Anda di halaman 1dari 10

196

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 3, NO. 2, APRIL 1997

InGaAsGaAs Quantum-Dot Lasers


D. Bimberg, N. Kirstaedter, N. N. Ledentsov, Zh. I. Alferov, P. S. Kopev, and V. M. Ustinov
Abstract Quantum-dot (QD) lasers provide superior lasing characteristics compared to quantum-well (QW) and QW wire lasers due to their delta like density of states. Record threshold current densities of 40 A1cm02 at 77 K and of 62 A1cm02 at 300 K are obtained while a characteristic temperature of 385 K is maintained up to 300 K. The internal quantum efciency approaches values of 80%. Currently, operating QD lasers show broad-gain spectra with full-width at half-maximum (FWHM) up to 50 meV, ultrahigh material gain of 105 cm01 , differential gain of 10013 cm2 and strong nonlinear gain effects with a gain compression coefcient of 10016 cm3 . The modulation bandwidth is limited by nonlinear gain effects but can be increased by careful choice of the energy difference between QD and barrier states. The linewidth enhancement factor is 0.5. The InGaAsGaAs QD emission can be tuned between 0.95 m and 1.37 m at 300 K. Index Terms Characteristic temperature, gain, linewidth enhancemant factor, quantum-dot gain, semiconductor lasers, threshold current density.

I. INTRODUCTION EMICONDUCTOR lasers having as active medium an uniform array of quantum dots (QDs) are expected to exhibit properties superior to three-dimensional (3-D), quantumwell (QW) and QW wire lasers [1][3]. Due to the three dimensional connement of carriers in a quantum dot with a size below or equal to the exciton Bohr radius the density of states becomes a delta function [4]. Consequently, no temperature dependent broadening of the gain function is possible if the sublevel splitting is sufciently large. QD lasers with such an atomic like density of states are expected to show ultra low-threshold current densities, ultrahigh temperature stability of threshold current, ultrahigh differential gain increased, cutoff frequency and chirpfree operation under direct current modulation. Potential device applications therefore range from high-power semiconductor lasers to high-speed light sources for ber-based data transmission. Practical realization of the fundamental advantages of QD lasers was proposed to be hampered both by a theoretically predicted othogonality of electron- and hole-wave function in QDs leading to a very small radiative recombination probality and by carrier capture delayed by more than nanoseconds due to the phonon bottleneck. Both effects together were predicted to lead to zero quantum efciency for dots of still macroscopic dimensions (100 nm) [5]. Recent numerical calculations [6] of electron, hole and exciton wavefunctions for actual InAsGaAs QD levels on the other hand did not
Manuscript received December 3, 1996. D. Bimberg and N. Kirstaedter are with Technical University-Berlin, 10623 Berlin, Germany. N. N. Ledentsov, Zh. I. Alferov, P. S. Kopev, and V. M. Ustinov are with the Ioffe Physical-Technical Institute, 194021, St. Petersburg, Russia. Publisher Item Identier S 1077-260X(97)04525-5.

conrm the complete wavefunction orthogonality of ground and excited states. In addition multiphonon relaxation allows efcient carrier capture [7] and experiments presented in detail below yield luminescence decay times of 2 ns for allowed transitions and carrier capture times of 20 ps. Consideration of the actual statistical nature of carrier capture into realistic dots having excited states and of the importance of the weak or vanishing thermal coupling between adjacent QDs lead recently to considerable revision of the original predictions for gain, threshold and chirp [8], [9]. In the last decade two different approaches have been explored to fabricate quantum dots. The rst one originally developed for quantum-wire lasers uses different variations of patterning techniques either prior to growth for selective growth or after growth (e.g., chemical etching with following overgrowth). Using patterning techniques InGaAsInGaAsP QD lasers were successfully realized, but they still show high threshold current density of 7.6 kA cm at 77 K [10]. The second approach simply uses strain induced self-organization of InGaAsGaAs quantum dots [11], [12] on low index substrates yielding threshold current densities as low as 60 A cm at room temperature [13]. Self-organization is also very successfully applied to the growth of InP on GaInP [14] and GaSb [15] or InSb [16] on (100)-GaAs resulting in the formation of 3-D-islands. Growth of InGaAs on (311)B-GaAs [17] shows formation of dots with a disk like shape. Their diameter of 60 nm is however comparatively large and no large impact of quantum effects could be yet demonstrated. Since the rst presentation of a QD injection laser [11] based on selforganized dots as gain medium large efforts were undertaken to decrease the threshold current density to its ultimate limit determined by the transparency condition of the dot states. For this purpose one has to increase the internal efciency and the carrier capture rate into the dots which present the most serious performance limitations for QD lasers. Since 1994 the threshold density has beeen indeed decreased by orders of magnitude as visualized in Table I. The threshold will soon approach the limit given by the actual dot size distribution (see Section III on threshold characteristics) and there will be now additional strong focus on other fundamental properties of QD lasers like the temperature and excitation dependence of the gain spectra and the modulation characteristics. II. REALIZATION
OF

QD LASERS

A schematic view of the band structure of a typical QD laser is presented in Fig. 1. The ideal QD laser consists of a 3-D-array of dots with equal size and shape surrounded by a higher bandgap material which connes the injected carriers. The whole structure is embedded in an optical waveguide consisting of lower and upper cladding layers.

