You are on page 1of 9

1188

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 34, NO. 7, JULY 1998

Self-Assembled Semiconductor Structures: Electronic and Optoelectronic Properties


Hongtao Jiang and Jasprit Singh
Abstract Strained epitaxy has been shown to produce pyramidal-shaped semiconductor dot structures by single-step epitaxy. The very high density of these dots (approaching teradots per wafer) and their ever improving uniformity suggest that these features could have important applications in future microelectronics. Understanding the structural and electronic properties of these quantum dots is therefore of great importance. In this paper, we examine some of the physics controlling the performance of devices that could be made from such structures. The self-assembled quantum dots are highly strained and we will examine the strain tensor in these quantum dots using a valence force-eld model. In this paper we will address the following issues. 1) What is the general nature of the strain tensor in self assembled quantum dots? 2) What are the electron and hole spectra for InAsGaAs quantum dots? 3) What are the important intersubband radiative and nonradiative scattering processes in the self assembled quantum dots? In particular, we will discuss how the electronphonon interactions are modied in the quantum dot structures. Consequences for uncooled intersubband devices such as lasers, detectors, and quantum transistors will be briey discussed. Index Terms Bandstructure, gain, quantum dots, scattering, strain.

I. INTRODUCTION VER THE PAST several decades, shrinking semiconductor feature size has been the most important driving force in microelectronics performance improvements. However, as we reach feature sizes approaching a few hundred angstroms, it is becoming increasingly difcult to nd the required technology tools (resists, masks, lithography, etc.). The need for semiconductor structures with such small dimensions arises not only from the fact that very dense circuits can be fabricated but also because of new and potentially useful physics that come into play in such small structures. Over the past decade, a number of theoretical studies have shown that certain classes of semiconductor devices (both electronic and optoelectronic) will see significant improvements in performance if one could use quasizero-dimensional systems or quantum dots. However, the largest dimensions of the dots have to approach 150 A to make them useful in room-temperature applications. Such small structures are extremely difcult to fabricate with conventional processing techniques. A promising approach that has gradually emerged over the last decade is based on the
Manuscript received October 16, 1997; revised March 13, 1998. This work was supported by the National Science Foundation under Grant ECS9628973 and by the U.S. Army Research Ofce under Grant DAAG55-97-1-0156. The authors are with the Department of Electrical Engineering and Computer Science, The University of Michigan, Ann Arbor, MI 48105 USA. Publisher Item Identier S 0018-9197(98)04556-4.

use of high strain epitaxy to create self-assembly of quantum dots. A number of groups have shown that high-strain epitaxy can be exploited to produce quantum dot structures by single-step epitaxy. In this approach, a few monolayers of a semiconductor having a large lattice mismatch ( 3%) with a substrate is deposited. The deposited lm assembles into three-dimensional (3-D) islands instead of a smooth atomically at lm. The size and separation of the selfassembled dots can be controlled by lattice mismatch, surface preparation, and growth conditions. It is known experimentally that these self-assembled dots have a pyramidal shape and have a highly complex strain tensor. The lateral dimensions of these structures can approach 100 A while the vertical dimensions approach 70 A. The packing densities of these structures can approach 10 cm . These structures can be fabricated from important semiconductor systems such as SiGeSi, InGaAsGaAs, InGaAsInP, etc. It is thus possible to envisage a number of electronic and optoelectronic technologies that could exploit such self-assembled systems. However, to fully exploit these structures, several important issues have to be understood. These include: 1) the basic growth issues controlling self-assembly of semiconductor lms into arrays of dots; 2) the strain tensor in quantum dot structures fabricated by self-assembly; 3) the electronic spectra of the electrons and holes in quantum dots; 4) radiative and nonradiative scattering processes in the dots; and 5) optical and transport properties in these dots. It is important to understand that in addition to the new and interesting physics related to these dots, there are enormous computational challenges offered by these structures. These arise primarily due to the 3-D nature of the semiconductor structures and the breakdown of simple effective mass pictures that are quite useful in describing unstrained semiconductor devices. In this paper, we will address some of the issues outlined above. Usually in most semiconductor device formalisms it is possible to make the effective mass approximation which allows one to decouple one band (e.g., the conduction band edge) from other bands. This allows great simplication in understanding device physics. However, in self-assembled structures, there is a very large built-in strain in the dots. The strong strain tensor causes a very large change in the effective bandgap and transitions in the quantum dots. For example, in InAs quantum dots on GaAs substrates [1][4], the ground state transition energy is 1.05 eV even though the bandgap of InAs is 0.4 eV. This large strain-driven effect suggests that the electronic spectra of the structure should not be calculated by a simple effective mass approach. However, to simplify the problem, calculations based on the effective mass picture have

