These values are slightly smaller than those obtained in the present simulations. Nevertheless, basic agreement between the earlier results and present simulations was obtained, proving the validity of the present approach.
Finally, the InxGa1−xN/GaN/In0.5Ga0.5N and InxGa1−xN/In0.5Ga0.5N/GaN structures described in Figure 1 have band profiles calculated from ab initio data by a projection of band states on the atom row states. This exact procedure was used previously for both AlN/GaN MQWs [16] and surface slab [48] simulations. The obtained results, presented in Figure 5, indicate that the Fermi level is located in the bandgap so the contribution of free carriers is small. In fact, the low InN bandgap leads to a small energy difference between the valence band maximum (VBM) and the Fermi energy in the fraction of the simulated system. It is expected that due to the relatively small InN bandgap ( E g D F T ( I n N ) = 0.75   e V ) compared to a wide GaN energy gap ( E g D F T ( G a N ) = 3.51   eV ), the smallest VBM-Fermi energy distance should be observed in Ga0.5In0.5N layer. It is not due to the combination of the energy change and the induced electric fields in case (a) that causes the highest VBM energy and is located in the graded InxGa1−xN layer interior about 1 nm from the In0.5Ga0.5N/InxGa1−xN interface. The estimated energy difference is about 0.1 eV; therefore, the screening by the mobile charge may contribute to the potential profile in the graded region. In case (b), the highest VBM energy is located at the In0.5Ga0.5N/GaN interface, due to the contribution of the polarization induced charge and dipole density. Again, we expect that this contribution may affect the obtained electric field jump at this interface. Therefore, the deviation in polarization difference obtained here and reported in the previous publications may be attributed to these factors [13,14,46].
From the point of view of possible applications, the emergence of the mobile carriers in addition to the immobile charge of the polarization-doping system is crucial, as discussed in [23]. It was found that in the case of AlxGa1−xN graded systems, the widening of the uniform part of GaN or In0.5Ga0.5N leads to penetration of the Fermi level in the conduction and the valence bands, so that the mobile carriers can screen the field. In AlxGa1−xN graded systems, the effect was more symmetric, the Fermi level was located at the midgap. In the case of the InxGa1−xN graded system, the Fermi level was located in the VBM vicinity. This opens the possibility of relatively easy emergence of holes which are particularly interesting in long wavelength devices, such as green and red laser diodes (LDs) and light-emitting diodes (LEDs). The solution of the p-type doping of In-containing parts of the device may be facilitated by the use of InxGa1−xN graded layers [22]. The design of such devices is possible through the use of drift-diffusion-based software [49,50,51,52,53,54]. This software was used in the simulation of nitride devices based on AlxGa1−xN graded systems [48]. The effect has immense technological importance as scattering by ionized defects is absent [55,56]. Thus, alloy scattering remains the dominant scattering process limiting carrier mobility [19,55,57]. Recently, polarization doping was successfully used in the design of AlGaN UVB LEDs proving that the concept could be used in nitride optoelectronic devices [30,58]. The present work opens its application in InGaN systems for the design of long wavelength LEDs and LDs.

4. Conclusions

Ab initio simulations of the InxGa1−xN/Ino.5Ga0.5N/GaN system proved the existence of the bulk polarization-doping immobile charge in the graded InxGa1−xN layer. It was shown that the bulk charge can be obtained using the following relation ρ p d = Δ P ( d x d z ) where the InN-GaN polarization difference is Δ P = Δ P G a N I n N = P G a N P I n N . The resulting polarization difference is Δ P G a N I n N = P G a N P I n N = 0.13   0.02   C / m 2 . The polarization difference is dependent on the lattice parameter via strong piezoelectric effects. Both these results, i.e., the magnitude and the piezo-dependence, are in accordance with the earlier critical evaluation of the polarization in [46] and the earlier obtained results in [13,14,46].
