You are on page 1of 11

952

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

Carrier Dynamics and High-Speed Modulation Properties of Tunnel Injection InGaAsGaAs Quantum-Dot Lasers
Pallab Bhattacharya, Fellow, IEEE, Siddhartha Ghosh, Student Member, IEEE, Sameer Pradhan, Jasprit Singh, Member, IEEE, Zong-Kwei Wu, J. Urayama, Kyoungsik Kim, and Theodore B. Norris, Member, IEEE

AbstractWe have performed pump-probe differential transmission spectroscopy (DTS) measurements on In0 4 Ga0 6 AsGaAsAlGaAs heterostructures, which show that at room temperature, injected electrons preferentially occupy the excited states in the dots and states in the barriers layers. The relaxation time of these carriers to the dot ground state is 100 ps. This leads to large gain compression in quantum-dot (QD) lasers and limits the attainable small-signal modulation 57 GHz. The problem can be alleviated by bandwidth to tunneling cold electrons into the lasing states of the dots from an adjoining injector layer. The design, growth, and steady-state and small-signal modulation characteristics of tunnel injection In0 4 Ga0 6 AsGaAs QD lasers are described and discussed. The tunneling times, directly measured by three-pulse DTS measurements, are 1.7 ps and independent of temperature. The measured small-signal modulation bandwidth for 7 th is = 23 GHz and the gain compression factor for this 3 frequency response is = 8 2 10 16 cm3 . The differential gain obtained from the modulation data is = 2 7 10 14 cm2 at room temperature. The value of the -factor is 0.205 ns and the maximum intrinsic modulation bandwidth is 43.3 GHz. Analysis of the transient characteristics with appropriate carrier and photon rate equations yield modulation response characteristics identical to the measured ones. The Auger coefficients are in the range 3.3 10 29 cm6 s to 3.8 10 29 cm6 s in the temperature range 15 C 85 C, determined from large-signal modulation measurements, and these values are smaller than those measured in separate confinement heterostructure QD lasers. The measured high-speed data are comparable to, or better than, equivalent quantum-well lasers for the first time. Index TermsCarrier dynamics, high-speed, quantum dots, semiconductor lasers.

I. INTRODUCTION UNCTION lasers with quantum dots (QDs) in the active region were demonstrated nearly a decade ago [1][5]. Since the first demonstration, great strides have been made in enhancing the performance characteristics of these unique devices. The coherently strained and relatively defect-free QDs are formed by the self-organization process when the lattice mismatch exceeds 1.8% [6][10]. Self-organized quantum-dot (QD) lasers are grown by metal-organic vapor phase epitaxy (MOVPE) or molecular beam epitaxy (MBE). Significant
Manuscript received November 4, 2002; revised March 10, 2003. The work was supported by the Army Research Office under Grant DAAD019-01-1-0331. The authors are with the Department of Electrical Engineering and Computer Science, University of Michigan, Ann Arbor, MI 481092122 USA. Digital Object Identifier 10.1109/JQE.2003.814374

milestones in the development of the QD lasers include demonstration of low threshold at room temperature [11], large differential gain [12], [13], high output power [14], wide spectral tunability [15], and better temperature insensitivity of the threshold current [16], [17] than quantum-well lasers. However, it is now clear that the carrier dynamics in QDs do not favor high-speed operation, such as direct modulation, of lasers [13], [18][22]. More specifically, detailed two- and three-pulse pump-probe DTS measurements made by us on QD heterostructures show that at temperatures above 100 K, a large number of injected electrons preferentially occupy the higher lying states of the dots and states in the adjoining barrier/wetting layer with higher density [23], [24]. Additionally a phonon bottleneck has also been observed in the same QD heterostructures, when electrons and holes are nongeminately captured (in different dots) [25]. Therefore, there exists a significant hot-carrier problem in QD lasers. We attribute this to be the principal reason for the inability to modulate QD lasers at high speeds. As the temperature is lowered, the hot carrier distribution is minimized and carriers relax from the higher lying states by efficient electron-hole scattering. Indeed, we have recorded small-signal modulation bandwidths of 30 GHz in QD separate confinement heterostructure (SCH) lasers at 80 K, whilst in the same devices, the highest measured GHz at 300 K [18]. The technique of tunnel injection (TI) in semiconductor lasers was proposed and demonstrated almost a decade ago to alleviate problems related to hot carriers in the active region and surrounding layers. By tunneling cold electrons directly into the lasing states of GaAs- and InP-based quantum-well lasers from an adjacent injector layer, the performance characteristics were enhanced and several deleterious effects were minimized [26]. Thus, enhanced modulation bandwidth, higher , reduced Auger recombination, and reduced chirp were demonstrated in 0.98- and 1.55- m lasers [27], [28]. The tunneling concept has also been utilized in vertical-cavity surface-emitting lasers (VCSELS) [29] and light-emitting diodes (LEDs) [30] for improved performance. The TI scheme can be applied to QD lasers, whereby electrons can be transported directly to the lasing states in the dots. Theoretical calculations made by Asryan and Luryi [31] indicate that large values of may be obtained in QD-TI lasers due to a minimization of carrier leakage from the active region and reduced parasitic recombination in the optical confinement layer and hot-carrier effects in the devices. By careful engineering of

0018-9197/03$17.00 2003 IEEE

BHATTACHARYA et al.: CARRIER DYNAMICS AND HIGH-SPEED MODULATION PROPERTIES OF TI InGaAsGaAs QD LASERS

953

(a)

Fig. 2. Calculated energy levels in a In 8-band k.p model.