00189197/98$10.00 1998 IEEE

JIANG AND SINGH: SELF-ASSEMBLED SEMICONDUCTOR STRUCTURES

1189

been reported in the literature [5][7]. It is important to verify how accurate such a picture is. Unlike higher dimensional systems, quantum dot structures have discrete density of electronic states and in this regard they behave like atoms rather than solids. This gives these structures many potential advantages over other structures. Some of the potential devices that could benet from zero-dimensional physics of quantum dots include the following. Single electron transistors: since each dot can only occupy a few (2 or 6) electrons, change in occupancy by a single electron can produce measurable changes in device characteristics. This could be exploited in applications such as high-density memories. The discrete nature of electron levels can be exploited for a very low threshold laser with threshold current which is expected to be rather temperature-insensitive. The discrete nature of electron levels in the conduction band can be exploited for intersubband detectors and lasers. The intersubband separation can be tuned for applications in the 1020- m regimea regime that is very difcult to address by semiconductor devices. If dots are designed with large intersubband separation (say larger than 50 meV), one can effectively decouple the electronphonon system. This can result in temperature-independent device performance. A number of devices that require cooling to low temperatures could operate at room temperature with undiminished performance. In order to be able to examine and exploit the possibilities mentioned above, one has to develop an understanding of the physics of the semiconductor quantum dots. II. FORMALISM In this paper, we will address the following issues. 1) What is the general nature of the strain tensor in self-assembled quantum dots? 2) What are the electron and hole spectra for InAsGaAs quantum dots and what are the differences between the theoretical results based on an effective mass approximation in which approximation and an eight-band the inuence of remote bands is included? 3) what is the nature of the electronphonon interaction along with other carrier thermalization mechanisms in self-assembled quantum dot structures? 4) What are some implications of our ndings in devices based on self-assembled semiconductor structures? We will pay particular attention to the conduction band states since these could play an important role in several intersubband devices. A. Strain Model and Numerical Simulations In this paper, we will focus on the InAsGaAs system although most of the approaches and conclusions will apply to other self-assembled structures. The lattice mismatch between InAs and GaAs is 7% making the critical thickness 1.7 monolayers. It is found experimentally [1] that once the thickness of InAs exceeds a few monolayers, InAs islands are formed on top of a wetting layer. For growth along the [001] direction, these islands are small pyramidal-shaped with a square base in the (001) plane. The limiting planes of the

TABLE I BOND-STRETCHING AND BOND-BENDING PARAMETERS IN VFF MODEL [10]

islands vary for different experiments [1], [6]. It appears that both the strain value and growth conditions inuence the limiting planes. For a small base angle, it is reasonable to assume the strain is biaxial over all the dot area and this was done by [6]. But if the base angle is not small, the strain distribution is expected to be quite different from the biaxial strain and this is veried by calculations done in [5] based on the use of the elastic continuum theory. We use the valence force-eld (VFF) model by Keating [9] and Martin [10] to calculate the strain distribution. This model has been shown to be successful in tting and predicting the elastic constants of elastic continuum theory, calculating strain distribution in a quantum well (QW) [11], and determining the atomic structure of IIIV alloy [12], [13], and it was also used by [7] to calculate the strain distribution in QDs. The VFF model is a microscopic theory which includes the bond stretching and bond bending and avoids the potential failure of elastic continuum theory in the atomically thin limit. The total VFF energy is taken as

