Excitonic transitions in Be-doped GaAs/AlAs multiple quantum well
Zheng Wei-Min1, †, , Li Su-Mei2, Cong Wei-Yan1, Wang Ai-Fang1, Li Bin3, Huang Hai-Bei4
School of Space Science and Physics, Shandong University, Weihai 264209, China
School of Information Engineering, Shandong University, Weihai 264209, China
Shanghai Institute of Technical Physics, Chinese Academy of Sciences, Shanghai 200083, China
School of Chemistry, the University of Melbourne, Victoria 3010, Australia

 

† Corresponding author. E-mail: wmzheng@sdu.edu.cn

Project supported by the National Natural Science Foundation of China (Grant No. 61178039) and the Natural Science Foundation of Shandong Province, China (Grant No. ZR2012FM028).

Abstract
Abstract

A series of GaAs/AlAs multiple-quantum wells doped with Be is grown by molecular beam epitaxy. The photoluminescence spectra are measured at 4, 20, 40, 80, 120, and 200 K, respectively. The recombination transition emission of heavy-hole and light-hole free excitons is clearly observed and the transition energies are measured with different quantum well widths. In addition, a theoretical model of excitonic states in the quantum wells is used, in which the symmetry of the component of the exciton wave function representing the relative motion is allowed to vary between the two- and three-dimensional limits. Then, within the effective mass and envelope function approximation, the recombination transition energies of the heavy- and light-hole excitons in GaAs/AlAs multiple-quantum wells are calculated each as a function of quantum well width by the shooting method and variational principle with two variational parameters. The results show that the excitons are neither 2D nor 3D like, but are in between in character and that the theoretical calculation is in good agreement with the experimental results.

1. Introduction

Quasi-two-dimensional semiconductor-based quantum well and superlattice structures have extensively been investigated because of their potential applications in compact Terahertz emitters, cascade lasers, small-size detectors, etc.[15] A characteristic feature of the optical spectra of quantum well and superlattice systems at low temperatures is the occurrence of excitons. The exciton is often treated as a ‘two-body’ combination involving an electron and a hole between which there is Coulombic interaction. In a three-dimensional (3D) crystal of semiconductor bulk material the envelope function ϕr describing the relative motion of excitons between an electron and a hole in the ground state has the following form:

where r2 = (xexh)2 + (yeyh)2 + (zezh)2 and λ is the Bohr radius of the exciton. It is clear that the function ϕr has the same 3D symmetry as the Hamiltonian describing the relative motion term. The scenario also holds true for a two-dimensional (2D) exciton confined to a plane, with the function ϕr now having the form

where

In a quantum well or superlattice structure, the problem is complicated by the quantum-well confined potential existing along the z-direction. Therefore, the exciton in a quantum confined system is neither 2D nor 3D like. However, in the majority of theoretical calculations of the transition and binding energies of the excitons in the literature, the shape of wave function representing the electron-hole relative motion is usually chosen as either the 2D[68] or 3D exciton[9,10] given by Eqs. (2) and (1), respectively. Choosing function ϕr, the exciton problem can then be solved to various degrees of sophistication. In addition, the exciton in an anisotropic or confined system is also studied by the model of fractional-dimensional space,[1114] in which the anisotropic or confined system can be treated as an isotropic system in the effective fractional-dimensional space, and excitonic bound-state energies and wave functions are obtained by solving the simple hydrogenic Schrödinger equation in the fractional-dimensional space.

In this paper, the envelope function ϕr describing the electron-hole relative motion in the ground state is taken as the intermediate between the 2D and 3D limit, which has the following form:

where and both ζ and λ are used as variational parameters, which are systematically varied in the numerical calculation so as to minimize the energy of the exciton. This fractional-dimensional choice for ϕr is rarely found in the literature. Then, within the effective mass and envelope function approximation, the transition recombination energies of the heavy- and light-hole excitons in GaAs/AlAs multiple-quantum wells are calculated each as a function of the quantum well width by the shooting method and variational principle with two variational parameters. Experimentally, we measure photoluminescence (PL) spectra at various temperatures for a series of GaAs/AlAs multiple quantum wells with Be δ-doping at the well center together with a single epilayer of GaAs with uniform Be doping (GaAs:Be). The transition recombination energies of the heavy- and light-hole excitons are measured, respectively for the samples with different quantum well sizes. Finally, the calculating results are compared with the experimental results.