Ga

AsGaAs QD system using

(b) Fig. 1. Carrier injection into a QD gain region in (a) a conventional separate confinement heterostructure laser and (b) by TI into the dot.

the QD heterostructure, we have demonstrated large modulation bandwidths in InGaAsGaAs self-organized QD lasers at room temperature, for the first time [32]. These lasers also demon0.5 ), -factors 1 [32], and strate ultra-low chirp ( as high as 363 K around room temperature [33]. Advantages of using the tunneling scheme have also been reported for self-organized InPInAlP QD lasers [34]. In what follows, we will first briefly review the concept of TI, followed by the results obtained from femtosecond pump-probe spectroscopy on QD heterostructures. The design, growth, fabrication, and characterization of the QD-TI lasers will be next described. The modulation characteristics are analyzed with the appropriate carrier and photon rate equations and the calculated results are in good agreement with the measured ones. Finally, large-signal modulation experiments and the determination of the Auger coefficients in these lasers are described. II. TI IN LASERS The carrier-injection process in a conventional separate confinement heterostructure (SCH) QD laser is depicted in Fig. 1(a). Electrons and holes are injected from the cladding layer regions into the active (guiding) region containing the QD layers. For efficient laser performance, the carriers should be able to enter the QDs and reach thermal equilibrium with the quasi-Fermi level. If thermalization into the dot lasing states does not occur in any reasonable amount of time ( stimulated emission time constant), the carrier distribution will become hot. While injected carriers lose energy and fill the lasing states, carrier heating simultaneously forces them out toward higher energies and causes leakage to adjoining layers. Recombination in these regions is believed to be the principal source of the temperature dependence of the threshold

current of lasers and small values of the coefficient [31]. The hot carrier distribution also leads to gain compression at the lasing energy, increased threshold current, increased Auger recombination, reduced small-signal modulation bandwidth, and higher chirp and linewidth-enhancement factor. In addition to quantum capture, diffusion of injected carriers has also to be included in considering the dynamics of injected carriers. The tunneling-injection scheme for electrons is illustrated in Fig. 1(b). If cold electrons are introduced into the QD ground states by direct or phonon-assisted tunneling, and the tunneling rate is comparable to the stimulated emission rate, the carrier distribution will be maintained close to a quasi-Fermi distribution, even at high injection levels. Thus carrier leakage and hot-carrier effects are minimized. It is important to note that the hole relaxation rates in QDs and other quantum confined structures are very fast due to multiplicity of the levels, band mixing and efficient hole-phonon coupling. In the tunneling injection scheme, fewer electrons will bypass the active region during the injection process. Gain compression effects will be minimized, since most of the electrons are injected close to the lasing energy. Additionally, the tunneling process acts as a filter and selects the dots whose ground states coincide with the tunneling energy level. We have observed that the inhomogeneously broadened linewidth of self-organized QD photoluminescence spectra, caused by the size distribution of the dots, is reduced by a factor of two in QD TI heterostructures [32]. Since the initial carrier distribution injected by tunneling has a smaller energy spread, the spectral purity of the laser is expected to improve. This has also been observed in quantum-well TI lasers. III. CARRIER DYNAMICS IN SELF-ORGANIZED QDs Self-organized In Ga AsGaAs QDs are formed on GaAs in the Stranski-Krastanow growth mode after the formation of an initial two-dimensional wetting layer. For a typical growth, the wetting layer is 56 monolayers (MLs) thick and the dot consists of 1 ML. Multiple coupled dot layers, separated by 1525 GaAs barriers, are generally incorporated in laser heterostructures. The typical density in a dot layer can be varied and cross-sectional high-resolution from (1-10) 10 cm transmission electron microscopy (HR XTEM) studies have confirmed that the individual dots are pyramidal to lens-like

954

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

(a)

(b)

(c)

Fig. 3. (a) Excited-state (n = 2) DT time scan in a In Ga AsGaAs QD at 10 K. (b) Three-pulse DT time scan corresponding to the gain recovery in the ground state of a In Ga AsGaAs QD; (c) DT time scan of the wetting layer and barrier region ( 20 meV of GaAs band edge) at various temperatures.

in shape with a base dimension of 1020 nm and a height of 68 nm. Theoretical bandstructure calculations [35] using an 8-band k.p description of both conduction and valence band states and a variety of experimental techniques, including photoluminescence and electroluminescence, reveal that there are a number of near-degenerate hole states and a few discrete electron states. Of these, the dominant ground and first excited states are separated by 50100 meV, depending on dot material, composition, heterostructure used, and growth parameters. Since the separation between these bands is greater than the longitudinal-optical (LO) phonon energy, the LO phonon scattering mechanism, that is mainly responsible for rapid carrier relaxation in bulk semiconductors and quantum wells, is suppressed. Therefore, we expect longer carrier relaxation times in QDs. The excited level in each dot has a twofold spatial degeneracy due to the symmetry of the dot geometry, in addition to the double-spin degeneracy. Fig. 2 shows the calculated energy levels in a In Ga AsGaAs QD. Two- and three-pulse femtosecond pump-probe DTS measurements, described in Section IV-B, were made on In Ga AsGaAs QD heterostructures, similar to SCH lasers, at different temperatures and for a range of excitation levels. The results of these experiments have been reported by us previously [23][25] and some relevant data are described and discussed here in the context of the present study. The differential transmission signal is proportional to the carrier population of the level probed. The carrier dynamics in the

QDs is therefore reflected in the transient of the differential transmission signal for the ground and excited states. Fig. 3(a) shows the excited state ( ) differential transmission signal for weak excitation ( 1 carrier per dot) in the GaAs barrier K. The population of the excited state shows a fast at decay, followed by a very slow one. We have recorded similar data at temperatures upto 300 K [23]. The fast excited-state ps and this component to ground-state relaxation is decreases and becomes slower as the sample temperature is raised. The slow component is due to nongeminate capture of electrons and holes amongst the dots and is observed when the number of injected carriers is smaller than the number of accessible dots. We have confirmed that the fast intersubband relaxation at low temperatures is mediated by electron-hole scattering [20]. As the temperature is raised, the rate of this process decreases and the intersubband relaxation rate also decreases. In practice, lasers are operated at room temperature and high injection levels. In the self-organized QDs, the localized QD states and the two-dimensional wetting layer states form an electronically coupled system and the number of wetting layer/barrier states is much larger than the number of available dot states at 300 K. Under these conditions, the injected electrons predominantly reside in the wetting layer and barrier states and the system cannot be described by equilibrium quasi-Fermi statistics. Evidence of injected electrons residing in higher lying states of the dot heterostructure are shown in Fig. 3(b) and (c), in which Fig. 3(b) is the three-pulse