1077260X/97$10.00 1997 IEEE

BIMBERG et al.: InGaAsGaAs QUANTUM-DOT LASERS

197

TABLE I HISTORY OF MAJOR STEPS IN THE DECREASE OF THRESHOLD CURRENT DENSITY OF QD LASERS. THE SECOND COLUMN LISTS THE DOT (QD) AND BARRIER MATERIAL, THE NUMBER OF QUANTUM DOT LAYERS (DOT LAYERS) GROWN ON TOP OF EACH OTHER, THE TYPE OF DOT FORMATION (FORMATION), THE SIZE OF THE DOT BASELENGTH (DOT SIZE), THE DOT GROWTH TEMPERATURE (Tgrowth ), THE AREA DOT DENSITY (DOT DENSITY) AND THE GROWTH REACTOR TYPE (REACTOR). THE THIRD COLUMN LISTS THE THRESHOLD AT VARIOUS TEMPERATURES AND THE FOURTH COLUMN THE CORRESPONDING LASING WAVELENGTH AND THE TYPE OF THE LASING TRANSITION. WL INDICATES LASING TRANSITION VIA WETTING LAYER STATES, QD3 VIA EXCITED QUANTUM DOT STATES AND QD VIA QD GROUND STATE

Fig. 1. Schematic bandstructure of a quantum dot laser with self-organized dots under forward bias. A 3-D-array of dots vertically aligned along the growth direction which is formed during the growth of multiple QD layers is illustrated schematically. Typically the dot area density in the (100)-plane is 1010 cm02 and the dot size distribution is around 10%. The distance 4 between the dot layers is 5 nm and the real dot density in the recombination volume with a thickness of 200 nm is 6 1015 cm03 for three QD layers.

Fig. 2. Modal gain spectrum of a single layer of quantum dots obtained after deposition of nominally 2ML of InAs derived from single pass amplication. The transition energies of the QD ground state and excited states and the WL are marked.

Spontaneous formation of QDs for currently operating QD lasers (see Table I) was successfully demonstrated at growth temperatures between 460 C and 550 C. The QDs are formed using a StranskiKrastanov growth mode. The areal dot densities range between 2 10 cm [19] and 1 10 cm [4] with a typical size distribution of 10% [22]. The low-dot density causes several problems concerning threshold and gain that will be described in the following section. If the cladding layer and the GaAs QD barrier are also grown at the same low temperature as the active dot layers they are a possible source for current leakage and nonradiative recombination. The QDs show intermixing with the surrounding barrier material if temperatures of 700 C are used to grow high-quality cladding layers. The limited dot areal density sets an upper value for the QD modal gain. The modal gain can be increased by stacking

several layers of QDs upon each other [23]. Using this approach ground state lasing [13] at room temperature has been demonstrated. The relaxation rate might actually exceed values of 10 electrons/s permitting a modulation bandwidth up to 200 GHz [24] if the energy difference between barrier and dot states is carefully tailored. III. GAIN
OF

SELF-ORGANIZED QD LASERS

Some of the most important properties which strongly inuence the threshold current density and modulation bandwidth of a semiconductor laser are the material gain and the differential material gain. The material gain of a QD laser can be derived from the complex susceptibility of an ensemble of atoms interacting with a time-dependent electromagnetic

198

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 3, NO. 2, APRIL 1997

eld. In case the dots show only one electron and one hole level the susceptibility is as follows:

(1) and are proportional to the absorption (material is the photon gain) and the refractive index, respectively. energy, the dot transition energy, the Bloch matrix element and the QD electron hole wavefunction overlap accounts for all integral. The phase coherence time mechanisms which cause a loss of coherence between the photon eld and the QD exciton. According to (1) the lineshape of the gain is Lorentzian and its width is determined by the phase coherence time. In the case of self-organized dots the transition energies of different dots are not identical due to size or shape uctuations. Thus the gain function is inhomogenously so that the broadened and one can dene a function probability of a dot having its emission energy between and is . is normalized such that the probability of nding a dot emission within an innite energy range becomes unity. If is multiplied with the ideal dot density of a 3-D-dot array in the recombination volume one gets the density of states for an inhomogenously broadened dot ensemble. The ideal dot density is the number of dots per volume where the number of dots is chosen as the total recombination volume divided by the dot volume .

(a)

(b) Fig. 3. (a) Material gain spectra as a function of excitation density ranging from 20250 A1cm02 for an ensemble of InAs dots having a base length of 7 nm and a size distribution of 13%. The WL is assumed to be a 0.5-nm-thick single QW. The inset shows the peak material gain (gmat ) at 1.28 eV and the differential material gain (dgmat =dN ) as a funcion of current density. (b) Arrhenius plot of WL- and QD electroluminescence (EL) well below threshold current density proving thermal equilibrium of carriers between QD and WL states.

(2) The factor of 2 in (2) accounts for the spin degeneracy. is the maximum of the dot distribution. The width of the Gaussian is given by . By combining (1) and (2), one gets the material gain for an inhomogenously broadened dot ensemble.

(3a) (3b) (3c) The Fermi functions determine the occupation probability of the single dot states by electrons and holes. If the minimum spectral distance between transitions from two different dots is much smaller than the single dot emission line width (determined by the phase coherence time) the gain spectrum of (3) is determined by the density of states of the dot ensemble. Since the total number of dots in the recombination volume of a typical ridge-waveguide laser exceeds a value of 10 the function gain spectrum of a single dot transition (demonstrated by Grundmann et al. [4]) transforms to a broad gain spectrum as shown in Fig. 2 for QDs formed after the deposition of two monolayer (ML) InAs. Due to the lateral separation of the dots carriers in different dots are thermally coupled only via wetting layer (WL) and GaAs barrier states [23]. The coupling between dots and the

WL is directly proved by the intensity ratio of QD and WL electroluminescence (EL) at different temperatures and constant injection level which follows an Arrhenius dependence [see Fig. 3(b)]. The temperature dependence of the intensity ratio can be used to derive an activation energy of the QD exciton to the WL of 25 meV. Based on the structural parameters of the laser, the luminescence spectra and the transition energies of ground and excited states one can derive the material gain spectra according to (3). Since the Bloch matrix element and the overlap integral are not exactly known for the case of a QD laser one has to calibrate the gain spectra [22] to absolute values by using the well known equilibrium value of the absorption coefcient of bulk GaAs. The derivation of the material gain versus current density in Fig. 3(a) is based on the experimentally determined decay in the recombination volume (see Fig. 1) time of carriers in order to calculate the carrier density in that volume from the injection current density [25].