(1) runs over the nearest where run over all the atomic sites, is the vector joining the sites neighbor(NN) sites of and , is the length of the bond, is the corresponding and equilibrium length in the binary constituents, and are the bond stretching and bond bending constants, respectively. and for InAs and GaAs and Martin [10] calculated obtained values that give satisfactory agreement with experiments. We use these values (see Table I) in this paper. For the bond-bending parameter of In-As-Ga, we take following [14]. To nd the strain tensor in the InAsGaAs self-assembled dots, we start from an arbitrary choice of the atomic positions and minimize the system energy using the Hamiltonian given above. Minimization of the total energy requires one to solve variables, where is the a coupled set of equations with total number of atoms. Direct solution of these equations is impractical in our case since the system contains more than 100 000 atoms. We use the approach taken by several authors [11], [13], [15] which has been shown to be quite efcient. In the beginning of the simulation, all the atoms are placed on the GaAs lattice. We allow atoms to deviate from this starting position and use periodic boundary conditions in the plane which is perpendicular to the growth direction. In each process, only one atom position is displaced and other atom positions are held xed. The displacement of that atom is determined by . All the atoms solving three linearized equations are displaced in sequence. The whole sequence is repeated

1190

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 34, NO. 7, JULY 1998

TABLE II DEFORMATION POTENTIAL [8]

valence band bound states. Another important point to note is symmetry is broken microscopically. There are only that two reection symmetries (110) and (1 0). This will cause a splitting in the energy levels corresponding to states with different excitation directions.
(a)

B. Electronic Spectrum The problem of bandstructure in semiconductor quantum structures has usually been solved by decoupling the conduction and valence band problems. The conduction band problem is typically solved using a scalar effective mass approach while the valence band problem is solved using a approach which includes the heavy hole and light hole coupling. This decoupling is possible because in problems typically encountered (e.g., quantum wells and wires or even quantum dots with small strain), there is minimal remote band inuence on the conduction band. However, in the selfassembled quantum dot this is not the case. We note the following: 1) the bandgap of bulk InAs is 0.4 eV while the effective gap of the dot is 1.0 eV; 2) the nature of the strain tensor is such that there is a strong spatial variation in strain; 3) the strain components are very large and the resultant splittings in the bands are comparable to the interband separations in the bulk material. All these considerations suggest that the simple decoupled conduction-valence band picture and the effective mass description may not be adequate. description where the inuence We use the eight-band of remote bands on the conduction and valence band states is included. In the presence of strain the Hamiltonian has the form (2) is the usual kinetic term and has the form of (3) where [17], [18], shown at the bottom of the next page, where

(b)

(c) Fig. 1. (a) A schematic plot of the self-assembled quantum dot geometry. Lines A and B are in the [001] direction. (b) Hydrostatic and biaxial components of the strain through line A. (c) Shear strain through line B.

until the maximum distance moved in the sequence is so small that there is essentially no change in the system energy. Once the positions of all the atoms is known, the strain distribution is derived straightforwardly in the manner described by Keating [9]. A typical strain distribution is shown in Fig. 1 for an InAsGaAs quantum dot with (101) limiting planes. The results are very much the same as [7]. The strain distribution at the top of the dot shows there is relaxation for which was proved qualitatively by experiment [7]. The biaxial strain changes sign from the bottom to the top of the dot which means there is more connement for heavy holes at the bottom and more connement for light holes at the top of the dot. It is important to note the shear strain is also large. From the strain distribution, we expect that there is a strong mixing between the light holes and the heavy holes for the

The strain part can be obtained by the following substitution because of the same underlying symmetry: etc. Here we neglect the spinorbit-dependent deformation potential [19]. In Tables II and III, we list the parameters used in this work. The four-band KohnLuttinger parameters can be converted to eight-band parameters according to [17], [19], and [20].