2. Experiment and results

A series of GaAs/AlAs multiple-quantum wells was grown on semi-insulating (100) GaAs substrates by molecular-beam epitaxy with Be acceptors δ-doped at the quantum-well center. The growth of the layers was performed under the exact stoichiometric condition using the technique of low-temperature growth, which ensures high-quality optical materials even at relatively low growth temperatures. Under these conditions, the quantum-well structures were grown at 550 ˚C or 540 ˚C without interruptions at the quantum well interfaces, which ensured negligible diffusion of the Be δ layers. Prior to the growth of the multiple-quantum wells, a GaAs buffer layer of 300 nm was grown. Each of the multiple-quantum well structures investigated contained an identical 5-nm-wide AlAs barrier, while each GaAs well layer was δ doped at the well center with Be acceptor atoms. The doping level, doping mode and main characteristics of each of the samples are summarized in Table 1. The sample 7 is a 5-μm-thick single GaAs:Be epilayer with uniform Be doping, which is used as a reference with the quantum well width being infinite and Be acceptors being under no quantum confinement. For the sample 5, the beryllium acceptors were uniformly doped in the 2-nm central region of a 20-nm GaAs quantum-well layer instead of the δ-doping at the well center. Furthermore, in comparison with other samples, sample 6 has a special structure, where an Al0.3Ga0.7As barrier was grown at the central 2-nm region of a 10-nm well layer and doped uniformly with Be acceptors, i.e., each of the 10-nm quantum wells is divided again into two 4-nm-wide quantum wells by the 2-nm-wide Al0.3Ga0.7As barrier.

Table 1.

Characteristics of the samples: the repeated period, quantum-well width (Lw), Be-doped concentration (P), doping mode and growth temperature (T) of the epitaxial layer.

.

Photoluminescence experiments were performed from liquid helium to room temperature using a Renishaw RM1000 Raman microscope. The samples were mounted on a cold finger of a continuous flow Helium cryostat. The optical excitation for PL experiments was provided by an argon-ion laser (514.5 nm). The laser beam was focused onto a sample, and the light reflected from the sample was collected and entered into a spectrometer for analysis. The excitation power was typically 5 mW.

The PL spectra with above-band-gap excitation have been measured at the various temperatures for the samples in Table 1. Figure 1 shows the PL spectra at 4, 20, 40, 80, 120, and 200 K for sample 5. Three peaks are clearly resolved at 4 K, with energy positions located at (1.5247±0.0002), (1.5287±0.0002), and (1.5337±0.0002) eV, respectively. The first labeled Be0X at (1.5247±0.0002) eV is the strongest peak and originates from the recombination of excitons bound to the neutral beryllium acceptor. The second strongest peak at (1.5287±0.0002) eV is attributed to the transition of a free heavy-hole exciton XCB1−HH1. The energy separation between the XCB1−HH1 and Be0X is 4 meV, which is the energy required to remove an exciton from the Be0X complex. The third strongest peak, XCB1−LH1, located at (1.5337±0.0002) eV arises from the recombination emission of free light-hole excitons. In addition, on the low energy side of the Be0X there is an even weaker peak in intensity than the XCB1−LH1, XCB1−HH1, and Be0X, which is attributed to the free-to-bound recombination eBe0 between an electron of the n = 1 quantized confined level and a hole bound to a Be acceptor at the center of the GaAs well. At 40 K below, the intensities of the XCB1−HH1 and XCB1−LH1 peak are both enhanced as the measuring temperature rises, while both positions slightly shift towards the low energy. When the temperature reaches up to 40 K, the intensity of the XCB1-HH1 exceeds that of Be0X. This is because the kinetic energy of free excitons increases with temperatures rising, which makes it hard that the free excitons are bound by acceptors to become bound excitons, and which gives rise to a reduction of the number of bound excitons. However, for temperatures above 80 K, as the measuring temperature further increases, neither Be0X nor eBe0 is detectable, with the energy positions of XCB1-LH1 and XCB1-HH1 observably shifting towards the low energy. This is attributed to the monotonic reduction of the band gap of GaAs bulk material with the temperatures rising. Furthermore, compared with sample 4 which has the same well width as sample 5 but is δ-doped with Be acceptors at the well center, sample 5 has slightly greater transition energies of the heavy- and light-hole excitons for sample 5, specifically, the XCB1-HH1 and XCB1-LH1 are (1.5285±0.0002) eV and (1.5335±0.0002) eV at 20 K for sample 5 respectively, while they are (1.5269±0.0002) eV and (1.5314±0.0002) eV, respectively, for sample 4.[15] Therefore, the differences in transition energies at 20 K between both samples are 1.6 meV and 2.1 meV for the XCB1-HH1 and XCB1-LH1, respectively. The doping sheet concentration of sample 4 with Be acceptors δ-doped at the well center is roughly two orders of magnitude higher than that of sample 5. Subsequently, the ions of δ-doped Be acceptors will form a spike potential in the quantum-well center, which forces the energy level of the ground state of electrons to be lowered with respect to the conduction band bottom, while the energy levels of the heavy- and light-holes shift towards the valence-band top. Consequently, the recombination emission energies of free excitons for sample 4 are less than those of sample 5.