(4)

BIMBERG et al.: InGaAsGaAs QUANTUM-DOT LASERS

199

Here, is the thickness of the recombination volume (see Fig. 1), is a factor which accounts for the different volumes occupied by QD, WL, and GaAs [22], is the density of states, is the Fermi function and the electron charge. The modal and material gain spectra in Figs. 2 and 3 are asymmetric due to the contribution of excited QD states already at comparatively low excitation in agreement with recent theoretical predictions [8]. The overlapping gain functions of ground and excited states enhance the material gain to a value of 9 10 cm at 250 A cm in the single layer QD laser investigated here. This value is one order of magnitude larger than for a 8 nm In Ga AsGaAs single QW laser [22] and in agreement with the prediction of Asada et al. [2]. The current density for the onset of gain saturation shown in Fig. 3(a) is much larger than the value estimated from the laser rate equation [26] in the case that all carriers are conned to the QD states. (5) Here, is the dot area density and is the exciton decay time of the InAs QDs. The factor of three accounts for one ground state and two excited states. From (5) the gain saturation level is 19 A cm for a dot area density of 4 10 cm and a QD decay time of 2 ns [22]. This saturation level is exceeded in real QD laser structures by one order of magnitude. In the next section it will be shown that current leakage due to the population of barrier states (WL, GaAs) and nonradiative recombination in our yet nonoptimized structures can explain the delayed saturation level. IV. THRESHOLD CHARACTERISTICS From the gain versus current density relation and the lasing condition one can easily calculate the threshold (transparency) current density by (6a) (6b) Here, is the fraction of the current density injected into QDs and contributing to radiative recombination, is the fraction of the current density from the barriers contributing to radiative recombination and the fraction of the current density contributing to the leakage current at threshold (transparency). is the sum of internal loss and mirror losses [see (10)]. Although the maximum material gain of the QD ensemble reaches a value of 10 cm according to (3) the connement factor dened as

Fig. 4. Threshold current density (solid line) as a function of the dot density. The calculation is done for a typical laser structure shown in Fig. 1 with an active layer thickness of 200 nm and total loss of 8 cm01 . The threshold current densities for single and triple QD layer lasers are marked. The transparency current density is also shown (dashed line). The dashed area marks the forbidden region due to gain saturation.

lateral coverage of 2% in the (100)-plane (dot areal density of 4 10 cm , dot base length 7 nm) and the QD exciton volume 80 nm [22] the connement factor for a single layer dot laser structure is estimated to be 1 10 . Thus for a maximum material gain of 10 cm a typical single-layer QD laser structure is expected to have a modal gain of 10 cm in agreement with our experimental results. This value is just sufcient for laser action in long cavity lasers. Short cavity lasers require a higher dot volume density. A way to overcome this limitation is the introduction of multiple layers of dots [23], [13]. The maximum material gain is calculated from the steady state rate equation with the radiative current density (assuming that all injected carriers are 100% conned to the dots and no leakage current exists) and the population inversion factor : (8a)
-

(8b) (8c)

(7) is rather low for a single layer QD laser. Here denotes the number of dots in the recombination volume (dened in Fig. 1), is the integral over the laser volume of the electric eld vector of the optical mode which is normalized to unity at its maximum value. Using a typical dot

Here, is the dot density and is the density of QD excitons calculated from the steady-state rate equation using the QD carrier decay time and the thickness of the recombination volume (see Fig. 1). Using (8) and the laser condition [see (6)] one can calculate the threshold current density as a function of the dot density (Fig. 4) assuming additionally that only one energy level exists for electrons and holes. For a total loss of 8 cm (e.g., for a long laser cavity of 2 mm and an internal loss of 3 cm ) the threshold current density for a single-layer QD laser structure with a gain width of 50 meV and a QD carrier decay time of 2 ns is as low as 2 A cm . Thus, even for a QD laser with typical dot size distribution of 10% and dot areal density of 4 10 cm the minimum threshold current density neglecting any current

200

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 3, NO. 2, APRIL 1997

Fig. 6. Threshold current density as a function of temperature for a single dot layer laser structure. The dots are placed in a 16-nmwide GaAsAlGaAs QW to reduce the escape probability of QD excitons into the barrier. This enhances T0 up to 385 K while maintaining a low threshold current density of 400 A1cm02 at 300 K.

Fig. 5. (a) Temperature dependence of threshold current density for a single layer QD laser. The large T0 between 50 K and 100 K reveals the temperature independent peak gain of the QD ensemble. (b) Above 100-K barrier population and nonradiative centers decrease the decay time of carriers in the QD ground state. The integrated electroluminescence (EL) intensity of QD, WL barrier and GaAs barrier at threshold unambiguously indicates the onset of current leakage above 100 K resulting in the decrease of T0 .