JIANG AND SINGH: SELF-ASSEMBLED SEMICONDUCTOR STRUCTURES

1191

TABLE III EIGHT-BAND k 1 p PARAMETERS [24]

The piezoelectric elds are also included as [5]. The inuence is around 5 meV on the conduction band and less than 1 meV on the valence band. Both the conduction band states and valence band states can be derived from the eight-band Hamiltonian. We use the nite-difference method to solve this equation numerically. The distance between two grid points is chosen to be 5.6533 A (GaAs lattice constant) in the direction and two times this plane. Harmonic averaging [21] of the distance in the parameters is used in the interface region. It should be noted that the simple central difference approach for representing rst-order derivatives will cause spurious states though this method has higher accuracy. We combine central difference and second-order upwind [22] to alleviate the problem of spurious states and use the implicitly restarted Arnoldi/Lanczos methods [23] to solve the eigenvalue problem. Typical matrix sizes that are needed for the eight-band approach are 10 10 . The matrices are very sparse and only a few eigenvalues and eigenstates are needed. However, the matrix size places signicant hurdles in understanding the physics of these selfassembled dots. C. Stark Effect An important class of optoelectronic devices utilize the Stark effect in QW structures. Due to the shift in the electron and hole envelope functions, an electric eld can modulate the subband energy levels in QWs. Both interband and intersubband optical modulators have been conceived and fabricated using this effect in QWs.

The self-assembled structures also offer the possibility of Stark-effect-based devices. In our formalism it is straightforward to include an electric-eld-related potential energy term. It is important to note that the electric eld can be applied along the growth direction or in the growth plane. Due to the asymmetric shape of the dot, we expect the Stark effect to be different for these cases. Also, due to the pyramidal shape of the dot, the Stark effect will be different for a eld along the growth direction and one in the opposite direction. D. Scattering Formalism In this section we will focus on the scattering mechanisms responsible for thermalization of carriers injected into the excited energy levels of the self-assembled dots. The dominant scattering mechanisms in quantum dot structures have been shown to be the following. 1) ElectronPhonon Scattering Due to a Single Phonon Process: This mechanism can cause extremely rapid scattering unless the subband level spacing is much larger than the optical phonon energies. In this case there could be a phonon bottleneck since there may be severe suppression in rst-order electronphonon interactions. For quantum dots with a lower strain, the intersubband separation may overlap with optical phonon energies in which case this would provide an efcient coupling mechanism with the help of another LA phonon [25]. For the quantum dot being considered in this paper, this scattering mechanism is unimportant. 2) ElectronMultiphonon Scattering: In cases where a single-phonon process cannot couple the electronic states, multiphonon processes can play an important role. The scattering rate due to two LO phonon emission can be written as

(4)

(3)

1192

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 34, NO. 7, JULY 1998

where denotes the energy of intermediate level is the energy difference between level and , and is the EinsteinBose distribution which is essentially zero at low is the matrix element for optical phonons temperature. and is given through Fr hlich interaction o (5) where is the permittivity of vacuum and 12.25, 15.15. In calculating the above result, we neglect the continuum state contribution and assuming the LO phonon is dispersionless and substitute the Dirac function with a Gaussian distribution with FWHM 0.1 meV. 3) ElectronHole Scattering: Due to the nature of the valence band states, the levels in the valence band are separated by energies smaller than the optical phonon energy. This allows holes to thermalize rapidly [26]. The hot electrons in the excited conduction band levels can scatter from these cool holes and transfer their energy to the holes. The holes then lose their energy via phonons. It is very important to include selfconsistent level broadening in the evaluation of these rates for discrete electronic states. 4) ElectronElectron Scattering: It is possible for electrons in the discrete excited subband levels to relax by giving their energy to another electron. The electron that gets the energy can be a hot electron in the continuum states or in another discrete excited subband levels. If it is in the continuum states, it will jump to a higher continuum state. If it is in a discrete excited subband level, it will jump to a continuum state. E. Radiative Processes The intersubband radiative rate is given by the expression (6) is the where is refractive index, is the speed of light, is the momentum matrix element. free-electron mass, and F. Intersubband Optical Spectra An important potential application of the self-assembled quantum dots is in the area of intersubband detectors and lasers. Within the framework of the eight-band model, the wavefunction can be written as (7)