The obtained results are affected by several factors, characteristic of the investigated systems. InN is a narrow band material which entails a relatively small difference of the VBM and Fermi energy. Therefore, the mobile band charge (holes) may affect the electric potential distribution used to determine the value of the bulk charge density.
The obtained results indicate the existence of the polarization induced charge density in nitride heterostructures, in accordance with the polarization difference. In addition, the potential jumps of the order of 1 V at the interfaces prove the existence of the surface dipole layer at nitride heterointerfaces.
These results could be potentially applied in the p-type polarization doping of the In-rich structures used in green and red LEDs and LDs technology. Therefore, application of such a design may be highly beneficial in the extension of the emitted light range to longer wavelengths.

Author Contributions

Conceptualization, P.S. and S.K.; Methodology, P.S., Y.K., I.G., Z.R.Z., G.M., E.M. and A.K.; Software, A.A. and K.S.; Validation, P.K., K.S. and M.L.; Formal analysis, J.P. and S.K. All authors have read and agreed to the published version of the manuscript.

Funding

The research was partially supported by the National Science Centre, Poland - grant number 2021/43/B/ST7/03162 and by Poland National Centre for Research and Development (grant number: TECHMATSTRATEG-III/0003/2019-00) and partially by Japan JST CREST (grant number JPMJCR16N2) and by JSPS KAKENHI (grant number JP16H06418). This research was carried out with the support of the Interdisciplinary Centre for Mathematical and Computational Modelling at the University of Warsaw (ICM UW) under grants no GB77-29, GB84-23 and G88-1169. We gratefully acknowledge Poland’s high-performance computing infrastructure PLGrid (HPC Centers: ACK Cyfronet AGH) for providing computer facilities and support within computational grant no. PLG/2022/015976.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Landau, L.D.; Lifshitz, E.M.; Sykes, J.B.; Bell, J.S.; Dill, E.H. Electrodynamics of Continuous Media; Pergamon Press: Oxford, UK, 1984. [Google Scholar]
  2. Feneberg, M.; Thonke, K. Polarization fields of III-nitrides grown in different crystal orientations. J. Phys. Condens. Matter 2007, 19, 403201. [Google Scholar] [CrossRef] [PubMed]
  3. Zúñiga-Pérez, J.; Consonni, V.; Lymperakis, L.; Kong, X.; Trampert, A.; Fernández-Garrido, S.; Brandt, O.; Renevier, H.; Keller, S.; Hestroffer, K.; et al. Polarity in GaN and ZnO: Theory, measurement, growth, and devices. Appl. Phys. Rev. 2016, 3, 041303. [Google Scholar] [CrossRef] [Green Version]
  4. Ashcroft, N.W.; Mermin, N.D. Solid State Physics; Thomson Learning: Toronto, ON, Canada, 1976. [Google Scholar]
  5. Monch, W. Semiconductor Surfaces and Interfaces; Springer: Berlin/Heidelberg, Germany, 1993. [Google Scholar]
  6. Eller, B.S.; Yang, J.; Nemanich, R.J. Electronic surface and dielectric interface states on GaN and AlGaN. J. Vac. Sci. Technol. A Vac. Surf. Film. 2013, 31, 050807. [Google Scholar] [CrossRef]