BHATTACHARYA et al.: CARRIER DYNAMICS AND HIGH-SPEED MODULATION PROPERTIES OF TI InGaAsGaAs QD LASERS

955

(a)

coupled In Ga As QD layers. The 95 In Ga As injector well is grown at 490 C, the QD layers are grown at 525 C and the rest of the structure is grown at 620 C. The conduction band profile in the active region is shown in Fig. 4(b). A critical aspect of the design is the alignment of the conduction band states of the injector layer with the bound states in the QDs, such that efficient phonon-assisted tunneling can take place from the injector layer to the dots. The bandgap of the injector layer is tuned for the optimum tunneling conditions and the tunneling rates are measured, as described in the next section. The wavelength of the luminescence peak for the dots is controlled by adjusting the InGaAs dot size during epitaxy such that the energy separation, in the conduction band, between the injector well states and the QD ground states is 36 meV at room temperature. This energy separation ensures LO phonon-assisted tunneling from the injector well to the dot ground states through the 20 undoped Al Ga As barrier layer. The injected holes rapidly thermalize to the dot ground states and the Al Ga As barrier prevents hole injection and recombination in the injector layer. Single-mode ridge waveguide lasers of cavity lengths varying from 2001300 m were fabricated by standard lithography, wet and dry etching, and metallization techniques. The devices were mounted on Cu heat sinks for the measurements. The cleaved facets were left uncoated for the measurements reported here. B. Measurement of Phonon-Assisted Tunneling Times Three-pulse pump-probe DTS measurements were made on tunneling heterostructures, which were identical to the lasers, except that all the layers were undoped. They were also grown on a GaAs buffer layer. The sample was anti-reflection coated, mounted on sapphire, and the GaAs substrate was removed by selective etching to eliminate any possible contribution of the substrate to the optical spectra. Measurements were made in the temperature range 10 K100 K. The photoluminescence spectra from the sample, measured at 12 K with 800-nm excitation, is shown in Fig. 5(a). A dominant emission from the QD with a peak at 980 nm and a weak shoulder, believed to be from the injector layer, at 950 nm are seen in the data. As stated earlier, a characteristic feature of the PL from TI QD heterostructures is the extremely small linewidth ( 23 meV) compared to normal QDs grown under identical conditions, for which the linewidth varies in the range 4060 meV. This is due to a selection process of the tunneling states in the dots, thereby artificially inducing dot uniformity for luminescence. The principle of the DT measurement is outlined in Fig. 5(b). A 100-fs 250-kHz amplified Ti:Sapphire laser is used to generate a white-light source from which 10-nm wide pulses are spectrally selected for the gain (800 nm), pump (980 nm), and probe pulses. The pump is tuned to generate carriers either in the GaAs barrier states or resonantly in the excited states of the dots. For DT spectral scans, the probe pulse consists of a dispersion-compensated near-infrared band between 820 and 1050 nm selected with a long-pass filter. Using a monochromator, we resolve the DT signal to 1 nm and detect the probe signal with a lock-in amplifier referenced to the 6-kHz mechanically chopped pump. The DT signal measures the change in the carrier occupation of the levels that are in resonance with the probe spectrum.

(b) Fig. 4. (a) TI QD laser heterostructure grown by molecular beam epitaxy. (b) Conduction band profile under flat-band conditions.

differential transmission signal corresponding to gain recovery in a laser heterostructure. A gain pulse to establish interband population inversion is followed 14 ps later by a pump pulse to deplete the electron-hole pairs in the ground state. The initial fast recovery is due to the excited- to ground-state transition and the slow component is due to carrier capture from higher lying states. Fig. 3(c) shows the differential transmission signal at GaAs barrier energy 20 meV with the pump tuned to the barrier energy. It is evident that as the temperature is increased, most of the carriers remain in the barrier and the dot excited states [23]. Therefore, a reduction in the rate of electron-hole scattering in the dots (and phonon scattering is not operative) and the hot carrier distribution lead to severe gain saturation at the QD ground-state lasing energy. This has also been recently verified experimentally by Mathews et al. [36]. The tunneling injection scheme should significantly reduce gain compression and carrier leakage to the cladding layers. IV. THE QD TI LASER A. Design and MBE Growth The laser heterostructure, designed with knowledge of the QD electronic states, is shown in Fig. 4(a). The device structure consists of 1.5- m-thick Al Ga As outer cladding layers 10 cm ) and GaAs contact layers on either ( side of the 0.15- m-thick undoped GaAs optical confinement layer (OCL). The active region consists of the In Ga As injector well, a 20 Al Ga As tunnel barrier and three

956

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

(a)

(a)

(b) Fig. 6. (a) Three-pulse DT signal from the TI heterostructure with gain pulse at 800 nm, pump pulse at 980 nm, and probe pulses at 950 nm and 9801000 nm (integrated). (b) Temperature-dependent two-pulse DT signal with the gain pulse at 800 nm and probe pulse at 980 nm.

(b) Fig. 5. (a) Low-temperature photoluminescence spectrum of the TI laser. (b) Principle of tunneling time measurements by three-pulse DTS.