G. Gain Spectra The gain of the material is given by

(9) is the momentum matrix element between elecwhere is the broadening factor tronic state and hole state , is the volume due to size and occupation uctuation, and of the active region. III. RESULTS In order to establish our formalism, we use quantum dot shapes with a base width of 124 A, a height of 62 A, and the one monolayer of wetting layer. This size is close to those that have been examined in [2], [4], and [7] and have been used to understand the experimental results reported in [2] and [4]. It is important to note that while the formalism presented here is quite general specic results on scattering rates will change if the dot size or shape changes. These results are quite sensitive to the intersubband separations. Our results should therefore be restricted as applicable to the pure InAsGaAs dots. The dependence of scattering rates on dots composition and size will be presented later. In [2], the conduction band states have been addressed using . The an effective mass description with a mass of valence band states are also calculated using a decoupled heavy holelight hole picture. The resulting levels are shown in Fig. 2(a). This description gives a single bound state in the conduction band well and the authors have used this picture to understand results given in [2]. The results shown in Fig. 2(a) use an electron effective mass which corresponds to that of bulk InAs. Recognizing that the effective mass should correspond to a larger gap material, results given in [7] use instead of . The resulting the effective mass spectra is shown in Fig. 2(b). The use of a larger effective mass gives two bound levels in the conduction band. However, as noted in [7], a more sophisticated Hamiltonian is needed to shed more light on the excited states. In Fig. 3, we show the results obtained from the eight-band model. We see that there are a number of excited states in the conduction band. The presence of the excited states is clearly seen from these experiments [3], [27]. From our calculations we nd that the rst and second transitions observed in [2] are due to the ground electronic states to ground hole states and excited electronic states to excited hole states. Unlike the case shown in Fig. 2(a), the second transition is not due to the ground electronic states to excited hole states. An interesting feature of the calculated spectra is the near four-fold degeneracy of the second excited states. These states have an exact double degeneracy due to time-reection symmetry and are almost degenerate due to different excitation directions. The second excited electronic state is 62 meV higher than the rst excited states. There is considerable mixture of wetting

is the envelope function part and is the central where cell part. and This transition matrix element between initial state has the form nal state

(8)

JIANG AND SINGH: SELF-ASSEMBLED SEMICONDUCTOR STRUCTURES

1193

(a)

Fig. 3. Electronic spectrum for a InAsGaAs dot with base width 124 A and height 62 A calculated using the eight-band k 1 p model described in the text.

(b) Fig. 2. (a) Electronic spectrum for an InAsGaAs dot with base diameter 80 A and height 42 A calculated by the simple effective mass approach [2]. Solid lines represents states that can be observed in photoluminescence spectra and dashed lines represents states that cannot be observed in photoluminescence spectra. (b) Electronic spectrum calculated for the dot with base width 124 A and height 62 A using the effective mass approach for conduction band (m3 = 0:04m0 ) and a four-band k 1 p model for valence band [7].

layer states in this state because it is less conned. The direction of excitation is in the direction. As found by [7], we also nd many conned hole states. The splitting between the ground and excited states are from 22 to 30 meV. There are no wavefunction overlap between second excited hole states and second excited electron states due to different excitation directions. In Fig. 4 we show our results for the Stark effect. The solid line is for the case where the eld is along the [100] direction

and the dotted line is for the case where the eld is along the [001] (growth) direction. In Fig. 4(a) we show how the transition between ground states shifts with eld. We see that when the eld is in the growth plane the Stark effect is quite strong. There is a red shift of about 22 meV for a 100-kV/cm eld. When the eld is along the [100] axis the Stark effect is symmetric with respect to the orientation of the eld due to the symmetry of the dot. For the case of the eld parallel to the growth direction, the Stark effect is relatively weak. Also, the effect can cause either blue shift or red shift depending on the eld polarity. When the eld is in the growth direction, there is a blue shift when the eld is not too large. This occurs because the electric eld pushes the heavy holes to the top of the dot where they have less connement [Fig. 1(b)]. When the eld is greater than 150 kV/cm, the effective bandgap begins to decrease. When the electric eld is in a direction opposite to the growth direction, there is the usual red shift. In Fig. 4(b) we show the effect of the electric eld on the intersubband separation in the conduction band of the InAsGaAs dot. We see that for the case when the eld is along the [100] direction there is an decrease in the intersubband separation. The decrease is not very large but can be used for modulating intersubband absorption. When the eld is along the growth direction, we again see an asymmetric (with regard to the eld orientation) shift. In Fig. 5 we show the scattering rates involving two LO phonons for the ground and rst excited states. We see that when the energy separation between the rst excited and is equal to the the scattering time ground states is 0.1 ps. However, the rate drops off very rapidly when

1194

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 34, NO. 7, JULY 1998

(a)

Fig. 5. Calculated electron relaxation rate =2LO for 2LO processes. E21 h!LO is energy of the 2LO resonance.