  7. Miller, D.A.B.; Chemla, D.S.; Damen, T.C.; Gossard, A.C.; Wiegmann, W.; Wood, T.H.; Burrus, C.A. Band-edge electroabsorption in quantum well structures—The quantum confined Stark effect. Phys. Rev. Lett. 1984, 53, 2173–2176. [Google Scholar] [CrossRef]
  8. Halsall, M.P.; Nicholls, J.E.; Davies, J.J.; Cockayne, B.; Wright, P.J. CdS/CdSe wurtzite intrinsic Stark superlattices. J. Appl. Phys. 1992, 71, 907–915. [Google Scholar] [CrossRef]
  9. Kaminska, A.; Strak, P.; Borysiuk, J.; Sobczak, K.; Domagala, J.Z.; Beeler, M.; Grzanka, E.; Sakowski, K.; Krukowski, S.; Monroy, E. Correlation of optical and structural properties of GaN/AlN multi-quantum wells—Ab initio and experimental study. J. Appl. Phys. 2016, 119, 015703. [Google Scholar] [CrossRef]
  10. Waltereit, P.; Brandt, O.; Trampert, A.; Grahn, H.T.; Menniger, J.P.; Ramsteiner, M.; Reiche, N.; Ploog, K.H. Nitride semiconductors free of electrostatic fields for efficient white light-emitting diodes. Nature 2000, 406, 865–868. [Google Scholar] [CrossRef]
  11. Ambacher, O.; Foutz, B.; Smart, J.; Shealy, J.R.; Weimann, N.G.; Chu, K.; Murphy, M.; Sierakowski, A.J.; Schaff, W.J.; Eastman, L.F.; et al. Two dimensional electron gas induced by spontaneous and piezoelectric polarization in undoped and doped AlGaN/GaN heterostructures. J. Appl. Phys. 2000, 87, 334–344. [Google Scholar] [CrossRef]
  12. Franciosi, A.; Van de Walle, C.G. Heterojunction band offset engineering. Surf. Sci. Rep. 1996, 25, 1–140. [Google Scholar] [CrossRef]
  13. Bernardini, F.; Fiorentini, V.; Vanderbilt, D. Accurate Calculation of Polarization-Related Quantities in Semiconductors. Phys. Rev. B 2001, 63, 193201. [Google Scholar] [CrossRef] [Green Version]
  14. Dreyer, C.E.; Janotti, A.; Van de Walle, C.G.; Vanderbilt, D. Correct Implementation of Polarization Constant in Wurtzite Materials and Impact on III-Nitrides. Phys. Rev. X 2016, 6, 021038. [Google Scholar] [CrossRef] [Green Version]
  15. Strak, P.; Kempisty, P.; Sakowski, K.; Kaminska, A.; Jankowski, D.; Korona, K.P.; Sobczak, K.; Borysiuk, J.; Beeler, M.; Grzanka, E.; et al. Ab initio and experimental studies of polarization and polarization related fields in nitrides and nitride structures. AIP Adv. 2017, 7, 015027. [Google Scholar] [CrossRef] [Green Version]
  16. Strak, P.; Kempisty, P.; Ptasinska, M.; Krukowski, S. Principal physical properties of GaN/AlN multiquantum well (MQWs) systems determined by density functional theory (DFT) calculations. J. Appl. Phys. 2013, 113, 193706. [Google Scholar] [CrossRef]
  17. Noguera, C.; Goniakowski, J. Polarity in nano-objects. Chem. Rev. 2013, 113, 4073–4105. [Google Scholar] [CrossRef]
  18. Jena, D.; Heikman, S.; Green, D.; Buttari, D.; Coffie, R.; Xing, H.; Keller, S.; DenBaars, S.; Speck, J.S.; Mishra, U.K.; et al. Realization of wide electron slabs by polarization bulk doping in graded III–V nitride semiconductor alloys. Appl. Phys. Lett. 2002, 81, 4395–4397. [Google Scholar] [CrossRef] [Green Version]
  19. Jena, D.; Heikman, S.; Speck, J.S.; Gossard, A.; Mishra, U.K.; Link, A.; Ambacher, O. Magnetotransport properties of a polarization-doped three-dimensional electron slab in graded AlGaN. Phys. Rev. B 2003, 67, 153306. [Google Scholar] [CrossRef] [Green Version]