Fig. 1. Series of PL spectra for sample 5 at 4, 20, 40, 80, 120, and 200 K, showing the XCB1−LH1, XCB1−HH1, Be0X, and eBe0 peaks.

Figure 2 illustrates the PL spectra of sample 6 at different temperatures. In comparison with Fig. 1, however, only two emitting peaks are observed at temperature 4 K, which are the XCB1−HH1 and Be0X located at (1.6659±0.0002) eV and (1.6577±0.0002) eV, respectively. The Be0X is a very weak shoulder peak, but neither XCB1−LH1 nor eBe0 can be resolved clearly. The quantum well size of sample 6 is 4 nm much narrower than that of sample 5, thus the electrons and holes in the GaAs quantum-well layer are confined more strongly by the quantum-well potential, resulting in the fact that the energy levels of the ground state and excited states of electrons holes rise relatively. Therefore, this makes it difficult that the heavy-hole excitons are captured by Be acceptors and that the light-hole excitons are formed by Coulomb interactions between an electron and a hole.[16]

Fig. 2. Series of PL spectra for sample 6 at temperatures of 4, 20, 40, 80, 120, and 200 K, showing the XCB1−HH1 and Be01X peaks.
3. Calculation and discussion
3.1. Theoretical model

Under the effective mass and envelope function approximations, for an exciton in the GaAs/AlAs quantum well, the Hamiltonian representing the interacting two-body electron-hole complex can be considered as the sum of three terms

where He and Hh are the one-particle Hamiltonians appropriate to the conduction and valence bands, respectively, of the GaAs/AlAs quantum well. The one-particle Hamiltonians are written as

The third term He−h on the right-hand side of Eq. (4) represents the electron-hole interaction, which is composed of two terms. One of these terms corresponds to the kinetic energy of the relative motion of the electron and hole in the xy plane (perpendicular to the growth axis), while the other represents the Columbic potential energy, i.e.,

where p is the quantum mechanical momentum operator for the in-plane component of the relative motion, and μ, the reduced mass of the electron–hole pair.

and r is simply the electron-hole separation, given by

The problem consists in finding the eigenfunctions Ψ and eigenvalues of the Schrödinger equation:

The two-body exciton wave function Ψ is taken as a product of three factors as follows:

where ψr is a variational wave function employed to minimize the total energy E of the system. The other two factors, ψe(ze) and ψh(zh), are simply the eigenfunctions of the one-particle Hamiltonians of the GaAs/AlAs quantum well:

One of the main advantages of this formalism is that it is independent of the forms of the one-particle Hamiltonians He and Hh, and indeed calculations can be performed on any system in which the standard electron and hole wave functions can be calculated.[15,17,18]

Multiplying Eq. (9) on the left by Ψ and integrating over all space, then the total exciton energy follows simply as the expectation value:

The wave function ψr representing the electron–hole relative motion in the ground state is chosen to be a function of a hydrogenic atom, given by

where the Bohr radius λ will be used as a parameter and systematically varied in order to minimize the total energy E of the system. A variable symmetry-type relative motion is chosen as

The second variational parameter, ζ, allows the exciton to be assumed to have any shape of wave function in a fractional-dimensional space. Traditionally, the case with ζ = 0 has become known as the two-dimensional exciton,[7,8] and ζ = 1, as a three-dimensional exciton.[9,10] The cases where ζ is allowed to take values other than 0 and 1 are rarely found in the literature.

3.2. Results and discussion

Using Eq. (11), the one-particle electron and hole eigenstates, ψe (ze) and ψh (zh), are calculated numerically by the shooting method, respectively, and then substituted into Eq. (12) together with ψr. The parameters of λ and ζ are systematically varied in order to minimize the total energy E of the system. In the numerical calculation, the effective mass values of the electron and hole, and are taken as the typical values for bulk GaAs, i.e., 0.067m0 and 0.62m0, respectively, where m0 is the rest mass of an electron in the free space. The relative dielectric permittivity εr is set to be 13.18.

Figure 3 displays the variational calculation results of recombination transition energies of heavy-hole XCB1−HH1 and light-hole XCB1−LH1 excitons, respectively, as a function of the quantum well width for the 5-period GaAs/5 nm AlAs multiple quantum wells, in which a finite structure resembles an infinite structure. It is clearly seen from Fig. 3 that the transition energies of both heavy-hole and light-hole excitons fall monotonically as the quantum well width increases, and are in good agreement with measured data of photoluminescence spectra. In addition, for the heavy-hole exciton XCB1−HH1 at 4.2 K, the transition energy tends smoothly towards the value of bulk GaAs:Be, i.e., (1.5142±0.0002) eV.[19] However, the reason is that the quantum confined effect on the electrons or holes in the wells becomes weak as the well thickness increases. Theoretically, when the well width goes to infinity, the quantum confined effect on the carriers in the wells tends toward zero, where the carriers are the same as those in the GaAs bulk.