When the pump and probe pulses are delayed with respect to each other, the transient dot-level population is resolved directly. The temperature of the sample is stabilized in a He flow cryostat with a feedback-heater controller. The gain pulse establishes an interband population inversion and sets the quasi-Fermi level in the active region. The pump pulse depletes the QD ground-state population. The carrier dynamics and the population of the various levels are then determined from the transmitted intensity of the probe pulse; a negative DT signal is the signature of gain in the system [24]. Measured DT data at 12 K are shown in Fig. 6(a). The signal at 950 nm corresponds to the injector layer population. The rapid decrease in the DT signal corresponds to phonon-assisted tunneling of electrons from this layer to the QD ground state and the slow recovery of the DT signal reflects the re-population of the layer from higher energy levels by phonon scattering. The initial drop in the 980-nm DT signal is due to pump-induced reduction in the QD ground state via stimulated emission. The fast recovery in the 980 nm DT signal corresponds again to the tunneling of electrons into the dot ground state. Fig. 6(b) shows the

DT signal in a two-pulse experiment (no gain pulse; pump tuned to 950 nm to inject electrons directly into the quantum-well injector layer) for various temperatures; the rise times are almost identical, as expected if the signal was due to tunneling. As mentioned earlier, the phononmediated hole relaxation rate is extremely fast and we have previously calculated and measured 0.6 ps [37]. We estimate the tunthe relaxation time to be neling time constant to be 1.7 ps. Therefore, the phonon-assisted tunneling rate will be comparable to the stimulated emission rate for a high injection rate and the carrier distribution can be maintained close to equilibrium. V. DEVICE CHARACTERISTICS A. DC Characteristics The steady-state characteristics of the lasers have been reported recently [33] and are briefly described here for completeness. The characteristics were measured under pulsed conditions (1 s pulses and 1% duty cycle) as a function of temperature and for various cavity lengths. Typical lightcurrent characteristics for a 400- m-long device is shown in Fig. 7(a). At room mA temperature (288 K), the lasers are characterized by m), A cm ( m), high slope (

BHATTACHARYA et al.: CARRIER DYNAMICS AND HIGH-SPEED MODULATION PROPERTIES OF TI InGaAsGaAs QD LASERS

957

(a) (a)

(b) Fig. 7. (a) Measured lightcurrent characteristics of a 200 m long single-mode TI QD laser at room temperature. (b) Plot of inverse differential quantum efficiency versus cavity length.

(b) Fig. 8. (a) Small-signal modulation response of 400 m long TI laser under varying injection currents at room temperature; (b) the plot of resonance frequency f of the modulation response versus square root of injection current (I I ).

efficiency ( W A for m), and high differential for m). From a plot quantum efficiency ( versus shown in Fig. 7(b), values of internal quantum of and cavity loss coefficient cm efficiency are obtained. From the measured threshold currents at different K for C C and temperatures, values of K for C C are derived. These are measured in these temperature ranges the highest values of in QD lasers. The high device efficiencies and the high values indicate minimization of carrier leakage from the gain reof gion and parasitic recombination in optical confinement layers. B. Small-Signal Modulation Characteristics The small-signal modulation response of the devices was measured with another set of devices at room temperature under cw biasing conditions with a HP 8562A electrical spectrum analyzer, a HP 8350B sweep oscillator, a low noise amplifier, and a New Focus high-speed detector. The frequency response for varying injection currents for a 200 m long laser cavity mA) is shown in Fig. 8(a). The continuous ( lines are best fit curves to the measured data. The spectral outputs at these injection currents confirm that lasing from the GHz ground state is maintained. A bandwidth of mA and this is the highest bandwidth is measured for

measured in any QD laser at room temperature. Fig. 8(b) shows the plot of the resonance frequency of the modulation response as a function of the square root of the output power. The modulation efficiency, which is the slope of this plot, 1.71 GHz mA . From this value of the modulation is efficiency and using a fill factor, which denotes the percent 28 , internal surface area covered by a layer of QDs, of (for the measured device) and quantum efficiency 10 , we derive a value of a confinement factor 10 cm at room temperature. A value ns was calculated using the damping factor of obtained from the best-fit curves to the measured modulation and cavity photon lifetime, response. Using the values of cm . A we get a gain compression factor, of 8.2 10 maximum intrinsic modulation bandwidth of 43.3 GHz is also obtained from this value of . The modulation characteristics of the TI lasers were also analyzed by taking into account the detailed carrier dynamics and the appropriate carrier and photon rate equations. The various transport processes operative in the active region are described in Fig. 9(a). We have separately measured of these time constants directly, as described in the previous sections. The rate

958

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

TABLE I PARAMETER VALUES USED IN NUMERICAL SIMULATION OF THE TI QD LASERS

(a)

(b) Fig. 9. (a) Transport processes in the active region of the TI laser. (b) Modulation response characteristics calculated under small-signal conditions with varying dc bias applied to the laser.

(5) , , , and are the carrier popIn these equations ulations in the injector well, the ground state of the dot, the excited state, and the barrier/wetting layer states, respectively. is the density of the dots per dot layer. The time constants are shown in Fig. 9(a). is the saturation gain corresponding is the photon density, and to the ground-state transition, and are the radiative recombination coefficient in the barrier/wetting layer and injector well, respectively. is the Auger recombination coefficient in the QDs including the barrier/wetting layer. However, at room temperature, the major contribution to this recombination process is from the barrier and wetting layer regions [38]. Furthermore, as shown in Section III, most carriers reside in these regions at room temperature. Therefore, Auger recombination has not been is the Auger recombiexplicitly considered in (1) and (2). and are the nation coefficient in the injector well. widths of the injector well and the wetting layer, respectively. is the number of dot layers in the gain region and is the width of the optical confinement layer. is the cavity loss, is injected current density, and is spontaneous emission lifetime. The parameters mentioned above were determined experimentally, as described in Section III. The Auger recombination coefficients were measured directly from large signal modulation measurements done on QD lasers, to be described in the next section. The values used for the various parameters are listed in Table I. The coupled-rate equations

equations for TI-QD laser with two dot-bound states in the conduction band are

(1)

(2)

(3)

(4)