  20. Rajan, S.; DenBaars, S.P.; Mishra, U.K.; Xing, H.; Jena, D. Electron mobility in graded AlGaN alloys. Appl. Phys. Lett. 2006, 88, 042103. [Google Scholar] [CrossRef]
  21. Ji, X.L.; Lee, K.T.; Nazar, L.F. A highly ordered nanostructured carbon-sulphur cathode for lithium-sulphur batteries. Nat. Mater. 2009, 8, 500–506. [Google Scholar] [CrossRef]
  22. Simon, J.; Protasenko, V.; Lian, C.; Xing, H.; Jena, D. Polarization-Induced Hole Doping in Wide–Band-Gap Uniaxial Semiconductor Heterostructures. Science 2010, 327, 60–64. [Google Scholar] [CrossRef] [Green Version]
  23. Ahmad, A.; Strak, P.; Kempisty, P.; Sakowski, K.; Piechota, J.; Kangawa, Y.; Grzegory, I.; Leszczynski, M.; Zytkiewicz, Z.R.; Muziol, G.; et al. Polarization doping—Ab initio verification of the concept: Charge conservation and nonlocality. J. Appl. Phys. 2022, 132, 064301. [Google Scholar] [CrossRef]
  24. Akasaki, I.; Amanol, H.; Nakamura, S. Nobel Prize. 2014. Available online: https://www.nobelprize.org/prizes/physics/2014/summary/ (accessed on 31 January 2023).
  25. Nakamura, S.; Fasol, G.; Pearton, S.J. The Blue Laser Diode: The Complete Story; Springer: Berlin, Germany, 2000. [Google Scholar]
  26. Raring, J.W.; Hall, E.M.; Schmidt, M.C.; Poblenz, C.; Li, B.; Pfister, N.; Kebort, D.; Chang, Y.-C.; Feezell, D.F.; Craig, R.; et al. State-of-the-art continuous-wave InGaN laser diodes in the violet, blue, and green wavelength regimes. In Laser Technology for Defense and Security VI; SPIE: Bellingham, WA, USA, 2010; Volume 7686. [Google Scholar] [CrossRef]
  27. Denault, K.A.; Cantore, M.; Nakamura, S.; DenBaars, S.P.; Sehadri, R. Monolithic translucent BAMgAl10O17:Eu2+ phosphors for laser driven solid state lighting. AIP Adv. 2013, 3, 072107. [Google Scholar] [CrossRef]
  28. DenBaars, S.P.; Feezell, D.; Kelchner, K.; Pimputkar, S.; Pan, C.-C.; Yen, C.-C.; Tanaka, S.; Zhao, Y.J.; Pfaff, N.; Farrell, R.; et al. Development of gallium-nitride-based light-emitting diodes (LEDs) and laser diodes for energy-efficient lighting and displays. Acta Mater. 2013, 61, 945–951. [Google Scholar] [CrossRef]
  29. Skierbiszewski, C.; Wasilewski, Z.; Grzegory, I.; Porowski, S. Nitride-based laser diodes by plasma-assisted MBE—From violet to green emission. J. Cryst. Growth 2009, 311, 1632–1639. [Google Scholar] [CrossRef] [Green Version]
  30. Sato, K.; Yamada, K.; Sakowski, K.; Iwaya, M.; Takeuchi, T.; Kamiyama, S.; Kangawa, Y.; Kempisty, P.; Krukowski, S.; Piechota, J.; et al. Effects of Mg dopant in Al-composition-graded AlxGa1−xN (0.45 ≤ x) on vertical electrical conductivity of ultrawide bandgap AlGaN p–n junction. Appl. Phys. Express 2021, 14, 096503. [Google Scholar] [CrossRef]
  31. Ordejon, P.; Drabold, D.A.; Grumbach, M.P.; Martin, R.M. Unconstrained minimization approach for electronic computations that scales linearly with system size. Phys. Rev. B 1993, 48, 14646–14649. [Google Scholar] [CrossRef] [PubMed]
  32. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab initio order-N materials simulation. J. Phys. Condens. Matter 2002, 14, 2745–2779. [Google Scholar] [CrossRef] [Green Version]