0 2

(b) Fig. 4. Stark effect for quantum dots. The points are results calculated by the eight-band model. (a) Ground state transitions. (b) Intersubband separations as the function of electric eld. The results are for the conduction band subbands.

the resonance condition is not satised. For the InAsGaAs dot considered in Fig. 3, the two phonon process becomes unimportant. We have also considered the process involving two LO phonons and one LA phonon. These processes can couple the conduction band states. However, this scattering rate is found to give scattering times larger than 1 ns. Other multiphonon processes are even more suppressed. Thus for the InAsGaAs self-assembled quantum dot we nd that phonon-induced scattering is rather ineffectual. In Fig. 6 we show the eh scattering assuming the hole density corresponds to one hole per dot. We see that the scattering rate is very much dependent on the hole lifetime . Hole lifetime allows us to relax the energy conservation restriction in the scattering rate via a broadening function with

Fig. 6. Calculated electronhole relaxation times e-h dependence on hole lifetime h .

width

given by (10)

Since the initial and nal states are all discrete states, relaxation of the energy conservation relation is crucial for the scattering process. We expect that the hole lifetime (essentially due to phonon scattering) is 1 ps. Thus we see from Fig. 6 that eh scattering causes an intersubband relaxation time for electrons of 100 ps.

JIANG AND SINGH: SELF-ASSEMBLED SEMICONDUCTOR STRUCTURES

1195

(a)

Fig. 7. Intersubband absorption spectra for the InAsGaAs dot with base width 124 A and height 62 A.

Our calculations also show that for the InAsGaAs quantum dot being considered, the electronelectron scattering is considerably less efcient than the eh scattering discussed above. This is because the large separation between the rst excited state and the ground state and the nal state is in high continuum states. As a result, ee scattering is not so important in this structure. However, in shallower dots we expect ee scattering to be quite important. Based on the scattering rates discussed above, we have calculated the relaxation times for electrons injected in InAsGaAs quantum dots. We nd that in structures where electrons and holes are injected into the dots the thermalization times for electrons into the ground state are about 100 ps. This value is quite close to experimental studies based on the time-resolved PL spectrum [2]. In structures where only electrons are injected, we nd that the relaxation times are much longer approaching a nanosecond. We nd the intersubband radiative rate is 4.75 10 s for the InAsGaAs dot under consideration. The optical absorption between the ground and excited levels is found to have a value cm (11)

(b) Fig. 8. Material gain at different injection level at 300 K. The second excited states are assumed completely degenerate. The number given in the legend represents the electron number per dot. (a) Material gain for InAsGaAs 3 dots. The volume of the active gain associated with dots is 1.6 105 A . (b) Material gain for In0:4 Ga0:6 AsGaAs dots. The volume of the active 3 gain associated with dots is 6 105 A .

where is the linewidth of the transition in megaelectronvolts. It must be noted that while this is a very large value, the absorbing region is quite small (a few hundred Angstroms in length). As a result, the uniformity in dot sizes is quite critical. For the case where electrons and holes are injected into the quantum dot, the radiative relaxation time is long compared to the nonradiative relaxation time. Of course, if the device is being used in the lasing mode, the radiative lifetime changes as (12) where is the photon occupation number in the laser.