leakage can reach values more than one order of magnitude lower than in high quality QW lasers which exhibit lowest threshold current densities of 50 A cm at 300 K [27]. From (8), we see that for longer decay times (i.e., lower oscillator strength) which are expected for dots with size smaller or equal to the exciton Bohr the transparency current further decreases. But simultaneously the reduced oscillator strength reduces the maximum gain, thus, increasing the threshold current to maintain the threshold gain. Note that the transparency current is proportional to the dot volume density and also to the number of dot layers if the carriers in the QDs are not thermally coupled. A. Temperature Dependence of Threshold and Loss Mechanism A nonequilibrium distribution of carriers between all dots in a dot ensemble should prohibit thermal broadening of the gain spectrum. If this condition is strictly fulllled the peak gain should remain constant as a function of temperature at constant injection level. Therefore the phenomenological description of the threshold temperature dependence according to the following equation results in an innite characteristic temperature . (9) due to carrier The rst observation of an enhancement of connement was reported by Arakawa et al. [1] for bulk lasers placed in a magnetic eld. A dramatic increase of for a true single layer QD laser up to 425 K is shown in Fig. 5. Unfortunately in this structure the breakdown of the nonequilibrium carrier distribution and the temperature dependent current leakage reduces the value of at temperatures above 100 K.

It was discussed in the last section that for this laser structure the carriers in the dots are thermally coupled to the WL and the GaAs barrier. Since the activation energy of a QD exciton (conned to a 7-nm InAsGaAs dot) is only 25 meV (see Fig. 3) one can expect two loss mechanisms: First of all, with increasing temperature the injected carriers start to populate the barrier states thus increasing the injection current to maintain the threshold gain of the QD laser. Secondly nonradiative recombination in the barrier might lead to current leakage. This leakage current is related to the quality of the barrier layer (in particular the low temperature GaAs) and does not present a principal limitation of the threshold current. It has been shown [28] that higher growth temperatures of the cladding layers and optimization of the dot size (leading to an increase of the activation energy) shifts the onset of current leakage to 220 K with an ultrahigh of 530 K between 80 K and 220 K. Another way to decrease the leakage current is to place the dot layers in a GaAsAlGaAs QW, thus, drastically decreasing the escape probability of carriers from the dots (see Fig. 6). With this approach the onset of current leakage has been increased up to room temperature while maintaining a high of 385 K between 80 K and 300 K [29]. An other important factor characterizing the laser quality is of the laser. This value the differential internal efciency can be derived from the dependence of the external differential quantum efciency on mirror losses according to (10a) (10b) Here, is the laser facet reectivity, the cavity length, and the internal loss of the laser. The differential internal efciency of the single-layer QD laser is about 40%. This value was improved to 50% by introducing multilayer QD structures (see Fig. 7). Values of 70% for a In Ga AsGaAs QD laser and of 81% for a In Ga AsGaAs QD laser were recently reported by Mirin et al. [19] and Ustinov et al. [30]. These values are not much lower than the recently reported record internal efciency of QW lasers of 99% [31]. The more typical internal efciency around 50% of current QD lasers demonstrates that the difference between the cal-

BIMBERG et al.: InGaAsGaAs QUANTUM-DOT LASERS

201

(a)

(b) Fig. 7. (a) The threshold density versus mirror losses yields a transparency current density below 20 A1cm02 and demonstrates clearly gain saturation for small cavity lengths. (b) External differential efciency for single and multiple layer QD laser as a function of mirror losses. The internal efciency obtained for a stack of ten QD layers is about 50% at 300 K.

Fig. 8. (a) Blueshift of the QD emission due to state lling. Each spectrum is marked with the corresponding laser cavity length. For short cavity lasers the QD gain is completely saturated and the emission jumps to WL states. (b) Light current characteristic of a six-fold layer QD laser with a 400-m-long and 5-m-wide laser cavity showing single-mode operation just above the threshold.

culated threshold current density of 2 A cm (see Fig. 4) for an quasi-ideal QD laser (with no leakage current and one electron and one hole state) and the measured value of 100 A cm for a single layer QD laser having both a total loss of 8 cm is still connected to current leakage. On the other hand the transparency current density for a single layer QD laser shown in Fig. 7 is below 20 A cm indicating that indeed ultralow threshold current densities are feasible. V. SPECTRAL CHARACTERISTICS The emission of QD lasers consisting of a dot ensemble with only one energy state for electrons and holes is expected to occur at the maximum of the PL emission and should not show any energy shift with injection current. In the case of real structures emission from both ground and excited states is often observed. Thus the line shape of the gain spectrum is no longer independent of the injection level causing a strong blueshift of the laser emission with injection current. Such a blueshift is demonstrated in Fig. 8. Obviously excited QD states provide higher modal gain here. Upon complete QD gain saturation the laser emission jumps to the WL state which has a much higher modal gain due to the larger connement factor. One of the most interesting features of a QD laser is its potential to produce longitudinal single-mode operation when the gain spectrum is more narrow than the longitudinal

mode spacing dened by the laser cavity length. None of the currently existing QD lasers exhibit a gain width below 1 meV, which is necessary to fullll this condition for stripe geometry lasers with a typical laser cavity length of 0.5 mm. For vertical-cavity surface-emitting lasers (VCSEL) there is a much larger chance to obtain single-mode lasing. The full-width at half-maximum (FWHM) of the QD PL ranges from 33 meV [20] to 50 meV [22] which is sufciently narrow for VCSELs to show single-mode operation. Nevertheless the experiment on FabryPerot lasers show an unexpected single-mode operation just above the threshold (see Fig. 8). For QW lasers it is well known that they can switch to one single mode at injection levels far above threshold despite their broad-gain spectra. Two possible mechanism might be responsible for the single-mode operation reported here. Due to inhomogenities of the dot distribution the gain spectra may have local minima near the main gain maximum. This effect will effectively increase the threshold of other longitudinal modes near the main emission mode. Consequently the rst suppressed mode will start lasing only at higher injection levels. In addition dots of different sizes might have different carrier capture times. Thus under stimulated emission conditions only dots that are rapidly relled contribute. The wavelength of the laser emission from QD states can be As composition and by varying tuned by varying the In Ga the size of the dots. With this approach the QD PL emission could be tuned between 0.95 and 1.37 m at 300 K [29]. Thus