  33. Troullier, N.; Martins, J.L. Efficient pseudopotentials for plane-wave calculations. Phys. Rev. B 1991, 43, 1993–2006. [Google Scholar] [CrossRef]
  34. Troullier, N.; Martins, J.L. Efficient pseudopotentials for plane-wave calculations. II. Operators for fast iterative diagonalization. Phys. Rev. B 1991, 43, 8861–8869. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Ruzsinszky, A.; Csonka, G.I.; Vydrov, O.A.; Scuseria, G.E.; Constantin, L.A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  37. Paszkowicz, W.; Adamczyk, J.; Krukowski, S.; Leszczynski, M.; Porowski, S.; Sokolowski, J.A.; Michalec, M.; Łasocha, W. Lattice parameters, density and thermal expansion of InN microcrystals grown by the reaction of nitrogen plasma with liquid indium. Philos. Mag. A 1999, 79, 1145–1154. [Google Scholar] [CrossRef]
  38. Leszczynski, M.; Teisseyre, H.; Suski, T.; Grzegory, I.; Bockowski, M.; Jun, J.; Porowski, S.; Pakula, K.; Baranowski, J.M.; Foxon, C.T.; et al. Lattice parameters of gallium nitride. Appl. Phys. Lett. 1996, 69, 73–75. [Google Scholar] [CrossRef]
  39. Ferreira, L.G.; Marques, M.; Teles, L.K. Approximation to density functional theory for the calculation of band gaps of semiconductors. Phys. Rev. B 2008, 78, 125116. [Google Scholar] [CrossRef] [Green Version]
  40. Riberio, M., Jr.; Fonseca, L.; Ferreira, L.G. Accurate prediction of the Si/SiO2 interface band offset using the self-consistent ab initio DFT/LDA-1/2 method. Phys. Rev. B 2009, 79, 241312. [Google Scholar] [CrossRef]
  41. Wu, J.; Walukiewicz, W.; Yu, K.M.; Ager, J.W.; Haller, E.E.; Lu, H.; Schaff, W.J.; Saito, Y.; Nanishi, Y. Unusual properties of the fundamental band gap of InN. Appl. Phys. Lett. 2002, 80, 3967–3969. [Google Scholar] [CrossRef] [Green Version]
  42. Matsuoka, T.; Okamoto, H.; Nakao, M.; Harima, H.; Kurimoto, E. Optical bandgap of wurtzite InN. Appl. Phys. Lett. 2002, 81, 1246–1248. [Google Scholar] [CrossRef]
  43. Wu, J.; Walukiewicz, W. Band gaps of InN and group III nitride alloys. Superlattices Microstruct. 2003, 34, 63–75. [Google Scholar] [CrossRef]
  44. Monemar, B.; Bergman, J.P.; Buyanova, I.A.; Amano, H.; Akasaki, I.; Detchprohm, T.; Hiramatsu, K.; Sawaki, N. The excitonic bandgap of GaN: Dependence on substrate. Solid-State Electron. 1997, 41, 239–241. [Google Scholar] [CrossRef]
  45. Yeo, Y.C.; Chong, T.C.; Li, M.F. Electronic band structures and effective-mass parameters of wurtzite GaN and InN. J. Appl. Phys. 1997, 83, 1429–1436. [Google Scholar] [CrossRef]
  46. Ahmad, A.; Strak, P.; Koronski, K.; Kempisty, P.; Sakowski, K.; Piechota, J.; Grzegory, I.; Wierzbicka, A.; Kryvyi, S.; Monroy, E.; et al. Critical evaluation of various spontaneous polarization models and induced electric fields in III-nitride multi-quantum wells. Materials 2021, 14, 4935. [Google Scholar] [CrossRef]
  47. Mahata, M.K.; Ghosh, S.; Jana, S.J.; Chakraborty, A.; Bag, A.; Mukhopadhyay, P.; Kumar, R.; Biswas, D. MBE grown AlGaN/GaN heterostructures on sapphire with ultra thin buffer. AIP Adv. 2014, 4, 117120. [Google Scholar] [CrossRef]