BHATTACHARYA et al.: CARRIER DYNAMICS AND HIGH-SPEED MODULATION PROPERTIES OF TI InGaAsGaAs QD LASERS

959

were solved by the fourth order RungeKutta method. The modulation response characteristics, shown in Fig. 9(b), were calculated under small-signal conditions with varying dc bias applied to the laser. The calculated modulation response agrees well with the experimentally measured response under similar injection conditions. C. Large-Signal Modulation Characteristics Auger recombination in small-band-gap materials affects the performance of lasers adversely, by increasing the threshold current and damping, and reducing the modulation bandwidth. Auger recombination rates and coefficients can be determined by measuring the turn-on delay of a laser under large-signal modulation. The turn-on dynamics of QD lasers have been theoretically investigated by Grundmann [39], [40] and measured in SCH QD lasers by us [41]. The turn-on delay can be expressed as (6) where is the recombination rate and is expressed as (7) in terms of carrier lifetime , the ShockleyReadHall coeffi, the radiative recombination coefficient , the carcient is the threshold rier density , and the Auger coefficient . carrier density and is the active volume of the laser. By measuring the stimulated emission delay times under large-signal modulation and calculation of the radiative recombination rates and the threshold carrier density, we can determine the Auger coefficient accurately in a self-consistent manner [41], [42]. The large-signal modulation measurements were made on the 400- m-long single-mode ridge waveguide TI lasers at different temperatures. The lasers are pulse biased with a 10% duty cycle to ( ) with 100-ps (20%80%) rise-time elecfrom trical pulses. The output is detected with an InGaAs photoreceiver. The delay time between the electrical and optical signal is measured with a high-speed digital sampling oscilloscope, taking into account the delays due to optical fiber, RF cables, and the photoreceiver. The delay times are measured between the 50% points in the applied electrical pulse and the output optical signal. Fig. 10(a) shows the experimentally determined turn-on delay times as a function of bias current density at different ambient temperatures. The increase in the bias current density decreases the recombination lifetime and consequently the delay time. The gain and spontaneous recombination rates are calculated using the Fermi golden rule with 8-band k.p description [35] is determined using the single-mode couof the bands and pled-rate equations with a measured value of cavity loss (8.2 cm for the 400- m cavity). These parameters are used to determine the Auger coefficient. The carrier distribution in the wetting layers, barrier layers and the injector well are taken into account in order to represent the carrier dynamics in the TI heterostructure more accurately. Recombination at traps is neglected due to the observations of small trap density [43] and negligible Stokes shift [44], [45] in the QD heterostructures. To

(a)

(b) Fig. 10. (a) Measured turn-on delay times in TI QD laser as a function of injection current density at different ambient temperatures. (b) Variation of Auger recombination coefficient with temperature. For comparison, Auger coefficients for a SCH QD laser are also shown.

match the experimental and theoretical threshold current densities at each temperature, the inhomogeneous broadening in the QDs is varied. The Auger coefficient is obtained by minimizing the root mean square error between the theoretical and experiwith mental delay times. The variation of Auger coefficient temperature in a TI QD laser is shown in Fig. 10(b). For comparison purposes, the measured coefficients in a SCH-QD laser [41] are also shown. As described earlier, electron-hole scattering is the dominant mechanism by which electrons relax from the excited state to the ground state in QDs. In this process, an excited-state electron scatters with a ground-state hole, thereby transferring its energy to the hole and relaxing to the ground state. The excited-state hole relaxes back to the ground state by rapid phonon scattering. Energy conservation in the process is achieved by self-consistent broadening of the hole levels due to hole-phonon scattering. The electron in the ground state can then contribute to Auger scattering. As the temperature increases, the population of ground-state holes decreases, due to their excitation to higher energy levels, and the electron-hole scattering and consequently Auger recombination decreases. This trend is observed in the SCH-QD lasers [41]. In the tunneling injection structure, electrons are supplied to the dot ground states directly by tunneling from the adjoining injector layer, and the temperature

960

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

dependence of the Auger recombination rate in the dot and barrier/wetting layer regions would be reduced. This is, indeed, observed in the data shown in Fig. 10(b). The values of the Auger rates in these regions are also lower, reflecting a smaller population of electrons in the dot excited states. VI. CONCLUSION In conclusion, we have demonstrated enhanced performance characteristics in TI In Ga AsGaAs self-organized QD and high slope and differential lasers. Large values of quantum efficiencies in these devices are attributed to the cold carrier injection and distribution maintained by the tunneling scheme in these devices. We have measured extremely fast ps). We also temperature independent tunneling times ( demonstrate extremely high modulation bandwidths, exceeding 20 GHz, at room temperature. The calculated modulation response, using coupled-rate equations, agrees well with the measured modulation response. Auger recombination coefficients at different temperatures have been determined from the turn-on delay time in the TI QD lasers under large signal modulation. The Auger recombination coefficients are 3 10 cm s and the values are smaller than those measured in SCH QD lasers. These data, together with our 0.5 and linewidth-enhancement reported values of chirp at the lasing wavelength [32], make the TI factor QD lasers serious contenders as high-performance sources for optical communication. ACKNOWLEDGMENT The authors thank Prof. D. Bimberg for useful discussions. REFERENCES
[1] N. Kirstaedter, N. N. Ledentsov, M. Grundmann, D. Bimberg, V. M. Ustinov, S. S. Ruvimov, M. V. Maximov, P. S. Kopev, Z. I. Alferov, U. Richter, P. Werner, U. Gsele, and J. Heydenreich, Low threshold, large T injection laser emission from (InGa)As quantum dots, Electron. Lett., vol. 30, pp. 14161417, 1994. [2] K. Kamath, P. Bhattacharya, T. Sosnowski, T. Norris, and J. Phillips, Room temperature operation of In Ga As/GaAs self-organized quantum dots lasers, Electron. Lett., vol. 32, pp. 13741375, 1996. [3] R. Mirin, A. Gossard, and J. Bowers, Room temperature lasing from InGaAs quantum dots, Electron. Lett., vol. 32, pp. 17321733, 1996. [4] H. Shoji, Y. Nakata, K. Mukai, Y. Sugiyama, M. Sugawara, N. Yokogama, and H. Ishikawa, Temperature dependent lasing characteristics of multi-stacked quantum dot lasers, Appl. Phys. Lett., vol. 71, pp. 7173, 1997. [5] D. Bimberg, M. Grundmann, and N. N. Ledenstov, Quantum Dot Heterostructures. Chichester, U.K.: Wiley, 1998. [6] L. Goldstein, F. Glas, J. Y. Marzin, M. N. Charasse, and G. L. Roux, Growth by molecular beam epitaxy and characterization of InAs/GaAs strained layer superlattices, Appl. Phys. Lett., vol. 47, pp. 10991101, 1985. [7] P. R. Berger, K. Chang, P. Bhattacharya, J. Singh, and K. K. Bajaj, Role of strain and growth conditions on the growth front profile of As on GaAs during pseudomorphic growth regime, Appl. In Ga Phys. Lett., vol. 53, pp. 684686, 1988. [8] S. Guha, A. Madhukar, and K. C. Rajkumar, Onset of incoherency and defect introduction in the initial stages of molecular beam epitaxy As on GaAs(100), Appl. Phys. growth of highly strained In Ga Lett., vol. 57, pp. 21102112, 1990. [9] D. Leonard, M. Krishnamurthy, C. M. Reaves, S. P. Denbaars, and P. M. Petroff, Direct formation of quantum-sized dots from uniform coherent islands of InGaAs on GaAs surfaces, Appl. Phys. Lett., vol. 63, pp. 32023204, 1993.