202

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 3, NO. 2, APRIL 1997

(a)

Fig. 9. Electroluminescence spectra (solid line) for an InAs QD laser with a vefold stacked dot layer at 1.05 times the threshold current density at 293 K. The inset shows the pulsed light current characteristic. The photoluminescence spectra (dotted line) at 293 K indicates that laser emission is due to the QD states. WL and GaAs transitions are also marked.

the rst optical transmission window of glass bers is already covered by emission from QDs grown on GaAs substrate. Using InSb instead of InAs 1.5- m emission is possible. It was shown by several groups that In Ga As QD lasers in the composition range of 0.3 1 can be grown by MBE [23], [13] but growth by MOCVD was thought to be more complicated because of the different growth kinetics as compared to MBE [32]. Recently, we discovered a MOCVD growth window for InAsGaAs QDs and demonstrated a three-layer InAsGaAs QD laser. The threshold at 77 K is as low as 21 A cm and lasing occurs via the QD ground state (see Fig. 9). The threshold current density increases to 320 kA cm at 293 K and the laser emission is red shifted by 45 meV going from 77 K to 293 K which is only slightly less than the InAs bandgap change [21]. If the ground state gain is enhanced by the existence of several QD layers, the laser operates far away from ground state gain saturation and all higher lying energy levels, especially the excited states are well below the transparency condition. Therefore lasing occurs in this case via the ground state and an increase of temperature does not increase the occupation of the excited states above the transparency level. Then the QD laser emission should follow the bandgap change of the dot material with temperature. VI. DYNAMIC PROPERTIES
AND

(b) Fig. 10. (a) Time resolved spectra of the QD ground state and excited state luminescence integrated over 5 meV spectral bandwidth for a single dot layer with dots grown by nominally 2ML InAs deposition. The inset shows the corresponding spectra of the QD structure where the energetic position of the QD ground (1) and excited (2) state as well as the peak energy of QD and WL luminescence are marked. (b) Time behavior of the integrated QD luminescence (scattered graph) of a threefold stacked dot layer with dots grown by nominally 2ML InAs deposition and a GaAs barrier width of 1.5 nm between adjacent dot layers. The excitiation puls has a full width half maximum of 38 ps and an energy of 1.7 eV. The solid line shows a t using (11). The inset demonstrates the dependence of the QD ground state decay time on excitation density for a single dot layer (uncoupled) and a twofold stacked dot layer (coupled) with a 1.5-nm-thick GaAs barrier of dots grown by nominally 2ML InAs deposition.

of the enhanced oscillator strength of the excited QD states and the bimolecular recombination kinetics of WL and GaAs barrier states [25]. From the streak camera recording of the spectrally integrated QD luminescence after short pulse excitation we deduce the rise time and the decay times of the ground (excited) state QD PL using a convolution of the QD PL signal [according (11)] [33] and of the excitation puls: (11) One gets a rise time for the QD luminescence of 20 10 ps which is almost the same as the rise time of 13 ps for the exciation pulse indicating that the QD PL rise time can be lower than 20 ps. The fast rise time of the QD luminescence after barrier excitation is attributed to fast capture of carriers from the barrier into the QDs. Therefore, the spontaneous recombination lifetime of several nanoseconds is at least more than two oders of magnitude longer than the carrier capture into the QDs. Thus, the phonon bottleneck problem as described in the introduction may be overcome in selforganized QD structures. The orders of magnitude larger differential gain of QD lasers should lead to an increased modulation bandwidth. Small-signal analysis of standard rate equations yields an

MODULATION BANDWIDTH

The time behavior of the ground state and excited state QD luminescence after short pulse excitation is shown in Fig. 10. The monoexponential decay of the ground state luminescence intensity over more than one order of magnitude proves the excitonic nature of the QD recombination. We do not observe any dependence of the ground state decay time on the excitation density. But the decay time reduces from 2.5 ns down to 1 ns by decreasing the GaAs barrier width from 5 nm to 1.5 nm between adjacent QD layers. This result may be explained by a higher oscillator strength [20] and a larger gain of the excited QD states compared to the ground state recombination (see Fig. 10). Additionally the luminescence decay time spectrally integrated over the QD, WL, and GaAs barrier emission decreases with excitation density because

BIMBERG et al.: InGaAsGaAs QUANTUM-DOT LASERS

203

(a)
0 Fig. 12. (a) Differential gain g 0 and differential refractive index nreal for a quasi-ideal QD ensemble with a symmetric gain curve and only one energy level for electrons and holes. (b) Calculated linewidth enhancement factor for the quasi-ideal QD ensemble and the experimentally reported single-layer QD ensemble with an asymmetric gain spectrum shown in Fig. 3.