  48. Krukowski, S.; Kempisty, P.; Strak, P. Foundations of ab initio simulations of electric charges and fields at semiconductor surfaces within slab models. J. Appl. Phys. 2013, 114, 143705. [Google Scholar] [CrossRef] [Green Version]
  49. Gummel, H. A self-consistent scheme for one-dimensional steady state transistor calculations. IEEE Trans. Electron Devices 1964, 11, 455–465. [Google Scholar] [CrossRef]
  50. Roosbroeck, V.V. Theory of Flow of Electrons and Holes in germanium and Other Semiconductors. Bell Syst. Tech. J. 1950, 29, 560–607. [Google Scholar] [CrossRef]
  51. Sakowski, K. Determination of the Properties of the Nitride Laser Diodes and Light-Emitting Diodes by Simulation Based on the Drift-Diffusion Model with the Discontinuous Galerkin Method. Ph.D. Thesis, Faculty of Mathematics Informatics and Mechanics University of Warsaw, Warszawa, Poland, 2017. [Google Scholar]
  52. Di Pietro, D.A.; Ern, A. Mathematical Aspects of Discontinuous Galerkin Methods; Springer: Berlin, Germay, 2012. [Google Scholar]
  53. Sakowski, K.; Marcinkowski, L.; Strak, P.; Kempisty, P.; Krukowski, S. Discretization of the drift-diffusion equations with the composite discontinuous Galerkin method Lecture Notes. Comp. Sci. 2016, 9574, 391–400. [Google Scholar]
  54. Sakowski, K.; Marcinkowski, L.; Strak, P.; Kempisty, P.; Krukowski, S. On the Composite Discontinuous Galerkin Method for Simulations of Electric Properties of Semiconductor Devices. Electron. Trans. Numer. Anal. 2019, 51, 75–98. [Google Scholar] [CrossRef] [Green Version]
  55. Armstrong, A.M.; Allerman, A.A. Polarization-induced electrical conductivity in ultra-wide band gap AlGaN alloys. Appl. Phys. Lett. 2016, 109, 222101. [Google Scholar] [CrossRef]
  56. Simon, J.; Wang, A.; Xing, H.; Rajan, S.; Jena, D. Carrier transport and confinement in polarization-induced three-dimensional electron slabs: Importance of alloy scattering in AlGaN. Appl. Phys. Lett. 2006, 88, 042109. [Google Scholar] [CrossRef]
  57. Pant, N.; Deng, Z.; Khioupakis, E. High electron mobility of AlxGa1-xN evaluated by unfolding the DFT band structure. Appl. Phys. Lett. 2020, 117, 242105. [Google Scholar] [CrossRef]
  58. Sato, K.; Yasue, S.; Yamada, K.; Tanaka, S.; Omori, T.; Ishizuka, S.; Teramura, S.; Ogino, Y.; Iwayama, S.; Miyake, H.; et al. Room-temperature operation of AlGaN ultraviolet-b laser diode at 298 nm on lattice-relaxed Al0.6Ga0.4N/AlN/sapphire. Appl. Phys. Express 2020, 13, 031004. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) InxGa1−xN/GaN/In0.5Ga0.5N and (b) InxGa1−xN/In0.5Ga0.5N/GaN supercells that were subject to ab initio calculations. The GaxIn1−xN layer is graded linearly, with (a) increasing and (b) decreasing In concentration towards the bottom of the ([0001] direction). Ga, In and N atoms are represented by yellow, green and blue balls, respectively.
Figure 1. (a) InxGa1−xN/GaN/In0.5Ga0.5N and (b) InxGa1−xN/In0.5Ga0.5N/GaN supercells that were subject to ab initio calculations. The GaxIn1−xN layer is graded linearly, with (a) increasing and (b) decreasing In concentration towards the bottom of the ([0001] direction). Ga, In and N atoms are represented by yellow, green and blue balls, respectively.
