Elsevier

Optics Communications

Volume 283, Issue 19, 1 October 2010, Pages 3703-3706
Optics Communications

Laser radiation effects on optical absorptions and refractive index in a quantum dot

https://doi.org/10.1016/j.optcom.2010.05.040Get rights and content

Abstract

An investigation of the laser radiation effects of a hydrogenic impurity in a quantum dot has been performed by using the matrix diagonalization method. We find that the laser field amplitude has an important influence on the linear, third-order nonlinear, and total absorption coefficients as well as the refractive index changes.

Introduction

In the last years, the study of the electronic and optical properties of low-dimensional semiconductor systems, such as quantum wells, quantum wires, quantum dots (QDs) and superlattices has been of great interest due to their importance for potential applications in electronic and optoelectronic devices. Of these systems, QDs have attained considerable theoretical and experimental attention due to their potential application in microelectronics and future laser technology [1], [2], [3]. QDs confine carriers in all three spatial dimensions and the many-body effects of particle–particle interactions show a broad range of electric structures similar to those of real atoms. The main features to consider in relation to QDs are geometrical shape, size, and the confining potential.

Especially the last two decades have witnessed an increasing interest in the theoretical and experimental investigations about the behavior of shallow hydrogenic donor, and acceptor impurities in semiconductor quantum heterostructures [4], [5], [6], [7], [8], [9], [10]. The study of impurities is one of the main problems in semiconductor low-dimensional systems because the presence of impurities in nanostructures influences greatly the electronic mobility and their optical properties [11]. Hence, the study of the impurity states in QDs has been extensively reported in the last few years [12], [13], [14], [15], [16], [17].

The linear and nonlinear optical properties are of particular interest, since they provide detailed information on the microscopic interactions of the quasiparticles. In semiconductor QDs, the nonlinear optical properties associated with optical absorptions are known to be greatly enhanced as compared to nonlinearity in bulk semiconductors. Hence, the nonlinear optical properties of QDs have been investigated both experimentally and theoretically by many authors [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], [28]. In the nonlinear optical absorption of QDs, the analysis of impurity states is inevitable because the confinement of quasiparticles in such structure leads to the enhancement of the oscillator strength of electron-impurity excitations. Since the operation of optoelectronic devices relies on the interaction of carriers with electromagnetic radiation, it is intriguing to investigate the effects of an intense laser field on the impurity states. Some investigations have been published about the effect of laser fields on nano-structure [29], [30]. However, to the best of our knowledge, there are only a few studies on the effect of an intense laser field on the optical transitions in a QD. In present work, we will focus on studying the effect of an intense, high-frequency laser field on the linear and nonlinear optical absorption coefficients (ACs) as the refractive index (RI) changes of a hydrogenic impurity in a disc-like QD by using the matrix diagonalization method and the compact density-matrix approach.

We consider one electron with a hydrogenic impurity confined by a disc-like QD, subjected to a laser field of frequency Ω, whose vector potential is given by A(t) = A0êxcosΩt. Here ê is the unit vector along the X axis. Following Refs. [29], [31], in the high-frequency limit, the Hamiltonian of a hydrogenic impurity confined by a disc-like parabolic QD can be written asH=p22me+Vdr,α0+VCdr,α0,with α0=α0e^x, where α0=(8π/Ωc)2Ie/me is the amplitude of the electron oscillation in the laser field, me being the effective mass of an electron and I the laser intensity. rP is the position vector (the momentum vector) of the electron originating from the center of the dot. The dressed parabolic and Coulomb potentials are defined as followsVdr,α0=14meω02rα0+r+α0,VCdr,α0=e22ϵ1rα0+1r+α0,where ω0 measures the strength of the confinement, and is the dielectric constant. The Hamilton can be rewritten asH=p22me+12meω02r2+12meω02α02e22ϵ1rα0+1r+α0.

From Eq. (4) a hydrogenic impurity in a two-center Coulomb field becomes apparent. The separation between the centers is 2α0, being proportional to I.

The Hamiltonian has cylindrical symmetry which implies the total orbital momentum L (i.e., Lz) is a conserved quantity, i.e., a good quantum number. Hence, the eigenstates of a hydrogenic impurity in a disc-like QD can be classified according to the total orbital angular momentum L. To obtain the eigenfunction and eigenenergy associated with the hydrogenic impurity in a disc-like QD, the Hamiltonian is diagonalized in the model space spanned by two-dimensional harmonic statesΨL=iciϕiωr,where is ϕiw(r)=Rnii(r)exp(-iliθ) ith two-dimensional harmonic oscillator eigenstate with a frequency ω and an energy (2ni + |i| + 1) ℏω. Rnℓ(r) is the radial wave function, given byRnℓ(r)=Nexpr2/2a2r||Ln||r2/a2,in which N is the normalization constant, a=/(meω), and Lnk(x) is the associated Laguerre polynomial. The radial and orbital angular momentum quantum members can have the following valuesn=1,2,,=0,±1,±2,.