In Fig. 7 we show the intersubband absorption calculated for the InAsGaAs quantum dot of Fig. 3. An important point to note is that the very strong transition is due to the central cell part of the state and not due to the envelope part. This is unlike the wide QW case where in the conduction band the intersubband transition is due to the envelope function part. The very strong value of the transition matrix element makes the self-assembled dots very important for intersubband detectors. In Fig. 8, we show results for the gain spectra of the selfassembled quantum dot structures. Since detailed experimental results have been reported on In Ga AsGaAs quantum dot lasers [29], we have included the results of this system as well. In both cases, 3-meV FWHM is used. The peak differential

1196

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 34, NO. 7, JULY 1998

gain is around 5 10 cm which is close to the experiments [28], [29]. An interesting aspect of the results in Fig. 8 is revealed when we examine gain versus injection. At medium injection level, gain can only be reached in ground state transitions. However, at high injection level, due to the degeneracy effect, the gain is higher for the excited state transitions. Thus, when these dots are used in a laser, lasing will occur at the ground state if the cavity loss is smaller than maximum gain at ground states. However, for single-layer quantum dot lasers, this condition may not be fullled and lasing will occur at the excited states. This has been observed by the experiment [29]. IV. CONCLUSIONS We have calculated the strain distribution in a typical self-assembled InAsGaAs dot using the VFF method. The electronic spectra has also been studied using an eight-band model. Our results show that the effective mass approach for the conduction band gives signicant errors and also suffers from the problem that there is no nonarbitrary choice for the effective mass. The eight-band model gives a series of bound states in the quantum dot conduction band in contrast to the effective mass approach which only gives one or two bound states. The experimental spectra shows the presence of several electronic states in the conduction banda result that is explained by the eight-band model. Our results show semiquantitative agreement with experiments based on the capacitance measurements as well as on the photoluminescence measurements. We have also evaluated radiative and nonradiative scattering processes for carriers injected into the self-assembled dots. These processes are of great importance in intersubband and interband detectors, modulators, and lasers. REFERENCES
[1] M. Grundmann, J. Christen, N. N. Ledentsov, J. B hrer, D. Bimberg, o S. S. Ruvimov, P. Werner, U. Richter, U. G sele, J. Heydenreich, V. o M. Ustinov, A. Yu. Egorov, A. E. Zhukov, P. S. Kopev, and Zh. I. Alferov, Ultranarrow luminescence lines from single quantum dots, Phys. Rev. Lett., vol. 74, pp. 40434046, 1995. [2] F. Adler, M. Geiger, A. Bauknecht F. Scholz, H. Schweizer, M. H. Pilkuhn, B. Ohnesorge and A. Forchel, Optical transitions and carrier relaxation in self assembled InAs/GaAs quantum dots, J. Appl. Phys., vol. 80, pp. 40194026, 1996. [3] K. H. Schmidt, G. Medeiros-Ribeiro, M. Oestreich, and P. M. Petroff, Carrier relaxation and electronic structure in InAs/GaAs self-assembled dots, Phys. Rev. B, vol. 54, pp. 1134611353, 1996. [4] M. Grundmann, N. N. Ledentsov, O. Stier, D. Bimberg, V. M. Ustinov, P. S. Kopev, and Zh. I. Alferov, Excited states in self-organized InAs/GaAs quantum dots: Theory and experiment, Applied Phys. Lett., vol. 68, pp. 979981, 1996. [5] M. Grundmann, O. Stier, and D. Bimberg, InAs/GaAs pyramidal quantum dots: Strain distribution, optical phonons, and electronic structure, Phys. Rev. B, vol. 52, pp. 1196911981, 1995. [6] J.-Y. Marzin and G. Bastard, Calculation of the energy levels in InAs/GaAs quantum dots, Solid State Commun., vol. 92, pp. 437442, 1994. [7] M. A. Cusack, P. R. Briddon, and M. Jaros, Electronic structure of InAs/GaAs self-assembled quantum dots, Phys. Rev. B, vol. 54, pp. R23002303, 1996.