Materials 16 01227 g001
Figure 2. Indium concentration (x) along the [000-1] axis in the structures presented in Figure 1. The zero coordinate is set at the bottom of the cell. The red circles and the blue squares represent indium concentration in the lattice plotted in Figure 1a,b, respectively. The line is for guiding the eye only.
Figure 2. Indium concentration (x) along the [000-1] axis in the structures presented in Figure 1. The zero coordinate is set at the bottom of the cell. The red circles and the blue squares represent indium concentration in the lattice plotted in Figure 1a,b, respectively. The line is for guiding the eye only.
Materials 16 01227 g002
Figure 3. The averaged potential profile obtained from ab initio calculations of the structures in Figure 1a. The zero coordinate value is set at the bottom of the cell. Solid blue lines represent averaged potential profiles, red dashed lines—parabolic approximation in the graded regions; green dashed lines—linear approximation in the uniform regions.
Figure 3. The averaged potential profile obtained from ab initio calculations of the structures in Figure 1a. The zero coordinate value is set at the bottom of the cell. Solid blue lines represent averaged potential profiles, red dashed lines—parabolic approximation in the graded regions; green dashed lines—linear approximation in the uniform regions.
Materials 16 01227 g003
Figure 4. The averaged electric potential profile obtained from ab initio calculations of the structures in Figure 1b. The zero distance is set at the bottom of the cell. Solid blue lines represent averaged potential profiles, red dashed lines—parabolic approximation in the graded regions; green dashed lines—linear approximation in the uniform regions.
Figure 4. The averaged electric potential profile obtained from ab initio calculations of the structures in Figure 1b. The zero distance is set at the bottom of the cell. Solid blue lines represent averaged potential profiles, red dashed lines—parabolic approximation in the graded regions; green dashed lines—linear approximation in the uniform regions.
Materials 16 01227 g004
Figure 5. Band profiles of the lattices presented in (a,b) These diagram show: left panel—energy bands in the momentum space, right panel—energy bands in the coordination space. The color at the top gives the scale of the density of states. The black line that is superimposed on the valence band represents the averaged potential multiplied by the electron charge.
Figure 5. Band profiles of the lattices presented in (a,b) These diagram show: left panel—energy bands in the momentum space, right panel—energy bands in the coordination space. The color at the top gives the scale of the density of states. The black line that is superimposed on the valence band represents the averaged potential multiplied by the electron charge.
Materials 16 01227 g005aMaterials 16 01227 g005b
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

Ahmad, A.; Strak, P.; Kempisty, P.; Sakowski, K.; Piechota, J.; Kangawa, Y.; Grzegory, I.; Leszczynski, M.; Zytkiewicz, Z.R.; Muziol, G.; et al. Polarization Doping in a GaN-InN System—Ab Initio Simulation. Materials 2023, 16, 1227. https://doi.org/10.3390/ma16031227

AMA Style

Ahmad A, Strak P, Kempisty P, Sakowski K, Piechota J, Kangawa Y, Grzegory I, Leszczynski M, Zytkiewicz ZR, Muziol G, et al. Polarization Doping in a GaN-InN System—Ab Initio Simulation. Materials. 2023; 16(3):1227. https://doi.org/10.3390/ma16031227

Chicago/Turabian Style

Ahmad, Ashfaq, Pawel Strak, Pawel Kempisty, Konrad Sakowski, Jacek Piechota, Yoshihiro Kangawa, Izabella Grzegory, Michal Leszczynski, Zbigniew R. Zytkiewicz, Grzegorz Muziol, and et al. 2023. "Polarization Doping in a GaN-InN System—Ab Initio Simulation" Materials 16, no. 3: 1227. https://doi.org/10.3390/ma16031227

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