Fig. 3. Plots of exciton transition energy versus quantum well width in a 5-period GaAs/5 nm AlAs multiple quantum wells for (a) heavy-hole excitons and (b) light-hole excitons.

Figure 4 shows the variations of parameter ζ in the variational calculation for the heavy and light-hole excitons, XCB1−HH1 and XCB1−LH1, with quantum well width for the 5-period GaAs/5 nm AlAs quantum wells, respectively. It can be seen from Fig. 4 that both of ζ rise as the quantum well thickness increases. Furthermore, the ζ value determines in physical meaning the shape dimensionality of spatial wave functions of excitons. Traditionally, when ζ = 0 and ζ = 1, the excitons correspond to two-dimensional (2D) and three-dimensional (3D) excitons, respectively. However, the variational calculation method used here differs from that employed by He in Refs. [11] and [13] with the model of fractional-dimensional space, in which an anisotropic system in the 3D space can be treated as an isotropic system in the effective fractional-dimensional space, and excitonic bound-state energies and wave functions are obtained by solving the simple hydrogenic Schrödinger equation in the fractional-dimensional space. Then, Kundrotas et al. extended the fractional-dimensional approach to analyze transitions of the free-electron-acceptor impurities located at the center of the quantum well, where effective fractional dimensionality, 1 < α < 3, is determined by experimental results, such as the PL spectra.[12]

Fig. 4. Plots of variational parameter ζ versus quantum well depth in the 5-period GaAs/5-nm AlAs quantum wells.
4. Conclusions

We investigate experimentally and theoretically the recombination transition energies of the heavy– and light-hole excitons, XCB1−HH1 and XCB1−LH1, for a series of GaAs/AlAs multiple quantum wells doped with Be in the wells or in barriers. The PL spectra are measured at the various temperatures, and the recombination emissions of the heavy- and light-hole excitons, XCB1−HH1 and XCB1−LH1, are clearly observed. The transition energies of the XCB1−HH1 and XCB1−LH1 is measured experimentally. In addition, within the effective mass and envelope function approximations the recombination transition energies of the heavy- and light-hole excitons are calculated numerically, respectively, as a function of the quantum-well size via the shooting method and variational method involving two variational parameters. The results show that the exciton in the ground state is neither 2D nor 3D like, but is in between in character, and that the theoretical numerical calculation is in good agreement with the result from the measured PL spectra.

Reference
1Chen X RSong Y XZhu L QQi ZZhu LZha F XGuo S LWang S MShao J 2015 Chin. Phys. Lett. 32 067301
2Ünsal Ö LGönül BTemiz M 2014 Chin. Phys. 23 077104
3Huang H BZheng W MCong W YMeng X YZhai J B 2013 Phys. Status Solidi 250 1352
4Halsall M PHarrison PWells J P RBradley I VPellemans H 2001 Phys. Rev. 63 155314
5Li S MZheng W MWu A LCong W YLiu JChu N NSong Y X 2010 Appl. Phys. Lett. 97 023507
6Sivalertporn KMouchliadis LIvanov A LPhilp RMuljarov E A 2012 Phys. Rev. 85 045207
7Ivchenko E LKavokin A VKochereshko V PPosina G RUraltsev I NYakovlev D RBicknell-Tassius R NWaag ALandwehr G 1992 Phys. Rev. 46 7713
8Monozon B SSchmelcher P 2010 Phys. Rev. 82 205313
9Kuo Y HLi Y S 2009 Phys. Rev. 79 245328
10Bellabchara ALefebvre PChristol PMathieu H 1994 Phys. Rev. 50 11840
11He X F 1991 Phys. Rev. 43 2063
12Kundrotas JČerškus AAšmontas SValušis GSherliker BHalsall M PSteer M JJohannessen EHarrison P 2005 Phys. Rev. 72 235322
13He X F 1990 Solid State Commun. 75 111
14Rønnow T FPedersen T GPartoens BBerthelsen K K 2011 Phys. Rev. 84 035316
15Zheng W MHalsall M PHarmer PHarrison PSteer M J 2002 J. Appl. Phys. 92 6039
16Feldmann JPeter GGöbel E ODawson PMoore KFoxon CElliott R J 1987 Phys. Rev. Lett. 59 2337
17Harrison PQuantum Wells, Wires and Dots: Theoretical and Computational Physics of Semiconductor NanostructuresEnglandWiley
18Zheng W MWang A FLu Y BZhang PHong D 2007 Semicond. Sci. Technol. 22 74
19Zheng W MHalsall M PHarmer PHarrison PSteer M J 2004 Appl. Phys. Lett. 84 735