(b) Fig. 11. (a) Relaxation oscillation frequency !R of a QD laser as a function of the square root of pulse power in the laser cavity for two different cavity lengths. The excitation was done by a square pulse of 5-ns length and rise-time below 100 ps. The ridge waveguide laser structure (ridge 5 m) consists of a six-fold stacked QD layer (GaAs barrier width 5 nm) with dots grown by nominall 2ML InAs deposition. (b) Calculated modulation bandwidth for different relaxation oscillation frequencies and capture times according to (12) using the parameters given in Fig. 11(a).

approximation of the modulation bandwidth dened as the 3dB value of the ratio of photon density and current density [34]:

(12a) (12b) (12c) (12d) Here, is the modulation frequency, the QD capture time, the QD carrier decay time, the photon roundtrip time in the laser cavity, the differential material gain, the connement factor, the group velocity in the laser, the gain compression coefcient, the photon density in the laser cavity at the bias level, the ratio of QD capture and emission time, the proportionality constant between the damping factor and the square of the relaxation . By measuring the relaxation osciloscillation frequency lation frequency as a function of the photon density (i.e., the power in the laser cavity) one can determine experimentally

the differential gain and the gain compression coefcient which is equal to the inverse of the saturation photon density in the laser resonator. From these values the coefcient -factor is obtained. If and remain linear functions of power the maximum modulation bandwidth is given by . The carrier density used for the calculation of the differential gain in Fig. 11 is related to the recombination volume dened by the lower and upper cladding layer of the laser structure (see Fig. 1). The experimental dependence of the relaxation oscillation frequency on pulse power shows a deviation from the expected proportionality [see (12)] to the square root of optical pulse power indicating the onset of gain compression at a power level of 50 mW. The gain compression coefcient has a minimum of 4 10 cm at 90 K equivalent to a power saturation coefcient of 100 mW in the laser cavity. The gain compression coefcient is more than one order of magnitude higher than typical values ( cm ) in InGaAsInGaAlAs QW lasers [35]. The differential gain of 2 10 cm obtained from the t according to (12) is close to the value derived in the previous section from the QD parameters. Since the gain compression is closely correlated to the capture time of carriers into the dots the physical origin of the enhanced gain compression in QD laser structures may be connected to the slower relaxation rate into the QDs. This mechanism reduces the modulation bandwidth even when the differential gain is as high as shown in Fig. 3. The measured -factor of 0.38 ns (corresponding to a maximum modulation bandwidth of 23 GHz) is a factor of 23 higher than best values reported for InGaAs QW [34] lasers at 100 K indicating that the modulation bandwidth is lower in present QD lasers than in QW lasers. VII. CHIRP Potential applications of a QD laser as directly modulated light source for data transmission via glass bers requires

204

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 3, NO. 2, APRIL 1997

minimum chirp. The chirp is determined from the shift of the wavelength a laser exhibits during current modulation. The physical origin of this shift is related to the coupling of the real and imaginary part of the complex susceptibility in the laser medium [see (1)]. A gain (imaginary part of the susceptibility) variation due to a change of the injection level changes the refractive index (real part of the susceptibility) which modies the phase of the optical mode in the laser cavity. The coupling strength between real and imaginary parts is dened by the linewidth enhancement factor which is dened as follows:

bandgap engineering of the laser structure. The QD emission wavelength can be tuned up to 1.5 m, which is important for ber-optic communication. The linewidth enhancement factor of currently operating laser is reduced due to a nearly symmetric gain spectrum to a value of 0.5, which is a factor of 3 better than in current QW lasers. REFERENCES
[1] Y. Arakawa and H. Sakaki, Multidimensional quantum well lasers and temperature dependence of its threshold current, Appl. Phys. Lett., vol. 40, pp. 939941, June 1982. [2] M. Asada, Y. Miyamoto, and Y. Suematsu, Gain and the threshold of three-dimensional quantum-box lasers, IEEE J. Quantum Electron., vol. QE-22, pp. 19151921, Sept. 1986. [3] K. J. Vahala, Quantum box fabrication tolerance and size limits in semiconductors and their effect on optical gain, IEEE J. Quantum Electron., vol. 24, pp. 523530, Mar. 1988. [4] M. Grundmann, J. Christen, N. N. Ledentsov, J. B ohrer, D. Bimberg, S. S. Ruvimov, P. Werner, U. Richter, U. G osele, J. Heydenreich, V. 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, May 1995. [5] H. Benisty, C. M. Sotomayor-Torres, and C. Weisbuch, Intrinsic mechanism for the poor luminescence properties of quantum-box systems, Phys. Rev. B, vol. 44, pp. 1094510948, Nov. 1991. [6] 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. 1196911980, Oct. 1995. [7] R. Heitz, M. Grundmann, N. N. Ledentsov, L. Eckey, M. Veit, D. Bimberg, V. M. Ustinov, A. Yu. Egorov, A. E. Zhukov, P. S. Kopev, and Zh. I. Alferov, Multiphononrelaxation processes in self-organized InAs/GaAs quantum dots, Appl. Phys. Lett., vol. 68, pp. 361363, Jan. 1996. [8] M. Grundmann and D. Bimberg, Gain and threshold of quantum dot lasers: Theory and comparison to experiments, Jpn. J. Appl. Phys., vol. 36, pp. 3642, June 1997. [9] M. Grundmann and D. Bimberg, Theory of random population for quantum dots, Phys. Rev. B., vol. 55, pp. 97409745, Apr. 1997. [10] H. Hirayama, K. Matsunaga, M. Asada, and Y. Suematsu, Lasing action of Ga0:67 In0:33 As/GaInAsP/InP tensile-strained quantum-box lasers, Electron. Lett., vol. 30, pp. 142143, Jan. 1994. [11] N. Kirstaedter, N. N. Ledentsov, M. Grundmann, D. Bimberg, V. M. Ustinov, S. S. Ruvimov, M. V. Maximov, P. S. Kopev, and Zh. I. Alferov, Low threshold, large T0 injection laser emission from (InGa)As quantum dots, Electron. Lett., vol. 30, pp. 14161417, Aug. 1994. [12] D. Bimberg, N. N. Ledentsov, M. Grundmann, N. Kirstaedter, O. G. Schmidt, M. H. Mao, V. M. Ustinov, A. Yu. Egorov, A. E. Zhukov, P. S. Kopev, Zh. I. Alferov, S. S. Ruvimov, U. G osele, and J. Heydenreich, InAs-GaAs quantum pyramid lasers: In situ growth, radiative lifetimes and polarization properties, Jpn. J. Appl. Phys., vol. 35, pp. 13111319, Feb. 1996. [13] N. N. Ledentsov, V. A. Shchukin, M. Grundmann, N. Kirstaedter, J. B ohrer, O. G. Schmidt, D. Bimberg, V. M. Ustinov, A. Yu. Egorov, A. E. Zhukov, P. S. Kopev, S. V. Zaitsev, N. Yu. Gordeev, and Zh. I. Alferov, Direct formation of vertically coupled quatum dots in stranski-krastanow growth, Phys. Rev. B, vol. 54, pp. 87438750, Sept. 1996. [14] A. Moritz, R. Wirth, A. Hangleiter, A. Kurtenbach, and K. Eberl, Optical gain and lasing in self-organized InP/GaInP quantum dots, Appl. Phys. Lett., vol. 69, pp. 212214, July 1996. [15] F. Hatami, N. N. Ledentsov, M. Grundmann, J. B ohrer, F. Heinrichsdorff, M. Beer, D. Bimberg, S. S. Ruvimov, P. Werner, U. G osele, J. Heydenreich, S. V. Ivanov, B. Ya. Meltser, P. S. Kopev, and Zh. I. Alferov, Radiative recombination in type-II GaSb/GaAs quantum dots, Appl. Phys. Lett., vol. 67, pp. 656658, July 1995. [16] E. R. Glaser, B. R. Benett, B. V. Shanabrook, and R. Magno, Photoluminescence studies of self-assembled InSb, GaSb, and AlSb quantum dot heterostructures, Appl. Phys. Lett., vol. 68, pp. 36143616, June 1996. otzel, and T. [17] J. Temmyo, E. Kuramochi, M. Sugo, T. Nishiya, R. N Tamamura, Quantum disk lasers with self-organized dot-like active regions, in Proc. Lasers and Electro-Opt. Soc. (LEOS95), 1995, vol. 1, pp. 7778.