[8] Ch. G. Van der Walle, Band lineups and deformation potentials in the model-solid theory, Phys. Rev. B, vol. 39, pp. 18711883, 1989. [9] P. N. Keating, Effect of invariance requirements on the elastic strain energy of crystals with application to the diamond structures, Phys. Rev., vol. 145, pp. 637645, 1966. [10] R. M. Martin, Elastic properties of ZnS structure semiconductors, Phys. Rev. B, vol. 1, pp. 40054011, 1969. [11] J. Bernard and A. Zunger, Is there an elastic anomaly for a (001) monolayer of InAs embedded in GaAs?, Appl. Phys. Lett., vol. 65, pp. 165167, 1994. [12] A. Silverman, A. Zunger, R. Kalish, and J. Adler, Atomic-scale structure of disordered Ga10x Inx P alloys, Phys. Rev. B, vol. 51, pp. 1079510816, 1995. [13] F. Glas, Single-stage calculation of the total energy of compositionally modulated III-V alloys, J. Appl. Phys., vol. 66, pp. 16671670, 1989. [14] M. Podgorny, M. T. Czyzyk, A. Balzarotti, P. Letardi, N. Motta, A, Kisiel, and M. Zimnal-Starnawska, Crystallographic structure of ternary semiconducting alloys, Solid State Commun., vol. 55, pp. 413417, 1985. [15] M. R. Weidmann and K. E. Newman, Simulations of elastic-network relaxation: The Si10x Gex random alloy, Phys. Rev. B, vol. 45, pp. 83888396, 1992. [16] T. Benabbas, P. Francois, Y. Androussi, and A. Lefebvre, Stress relaxation in highly strained InAs/GaAs structures as studied by nite element analysis and transmission electron microscopy, J. Appl. Phys., vol. 80, pp. 27632767, 1996. [17] G. L. Bir and G. E. Pikus, Symmetry and Strain-Induced Effects in Semiconductors. New York: Wiley, 1974. [18] G. Bastard, Wave Mechanics Applied to Semiconductor Heterostructures. New York: Halsted Press, 1988. [19] F. H. Pollak, Semiconductors and Semimetals, vol. 32, pp. 1753, 1990. [20] J. Singh, Physics of Semiconductors and Their Heterostructures. New York: McGraw-Hill, 1993. [21] T. L. Li and K. J. Kuhn, Effects of mass discontinuity on the numerical solutions to quantum wells using the effective mass equation, J. Comp. Phys., vol. 110, pp. 292300, 1994. [22] W. Shyy, A study of nite difference approximations to steady-state convection-dominated ow problems, J. Comp. Phys., vol. 57, pp. 415438, 1985. [23] R. Lehoucq, D. C. Sorensen, and C. Yang, ARPACK Users Guide: Solution of Large Scale Eigenvalue Problems by Implicitly Restarted Arnoldi Methods, submitted for publication. [24] O. Madelung, M. Schulz, and H. Weiss, Eds., Numerical Data and Functional Relationships in Science and Technology, Landolt-B rnstein. o New York: Springer, 1985, vols. 17a and 21a. [25] T. Inoshita, and H. Sakaki, Electron relaxation in a quantum dot: Signicance of multiphonon processes, Phys. Rev. B, vol. 46, pp. 72607263, 1992. [26] I. Vurgaftman, K. Yeom, J. Hinckley, and J. Singh, presented at the Hot Carrier Conf., Chicago, IL, Aug. 1995. [27] M. Fricke, A. Lorke, J. P. Kotthaus, G. Medeiros-Ribeiro, and P. M. Petroff, Shell structure and electron-lectron interaction in selfassembled InAs quantum dots, Europhys. Lett., vol 36, pp. 197202, 1996. [28] D. Bimberg, N. Kirstaedter, N. N. Ledenstov, Zh. I. Alferov, P. S. Kopev, and V. M. Ustinov, InGaAs-GaAs quantum dot lasers, IEEE J. Select. Topics Quantum Electron., vol. 3, pp. 196205, 1997. [29] P. Bhattacharya, K. Kamath, J. Singh, D. Klotzkin, J. Phillips, H. Jiang, N. Chervela, T. Norris, T. Sosnowski, J. Laskar, and R. Murty, In(Ga)As/GaAs self-organized quantum dot lasers: DC and smallsignal modulation properties, IEEE J. Quantum Electron., submitted for publication.

Hongtao Jing, photograph and biography not available at the time of publication.

Jasprit Singh (M88), photograph and biography not available at the time of publication.

You might also like