This single-particle basis is used by the matrix diagonalization method in order to expand the Hamiltonian H. Here ω is the adjustable parameter, and in general not equal to ω0.

Let N = 2n + l. Let {ΨK} denote the set of basis functions including all the ΨK having their N smaller or equal to an upper limit Nmax. It is obvious that the total number of basis functions of the set is determined by Nmax. After the diagonalization we obtain the eigenvalues and eigenvectors. Evidently, the eigenvalues depend on the adjustable parameter ω. In our calculation, ω serves as a variational parameter to minimize the low-lying state energy. The matrix diagonalization method consists in spanning the Hamiltonian for a given basis and extracting the lowest eigenvalues of the matrix generated. The better the basis describes the Hamiltonian, the faster the convergence will be. The most common basis chosen is the one that describes the Hamiltonian at zero order.

The optical absorption calculation is based on the compact density-matrix approach, for which the total AC is given by [32]α(υ,I)=α(1)(υ)+α(3)(υ,I),whereα(1)(υ)=ωμϵRσsMfi2ΓfihυΔEfi2+Γfi2,andα(3)(υ,I)=υμϵRI2ϵ0nrcσsMfi2ΓfihυΔEfi2+Γfi224Mfi2MffMii23ΔEfi24ΔEfihυ+h2υ22Γfi2ΔEfi2+(Γfi)2,are the linear and third-order nonlinear optical ACs, respectively. In the above equations, ΔEfi = Ef  Ei denotes difference of the energy between lower and upper levels, c is the speed of light, I is the intensity of electromagnetic field, σs is the electron density, μ is the permeability of the system, nr is the RI and ϵR is the real part of the permittivity. Mfi = Ψi|x|Ψf〉 is the electric dipole moment of the transition from the Ψi state to the Ψf state. Here Γ is the phenomenological operator. Diagonal matrix element Γff of operator Γ, which is called as relaxation rate of fth state, is the inverse of the relaxation time Tf for the state |f〉, namely Γff = 1 / Tff. Whereas nondiagonal matrix element Γfi (f  i) is called as the relaxation rate of fth state and kth state. In a disc-like QD the dipole transitions are allowed only between states satisfying the selection rules Δl =  1, where l is the angular momentum quantum number. Hence, in the present work, we restrict our study to the transition of the L = 0 state to the L = 0 state.

The linear, and third-order nonlinear RI changes are given byΔn(1)(υ)nr=12nr2ϵ0σsMfi2EfihυEfihυ2+Γfi2,andΔn(3)(υ)nr=μcI4ϵ0nr3σsMfi2Efihυ2+Γfi224EfihυMfi2MffMii2Efi2+Γfi2EfihυEfiEfihυΓfi2Γfi22Efihυ.

The total RI change can be written asΔn(υ)nr=Δn(1)(υ)nr+Δn(3)(υ)nr.

In what follows the energy unit is meV and the length unit is nm. All calculations were performed using the following parameters: me = 0.067m0, where m0 is the free electron mass, ϵ = 12.4, σs = 5.0 × 1024 m 3, I = 1.5 × 1019W/m2, Γff = 1 ps 1, Γfi = 1/0.14 ps 1, and nr = 3.2. These parameters are suitable for GaAs–Ga1  xAlxAs QDs. Fig. 1, Fig. 2 show the linear, third-order nonlinear and total optical ACs and the total optical RI changes as a function of the incident photon energy for four different laser field values, i.e., α0 = 0.0, 1.0, 3.0 and 5.0 nm, respectively. The confining energy ℏω0 is set to be 15.0 meV. When α0 = 0, there is no laser field. The laser-field-induced effect is clear. It is readily seen that as the laser field value increases, the AC and the RI peaks will move to the left side, which show a laser-field-induced red shift of the resonance in QDs. This result is obviously different from that of quantum wells [29]. This redshift is because the energy difference between the ground state and the excited state in QDs decreases with increasing α0. Obviously this is the result of the laser-dressed Coulomb interaction. However, the peak values of the ACs and the RI changes are not monotonic functions of α0. As seen in Fig. 1, Fig. 2, we find that the magnitude of the ACs and the RI changes first increases with α0 up to a critical laser field value, α0  3.0 nm, and for further large α0 value they begin to decrease. On the one hand, we find that in the magnitude of ACs and RI considerably increases as the laser field inducing. Hence, we can say that in order to obtain the larger ACs and the RI changes in QDs, we should induce the laser field.