(13) is the carrier density, are the real and Here, imaginary parts of the complex refractive index, is the the velocity of light in vacuum, is the photon energy, material gain, the differential gain, and the differential refractive index. The linewidth enhancement factor can be calculated from the gain spectrum via the KramersKronig (KK) relation. In the case of a QD laser with a dot ensemble showing a perfect Gaussian energy distribution and only one energy level for electrons and for holes the gain spectrum is perfectly symmetric around the peak gain energy. In this case, the differential gain is also symmetric to the peak gain energy position where lasing occurs. Thus the differential refractive index change computed via the KK relation is exactly zero at the lasing energy (i.e., peak gain position). Therefore the linewidth enhancement factor is zero at the lasing energy and QD lasers with a quasi ideally distributed (symmetric) ensemble of QDs exhibit chirpfree operation. This property is shown in Fig. 12. In the case of real QD lasers the contribution of excited states might cause a non symmetric gain curve (see Fig. 3) which increases the linewidth enhancement factor depending on excitation level to about 0.5. This value is considerably lower than values of 12 [36], which are considered as being very good for QW laser structures. VIII. CONCLUSION We have described lasing characteristics of (InGa)As QD lasers. Already the rst preliminary results indicate that most of the laser key parameters are signicantly improved as compared to QW lasers due to the introduction of QDs as an active medium. The single layer QD structures exhibit an ultrahigh material gain of 10 cm only limited by the dot size distribution. A threshold density of 2 A cm for a laser structure with a QD ensemble having a gain width of 50 meV and no current leakage is predicted. The threshold current density of currently operating lasers is as low as 40 A cm at 77 K and 62 A cm at 300 K while a of 385 K is obtained up to room temperature. The main current leakage is connected to barrier population and nonradiative loss in the barrier. Both effects can be suppressed in the future by growth optimization and careful