[10] J. Pamulapati, P. K. Bhattacharya, J. Singh, P. R. Berger, C. W. Snyder, B. G. Orr, and R. L. Tober, Realization of in-situ sub two-dimensional quantum structures by strained layer growth phenomena in InGaAs system, J. Electron. Mater., vol. 25, pp. 479483, 1996. [11] G. T. Liu, A. Stintz, H. Li, K. J. Malloy, and L. F. Lester, Extremely low room-temperature threshold current density diode lasers using Ga As quantum well, Electron. Lett., vol. 35, InAs dots in In pp. 11631165, 1999. [12] N. Kirstaedter, O. Schmidt, N. Ledenstov, D. Bimberg, V. Ustinov, A. Egorov, A. Zhukov, M. Maximov, P. Kopev, and Z. I. Alferov, Gain and differential gain of single layer InAs/GaAs quantum dot injection lasers, Appl. Phys. Lett., vol. 69, pp. 12261228, 1996. [13] D. Klotzkin, K. Kamath, K. Vineberg, P. Bhattacharya, R. Murty, and J. Laskar, Enhanced modulation bandwidth (20 Ghz) of In Ga As/GaAs self-organized quantum-dot lasers at cryogenic temperatures: Role of carrier relaxation and differential gain, IEEE Photon. Technol. Lett., vol. 10, pp. 932934, 1998. [14] D. Bimberg, M. Grundmann, F. Heinrichsdorff, N. N. Ledentsov, V. M. Ustinov, A. E. Zhukov, A. R. Kovsh, M. V. Maximov, Y. M. Shernyakov, B. V. Volovik, A. F. Tasulnikov, P. S. Kopev, and Z. I. Alferov, Quantum dot lasers: Breakthrough in optoelectronics, Thin Solid Films, vol. 367, pp. 235249, 2000. [15] P. M. Varangis, H. Li, G. T. Liu, T. C. Newell, A. Stintz, B. Fuchs, K. J. Malloy, and L. F. Lester, Low-threshold quantum dot lasers with 201 nm tuning range, Electron. Lett., vol. 36, pp. 15441545, 2000. [16] M. V. Maksimov, N. Y. Gordeev, S. V. Zaitsev, P. S. Kopev, I. V. Kochnev, N. N. Ledentsov, A. V. Lunev, S. S. Ruvimov, A. V. Sakharov, A. F. Tsatsulnikov, Y. M. Shernyakov, Z. I. Alferov, and D. Bimberg, Quantum dot injection heterolaser with ultrahigh thermal stability of the threshold current up to 50 C, Semiconductors, vol. 31, pp. 124126, 1997. [17] O. B. Shchekin and D. G. Deppe, 1.3 m InAs quantum dot laser with T = 161 K from 0 to 80 C, Appl. Phys. Lett., vol. 80, pp. 32773299, 2002. [18] K. Kamath, J. Phillips, H. Jiang, J. Singh, and P. Bhattacharya, Smallsignal modulation and differential gain of single-mode self-organized In Ga As/GaAs quantum dot lasers, Appl. Phys. Lett., vol. 70, pp. 29522953, 1997. [19] D. Klotzkin, K. Kamath, and P. Bhattacharya, Quantum capture times at room temperature in high-speed In Ga As-GaAs self-organized quantum dot lasers, IEEE Photon. Technol. Lett., vol. 9, pp. 13011303, 1997. [20] P. Bhattacharya, K. K. Kamath, J. Singh, D. Klotzkin, J. Phillips, H. Jiang, N. Chervela, T. B. Norris, T. Sosnowski, J. Laskar, and M. R. Murty, In(Ga)As-GaAs self-organized quantum dot lasers: DC and small-signal modulation properties, IEEE Trans. Electron Devices, vol. 46, pp. 871883, 1999. [21] D. Klotzkin and P. Bhattacharya, Temperature dependence of dynamic and DC characteristics of quantum well and quantum-dot lasers: A comparative study, J. Lightwave Technol., vol. 17, pp. 16341642, 1999. [22] P. Bhattacharya, D. Klotzkin, O. Qasaimeh, W. Zhou, S. Krishna, and D. Zhu, High-speed modulation and switching characteristics in In(Ga)As-GaAs self-organized quantum-dot lasers, IEEE J. Select. Topics Quantum Electron., vol. 6, pp. 426438. [23] J. Urayama, T. B. Norris, H. Jiang, J. Singh, and P. Bhattacharya, Temperature-dependent carrier dynamics in self-assembled InGaAs quantum dots, Appl. Phys. Lett., vol. 80, pp. 21622164, 2002. [24] K. Kim, J. Urayama, T. B. Norris, J. Singh, J. Phillips, and P. Bhattacharya, Gain dynamics and ultrafast spectral hole burning in In(Ga)As self-organized quantum dots, Appl. Phys. Lett., vol. 81, pp. 670672, 2002. [25] J. Urayama, T. Norris, J. Singh, and P. Bhattacharya, Observation of phonon bottleneck in quantum dot electronic relaxation, Phys. Rev. Lett., vol. 86, pp. 49304933, 2001. [26] P. Bhattacharaya, J. Singh, H. Yoon, X. Zhang, A. Gutierrez-Aitken, and Y. Lam, Tunneling injection lasers: A new class of lasers with reduced hot carrier effects, IEEE J. Quantum Electron., vol. 32, pp. 16201629, 1996. [27] X. Zhang, A. Gutierrez-Aitken, D. Klotzkin, P. Bhattacharya, C. Caneau, and R. Bhat, 0.98 m multiple-quantum-well tunneling injection laser with 98-GHz intrinsic modulation bandwidth, IEEE J. Select. Topics Quantum Electron., vol. 3, pp. 309314, 1997. [28] H. Yoon, A. L. Gutierrez-Aitken, R. Jambunathan, J. Singh, and P. Bhattacharya, A cold InP-based tunneling injection laser with greatly reduced Auger recombination and temperature dependence, IEEE Photon. Technol. Lett., vol. 7, pp. 974976, 1995.