In Fig. 3, Fig. 4, in order to see clearly the quantum confinement effect on the optical properties, we set ℏω0 = 50.0 meV and plot the linear, third-order nonlinear and total optical ACs and the total optical RI changes as a function of the incident photon energy for four different laser field values, i.e., α0 = 0.0, 1.0, 3.0 and 5.0 nm, respectively. Fig. 3, Fig. 4 show that the qualitative properties of ACs and RI changes are in good agreement with those of Fig. 1, Fig. 2. However, the quantitative differences are also obvious. By comparing the AC and the RI peaks of Fig. 3, Fig. 4 with those of Fig. 1, Fig. 2 we can see that the AC peak increases and the RI peak decreases in magnitude for large confinement strength. On the other hand, we note that as the confinement strength ℏω0 increases, the ACs and the RI peak positions will move to the right side, which show a confinement-induced blue shift of the resonance in QDs.

In conclusion, we have calculated the ACs and the RI changes of a hydrogenic impurity in a disc-like QD in an intense laser field by using the matrix diagonalization method and the compact density-matrix approach. The results obtained have been presented as a function of the incident photon energy for the different values of laser field and confinement strength. We find that the linear, nonlinear third-order, and total optical ACs as well as the RI changes are strongly affected by the laser radiation and confinement strength of QDs. In order to obtain the larger ACs and the RI changes in QDs, we should induce the laser field.

Section snippets

Acknowledgment

This work is financially supported by the National Natural Science Foundation of China under grant No. 10775035.

References (32)

  • R. Greene et al.

    Solid State Commun.

    (1983)
  • C. Riva et al.

    Physica E

    (2004)
  • S. Yilmaz et al.

    Physica E

    (2007)
  • H. Sari et al.

    Physics Letters A

    (2003)
  • C. González-Santander et al.

    Physics Letters A

    (2010)
  • M.R.K. Vahdani et al.

    Physics Letters A

    (2009)
  • L. Jacak et al.

    Quantum Dots

    (1998)
  • A.D. Yoffe

    Adv. Phys.

    (2001)
  • Y. Masumoto et al.

    Semiconductor Quantum Dots

    (2002)
  • C. Aldrich et al.

    Phys. Status Solidi B

    (1979)
  • S. Chaudhury et al.

    Phys. Rev. B

    (1984)
  • N.C. Jarosik

    Phys. Rev. Lett.

    (1985)
  • Z. Xiao

    Phys. Status Solidi B

    (1995)
  • A. Latge'

    Phys. Rev. B

    (1996)
  • T. Szwacka et al.

    Condens. Matter

    (1996)
  • T. Ando et al.

    Mesoscopic Physics and Electronics

    (1998)
  • Cited by (20)

    • Optical properties of two dimensional fractal shaped nanostructures: Comparison of Sierpinski triangles and Sierpinski carpets

      2020, Optics Communications
      Citation Excerpt :

      Among the optical properties of the quantum dot nanostructures, the absorption coefficient has a special situation. So far effects of the laser radiation [16], donor impurity [17], an exciton [18], electric and magnetic fields [19], Rashba spin–orbit interaction [20], two electrons [21], three electrons [22], electron–phonon interaction [23], temperature and hydrostatic pressure [24], interdiffusion [25], etc. on the absorption coefficient of quantum dots have been investigated. Along with these effects, effects of different shape of the quantum dot and type of the confining potential such as parabolic quantum dots [26], Gaussian potentials [27], generalized Hulthén potential [28], modified Poschl–Teller​ potential [29], disk-like quantum dots [30], capping layer in spherical layer quantum dots [31], core–shell nanodots [32], etc. on the absorption coefficient of quantum dots have extensively been explored.

    • Impurity-modulated optical response of a disc-shaped quantum dot subjected to laser radiation

      2020, Photonics and Nanostructures - Fundamentals and Applications
    • Self-assembled strained pyramid-shaped InAs/GaAs quantum dots: The effects of wetting layer thickness on discrete and quasi-continuum levels

      2014, Physica E: Low-Dimensional Systems and Nanostructures
      Citation Excerpt :

      Although the role of WL is significant in electronic and optical properties of QDs [16–25], there has been noticeably fewer research on the effects of WL on QD properties [16]. In spite of discarding WL in theoretical works [26–34], self-assembled QDs cannot be grown without WL. This ignorance in theoretical works may arise from mathematical complexities in making analytical models.

    View all citing articles on Scopus
    View full text