BIMBERG et al.: InGaAsGaAs QUANTUM-DOT LASERS

205

[18] H. Shoji, K. Mukai, N. Ohtsuka, M. Sugawara, T. Uchida, and H. Ishikawa, Lasing at three-dimensionally quantum-conned sublevel of self-organized In0:5 Ga0:5 As quantum dots by current injection, IEEE Photon. Technol. Lett., vol. 7, pp. 13851387, Dec. 1995. [19] R. Mirin, A. Gossard, and J. Bowers, Room temperature lasing from InGaAs quantum dots, in Proc. Int. Conf. Quantum Devices and Circuit, 1996. [20] K. Kamath, P. Bhattacharya, T. Sosnowski, T. Norris, and J. Phillips, Room temperature operation of In0:4 Ga0:6 As/GaAs self-organized quantum dot lasers, Electron. Lett., vol. 32, pp. 13741375, July 1996. [21] F. Heinrichsdorff, M.-H. Mao, N. Kirstaedter, A. Krost, D. Bimberg, A. O. Kosogov, and P. Werner, Room temperature lasing from vertically stacked InAs/GaAs quantum dots grown by metalorganic chemical vapor deposition, Appl. Phys. Lett., to be published. [22] N. Kirstaedter, O. G. Schmidt, N. N. Ledentsov, D. Bimberg, V. M. Ustinov, A. Yu. Egorov, A. E. Zhukov, M. V. Maximov, P. S. Kopev, and Zh. I. Alferov, Gain and differential gain of single layer InAs/GaAs-quantum dot injection lasers, Appl. Phys. Lett., vol. 69, pp. 12261228, Aug. 1996. [23] O. G. Schmidt, N. Kirstaedter, N. N. Ledentsov, M.-H. Mao, D. Bimberg, V. M. Ustinov, A. Y. Egorov, A. E. Zhukov, M. V. Maximov, P. S. Kopev, and Zh. I. Alferov, Prevention of gain saturation by multilayer quantum dot lasers, Electron. Lett., vol. 32, pp. 13021303, July 1996. [24] H. Nakayama and Y. Arakawa, Calcuation of lasing characteristics in quantum dot lasers considering interaction of electrons with LO phonons, in Proc. 15th IEEE Int. Semiconductor Laser Conf. Haifa, Israel, 1996, pp. 4142. [25] E. H. B ottcher, N. Kirstadter, M. Grundmann, and D. Bimberg, Nonspectroscopic approach to the determination of the chemical potential and band-gap renormalization in quantum wells, Phys. Rev. B, vol. 45, pp. 85358541, Apr. 1992. [26] Y. L. Lau, Ultalow threshold quantum well lasers, in Quantum Well Lasers, P. S. Zory, Ed. San Diego, CA: Academic, pp. 189215, 1993. [27] N. Chand, E. E. Becker, J. P. van der Ziel, S. N. G. Chu, and N. K. Dutta, Excellent uniformity and very low ( 50 Acm02 ) threshold current density strained quantum well diode lasers on GaAs substrate, Appl. Phys. Lett., vol. 58, pp. 17041706, Apr. 1991. [28] Zh. I. Alferov, N. Yu. Gordeev, S. V. Zaitsev, P. S. Kopev, I. V. Kochnev, V. V. Komin, I. L. Krestnikov, N. N. Ledentsov, A. V. Lunev, M. V. Maximov, S. S. Ruvimov, A. V. Sakharov, A. F. Tsapulnikov, and Yu. M. Shernyakov, A low threshold injection heterojunction laser based on quantum dots, produced by gase-phase epitaxy from organometallic compounds, Semiconductors, vol. 30, pp. 197200, Feb. 1996. [29] N. N. Ledentsov, Ordered arrays of quantum dots, in Proc. 23th Int. Conf. Semiconductors (ICPS96), 1996, pp. 1925. [30] V. M. Ustinov, A. Yu. Egorov, A. R. Kovsh, A. E. Zhukov, M. V. Maximov, A. F. Tsatsulnikov, N. Yu. Gordeev, S. V. Zaitsev, Yu. M. Shernyakov, N. A. Bert, P. S. Kopeev, Zh. I. Alferov, N. N. Ledentsov, J. B ohrer, D. Bimberg, A. O. Kosogov, P. Werner, and U. G osele, Lowthreshold injection lasers based on vertically coupled quantum dots, in Proc. 9th Int. Conf. MBE, 1996. [31] G. Zhang, High power and high efciency operation of Al-free InGaAs/GaInAsP/GaInP GRINSCH SQW lasers ( 0.98 m), Electron. Lett., vol. 30, pp. 12301232, July 1994. [32] F. Heinrichsdorff, A. Krost, M. Grundmann, D. Bimberg, A. O. Kosogov, and P. Werner, Self-organization process of InGaAs/GaAs quantum dots grown by metalorganic chemical vapor deposition, Appl. Phys. Lett., vol. 68, pp. 32843286, June 1996. [33] H.-J. Polland, K. Leo, K. Rother, K. Ploog, J. Feldmann, G. Peter, E. O. G obel, K. Fujiwara, T. Nakayama, and Y. Ohta, Trapping of carriers in single quantum wells with different conguration of the connement layers, Phys. Rev. B, vol. 38, pp. 76357647, Oct. 1988. [34] R. Nagarajan, I. Ishikawa, T. Fukushima, R. Geels, and J. E. Bowers, High speed quantum well-lasers and carrier transport effects, IEEE J. Quantum Electron., vol. 28, pp. 19902008, Oct. 1992.

[35] A. Grabmeier, A. Hangleiter, G. Fuchs, J. E. A. Whiteaway, and R. W. Glew, Low nonlinear gain in InGaAs/InGaAlAs seperate connement multiquantum well lasers, Appl. Phys. Lett., vol. 59, pp. 30243026, Dec. 1991. [36] R. Raghuraman, N. Yu, R. Engelmann, H. Lee, and C. L. Shieh, Spectral dependence of the differential gain, mode shift, and linewidth enhancement factor in a InGaAs-GaAs strained-layer single-quantumwell laser operated under high-injection conditions, IEEE J. Quantum Electron., vol. 29, pp. 6975, Jan. 1993.

D. Bimberg (M92) was born in Schrozberg, Germany, on July 10, 1942. He received the Diploma in physics and the Ph.D. degree from the Goethe University, Frankfurt, Germany, in 1968 and 1971, respectively. From 1972 to 1979, he was a Senior Scientist with the Max Planck Institute for Solid State Research. From 1979 to 1981, he was an Associate Professor with the Department of Electrical Engineering, University of Aachen, Aachen, Germany. He presently holds the Chair of Applied Solid State Physics at the Technical University of Berlin, Berlin, Germany. He authored more than 300 papers, patents, and books. His research interests include the physics of microstructures and microstructures devices, integrated optics, high-speed modulation of semiconductor lasers, and transition metals in IIIV materials.

<

N. Kirstaedter was born in Berlin, Germany, in 1964. He received the M.S. and the Ph.D. degrees in physics from the Technical University of Berlin, Germany, in 1990 and 1996, respectively. He is presently working on fabrication, characterization, and modeling of QD lasers.

N. N. Ledentsov, photograph and biography not available at the time of publication.

Zh. I. Alferov, photograph and biography not available at the time of publication.

P. S. Kopev, photograph and biography not available at the time of publication.

V. M. Ustinov, photograph and biography not available at the time of publication.

Anda mungkin juga menyukai