BHATTACHARYA et al.: CARRIER DYNAMICS AND HIGH-SPEED MODULATION PROPERTIES OF TI InGaAsGaAs QD LASERS

961

[29] D. L. Huffaker and D. G. Deppe, Improved performance of oxide-confined vertical-cavity surface-emitting lasers using tunnel injection in the active region, Appl. Phys. Lett., vol. 71, pp. 14491451, 1997. [30] T. C. Wen, S. J. Chang, L. W. Wu, Y. K. Su, W. C. Lai, C. H. Kuo, C. H. Chen, J. K. Sheu, and J. F. Chen, InGaN/GaN tunnel-injection blue light-emitting diodes, IEEE Trans. Electron Devices, vol. 49, pp. 10931095, 2002. [31] L. Asryan and S. Luryi, Tunnel-injection quantum-dot laser: Ultra-high temperature stability, IEEE J. Quantum Electron., vol. 37, pp. 905910, 2001. [32] S. Ghosh, S. Pradhan, and P. Bhattacharya, Dynamic characteristics of high-speed In Ga As/GaAs self-organized quantum dot lasers at room temperature, Appl. Phys. Lett., vol. 81, no. 3055, 2002. [33] S. Pradhan, S. Ghosh, and P. Bhattacharya, Temperature dependent steady-state characteristics of high-performance tunnel injection quantum dot lasers, Electron. Lett., vol. 38, no. 1449, 2002. [34] G. Walter, N. Holonyak Jr., J. H. Ryou, and R. D. Dupuis, Coupled InP quantum-dot InGaP quantum well InP-InGaPIn(AlGa)P-InAlP heterostructure diode laser operation, Appl. Phys. Lett., vol. 79, pp. 32153217, 2001. [35] H. Jiang and J. Singh, Strain distribution and electronic spectra of InAs/GaAs self-assembled dots: An eight-band study, Phys. Rev. B, vol. 56, pp. 46964701, 1997. [36] D. R. Matthews, H. D. Summers, P. M. Smowton, and M. Hopkinson, Experimental investigation of the effect of wetting-layer states on the gain-current characteristic of quantum-dot lasers, Appl.Phys. Lett., vol. 81, pp. 49044906, 2002. [37] T. Sosnowski, T. Norris, H. Jiang, J. Singh, K. Kamath, and P. Bhattacharya, Rapid carrier relaxation in In Ga As/GaAs quantum dots characterized by differential transmission spectroscopy, Phys. Rev. B, vol. 57, pp. R9423R9426, 1998. [38] H. Jiang, Strain Induced Physical Phenomena in InGaAs/GaAs and InGaN/GaN Heterostructures, PhD, Univ. Michigan, Ann Arbor, 1999. [39] M. Grundmann, How a quantum-dot laser turns on, Appl. Phys. Lett., vol. 77, pp. 14281430, 2000. [40] , Feasibility of 5 Gbit/s wavelength division multiplexing using quantum dot lasers, Appl. Phys. Lett., vol. 77, pp. 42654267, 2000. [41] S. Ghosh, P. Bhattacharya, E. Stoner, H. Jiang, J. Singh, S. Nuttinck, and J. Laskar, Temperature-dependent measurement of Auger recombination in self-organized In Ga As/GaAs quantum dots, Appl. Phys. Lett., vol. 79, pp. 722724, 2001. [42] L. Davis, Y. Lam, D. Nichols, J. Singh, and P. Bhattacharya, Auger recombination rates in compressively strained In Ga As/InGaAsP/InP (0:53 x 0:73) multiquantum well lasers, IEEE Photon. Technol. Lett., vol. 5, pp. 120122, 1993. [43] C. Walther, J. Bollmann, H. Kissel, H. Kirmse, W. Nuemann, and W. T. Masselink, Characterization of electron trap states due to InAs quantum dots in GaAs, Appl. Phys. Lett., vol. 76, pp. 29162918, 2000. [44] M. Grundmann, N. N. Ledenstov, R. Heitz, L. Eckey, J. Christen, J. Bohrer, D. Bimberg, S. S. Ruvimov, P. Werner, U. Richter, J. Heydenreich, V. M. Ustinov, A. Y. Egorov, A. E. Zhukov, P. S. Kopev, and Z. I. Alferov, InAs/GaAs quantum dots radiative recombination from zero-dimensional states, Physica Status Solidi B, vol. 188, pp. 249258, 1995. [45] M. Paillard, X. Marie, E. Vanelle, T. Amand, V. K. Kalevich, A. R. Kovsh, A. E. Zhukov, and V. M. Ustinov, Time-resolved photoluminescence in self-assembled InAs/GaAs quantum dots under strictly resonant excitation, Appl. Phys. Lett., vol. 76, pp. 7678, 2000.

 

Pallab Bhattacharya (M78SM83F89) received the Ph.D. degree from the University of Sheffield, Sheffield, U.K., in 1978. He is currently the James R. Mellor Professor of Engineering in the Department of Electrical and Engineering and Computer Science at the University of Michigan, Ann Arbor. His teaching and research interests include liquid-phase and molecular beam epitaxy of elemental and III-V compound semiconductors, materials characterization, electronic and optoelectronic devices, and optoelectronic integrated circuits. He was on the faculty of Oregon State University, Corvallis, from 1978 to 1983, and since 1983, he has been with the University of Michigan. He was an Invited Professor at the Ecole Polytechnique Federale de Lausanne, Switzerland, from 1981 to 1982. He has published over 300 technical articles in archival journals. He is author of the textbook Semiconductor Optoelectronic Devices (Englewood Cliffs, NJ: Prentice-Hall, 1994, 1st ed. and 1997, 2nd ed.) and was the editor for Properties of Lattice-Matched and Strained InGaAs (U.K.: INSPEC, 1993) and Properties of III-V Quantum Wells and Superlattices (U.K.: INSPEC, 1996). Dr. Bhattacharya is an Editor of the IEEE TRANSACTIONS ON ELECTRON DEVICES. He has served on the Advisory Board of the Electrical and Communications Systems Division at the National Science Foundation. He has also served on several other committees and panels in academia, government, industry, and technical conferences. He has received the Parker Rhodes Scholarship from the University of Sheffield, the Research Excellence Award from the University of Michigan, the Alexander von Humboldt Award, the SPIE Technology Achievement Award, the John Simon Guggenheim Award, the IEEE (LEOS) Distinguished Lecturer Award, and the S. S. Attwood Award from the University of Michigan. He is a member of the American Physical Society and a fellow of the Optical Society of America. Dr. Bhattacharya is the Editor of the IEEE Transactions on Electron Devices and has edited Properties of Lattice-Matched and Strained InGaAs (UK: INSPEC, 1993) and Properties of III-V Quantum Wells and Superlattices (UK: INSPEC, 1996). He has served on the Advisory Board of the Electrical and Communications Systems Division at the National Science Foundation. He has also served on several other committees and panels in academia, government, industry, and technical conferences. He has received the Research Excellence Award from the University of Michigan, the Alexander von Humboldt Award, the John Simon Guggenheim Fellowship, the IEEE (LEOS) Distinguished Lecturer Award, the S. S. Attwood Award from the University of Michigan, the lEEE (EDS) Paul Rappaport Award, the SPIE Technical Achievement Award, the Distinguished Faculty Achievement Award for the University of Michigan, the IEEE (LEOS) Engineering Achievement Award, and the Optical Society of America Nick Holonyak, Jr. Award. He is a Fellow of the IEEE and OSA.

Siddhartha Ghosh (S97) received the Ph.D. degree in electrical engineering from the University of Michigan, Ann Arbor, in January 2003. His dissertation focused on application of self-organized quantum dots to high-speed lasers and spin polarized light sources. He is currently a post-doctoral Research Fellow in the solid-state electronics laboratory at the University of Michigan. His research interests include molecular beam epitaxy (MBE) of III-V compound semiconductor optoelectronic devices, their fabrication, and high-speed characterization. He has also worked on MBE growth of III-Mn-V diluted magnetic semiconductors and their applications to spin-polarized LEDs. Dr. Ghosh has received Best Student Paper Awards at the 19th North American Molecular Beam Epitaxy (NA-MBE) Conference, Tempe, AZ, and at the 20th NA-MBE, Providence, RI.

962

IEEE JOURNAL OF QUANTUM ELECTRONICS, VOL. 39, NO. 8, AUGUST 2003

Sameer Pradhan received the B.E. (Hons.) degree in electronic engineering from the University of Bombay, Bombay, India, in 1998, and the M.S. degree in electrical engineering in 2001 from the University of Michigan, Ann Arbor, where he is currently working toward the Ph.D. degree in electrical engineering. His research interests include design, fabrication, and characterization of high-speed optoelectronic integrated circuits and devices such as photoreceivers, phototransceivers, and lasers for optical communication.

J. Urayama, photograph and biography not available at the time of publication.

Jasprit Singh (M88) received the Ph.D. degree in solid-state physics from the University of Chicago, Chicago, IL, in 1980, where he worked in the area of disordered semiconductors. He then spent three years at the University of Southern California, Los Angeles, two years at Wright Patterson AFB, OH, and has been with the University of Michigan, Ann Arbor, since 1985. His area of research is in physics and the design of semiconductor heterostructure devices. He has worked on electronic and optoelectronic devices based on traditional III-V materials, HgCdTe, SiGe, and the nitrides. Recently, his work has focused on exploiting ferroelectric materials within conventional semiconductor devices. He has written eight textbooks, including the most recent Semiconductor Devices: Basic Principles (New York: Wiley, 2001) and Electronic and Optoelectronic Properties of Semiconductor Structures (Cambridge, U.K.: Cambridge Univ. Press, 2003).

Kyoungsik Kim received the B.S. and M.S. degree in physics from Seoul National University, Seoul, Korea, in 1992 and 1994, respectively. He was a full-time Lecturer of Physics at the Korean Air Force Academy, Cheongwon, Korea, for military service from 1994 to 1997. He is currently a Research Assistant at the Center for Ultrafast Optical Science, The University of Michigan, Ann Arbor.

Zong-Kwei Wu was born in Taiwan, R.O.C., in 1975. He received the B.S. degree in electrical engineering from National Taiwan University, Taipei, Taiwan, R.O.C., in 1998. He is currently working toward the Ph.D. degree in electrical engineering at the University of Michigan, Ann Arbor. His research includes ultrafast carrier dynamics in quantum-dot systems.

Theodore B. Norris (M95) received the B.A. degree in physics (with highest honors) from Oberlin College, Oberlin, OH, in 1982, and the Ph.D. degree in physics from the University of Rochester, Rochester, NY, in 1989. His postdoctoral research was perfomed at Thomson-CSF, France, during 1989-1990. He is a Professor in the Electrical Engineering and Computer Science Department and the Center for Ultrafast Optical Science, University of Michigan, Ann Arbor. His research interests include application of femtosecond optical techniques to the physics of semiconductor nanostructures, in developing new ultrafast optical and optoelectronic measurement techniques, including applications to nanostructures and to biological imaging. Dr. Norris is a member of the American Physical Society, and is a Fellow of the Optical Society of America.

You might also like