Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Emission properties and photon statistics of a single quantum dot laser

Open Access Open Access

Abstract

A theoretical description for a single quantum-dot emitter in a microcavity is developed. We analyze for increasing steady-state pump rate the transition from the strong-coupling regime with photon antibunching to the weak-coupling regime with coherent emission. It is demonstrated how Coulomb interaction of excited carriers and excitation-induced dephasing can strongly modify the emission properties. Our theoretical investigations are based on a direct solution of the Liouville-von Neumann equation for the coupled carrier-photon system. We include multiple carrier excitations in the quantum dot, their Coulomb interaction, as well as excitation-induced dephasing and screening. Similarities and differences to atomic systems are discussed and results in the regime of recent experiments are interpreted.

©2010 Optical Society of America

Recent advances in the active field of cavity quantum electrodynamics have opened the doors to a fascinating regime of physics, where a single emitter strongly interacts with the electromagnetic field of a resonator. Experimental achievements in this field date back to the micro-maser with individual atoms traversing a microwave cavity [1] and its optical analog [2], and the emission of a train of individual photons from a single atom [3]. Single-atom laser emission is connected with the ultimate quantum limit where one photon is sufficient to saturate the atomic transition and novel properties unknown from conventional lasers are seen, which include nonclassical and more orderly light emission [4].

Single quantum dot (QD) lasers have been proposed as the solid-state analog to the single-atom laser [5]. In the past decade, considerable progress has been made in the fabrication of semiconductor microcavities with quasi-three-dimensional photon confinement and single QD emitters. These systems have demonstrated single-photon generation on demand [6, 7] and strong coupling [8, 9]. The laser emission has been studied with a small number of QD emitters [10, 11, 12, 13] and the goal of current investigations is the realization of a single-QD laser. In recent experiments [14, 15, 16] the dominance of a single QD in the laser emission was demonstrated.

The discrete level structure of QDs and the resulting similarity to atomic systems has been widely used throughout the literature by invoking atomic models to describe QD-based systems. In the limit of a single-emitter laser, the question what differences to expect between atomic and QD systems becomes even more prominent than for the many-emitter case where inhomogeneous broadening due to QD size and composition fluctuations masks experimental observations. Present laser experiments with dominating contributions from single QDs [14, 15, 16] are considered to show stimulated emission in the weak-coupling regime. In this paper, we analyze these experimental situations and predict results for further cavity quality improvements. A central point of our investigations are conceptual differences in the physics of atomic and QD emitters.

In an atomic system one usually assumes optical transition between two electronic configurations (the “laser levels”) that are resonantly coupled to a high-quality cavity mode. In the corresponding theoretical models, this coupling is described by the Jaynes-Cummings Hamiltonian. Additional configurations are used for the pump process. The theoretical description of the single-atom laser was introduced in [17]. On the other hand, a microscopic description of semiconductor systems usually starts from the electronic single-particle states. The occupation of these states with electrons and holes in the conduction and valence band states, respectively, leads even for a single QD to various system configurations with energies depending on the Coulomb interaction of the excited carriers. This is very much like in a multi-electron atom and it is what has earned QDs the name of artificial atoms.

Yet, the analogy breaks down in several important respects, related to the strength of the configuration (Coulomb) interaction and its consequences. In atoms the interaction induces a large energetic separation between configurations, allowing in atomic laser models to consider only a limited number of configurations and neglect those which lie outside the energetic window of interest. For instance, just two configurations are supposed to interact with the cavity mode. We argue below that in QDs such a restriction is usually not possible. The involvement of multiple configurations in the same interaction process changes the physics of the QD laser and requires a more flexible model.

To begin with, QDs are much larger than atoms, which leads to proportionally reduced Coulomb interaction of carriers in various states. This considerably weakens the strength of the configuration interaction. Secondly, in an atom a change in the configuration when an electron changes its orbital amounts to an important modification in the charge distribution, and the configuration energy is substantially shifted by the direct (Hartree) Coulomb interaction. On the contrary, in QDs, to a good accuracy [18] the conduction and valence band states have pair-wise equal wave functions and a carrier jump between such states does not modify the charge distribution in the dot.

As a typical example we consider self-assembled QDs that currently have many technological applications. The confinement potential gives rise to localized states, both for electrons and holes, below a quasi-continuum of wetting layer (WL) and bulk states [18]. For a QD laser, the lowest confined QD states of the electrons and holes provide the laser levels when tuned to the cavity mode. Additional levels are necessary for the pump process. Here one has to distinguish between different situations. In the limit of a shallow QD with only one confined state for electrons or holes, the quasi-continuum of delocalized WL states at higher energies would need to be used for the pump process. Efficient carrier relaxation between the WL and QD states due to carrier-carrier and carrier-phonon interaction easily facilitates the laser action [19]. Alternatively, deeper confinement potentials provide additional localized excited QD states in each band, which can be used directly for an optical pump process or as intermediate levels for carrier relaxation from the WL to the laser levels. Various system configurations arise from different occupations of these QD single-particle states with carriers1.

In the following we investigate the simplest possible QD system suitable for laser action under optical pumping2. Two confined QD states for holes (∣1〉, ∣2〉) and electrons (∣3〉, ∣4〉) are considered, as shown in Fig. 1. The pump process and the carrier relaxation are indicated by blue arrows (solid lines). Carrier transitions due to photon emission into and reabsorption from the laser mode are represented by the red arrow (dashed line) and optical processes coupled to non-lasing modes are shown in magenta (dotted lines). (Note that the conduction-valence-band picture and not the electron-hole picture is used.) In the ground state system configuration, the valence-band states are occupied with electrons. All states are spin-degenerate, however, optical transitions between carriers from the respective spin-subsystems couple to different light polarizations. The two spin-subsystems can be treated separately provided that the Coulomb interaction between configurations with opposite spin carriers is smaller than the line broadening. Also, to keep the discussion and the model as simple as possible, we do not consider the option of an additional degeneracy of the excited states.

 figure: Fig. 1.

Fig. 1. Localized single-particle states ∣1〉 - ∣4〉 in the QD confinement potential and carrier transitions for an optically pumped QD laser. The pump process P excites holes and electrons in the states ∣1〉 and ∣4〉 while carrier relaxation with rates γ 34 and γ 12 couples the pump levels to the laser levels; γ 23, γ 14 are the laser- and pump-level spontaneous emission rate into non-lasing modes.

Download Full Size | PDF

For one spin-subsystem, with two carriers in the ground-state we obtain the 6 possible system configurations of Fig. 2 which are connected by the transition processes depicted in Fig. 1. As we outline subsequently, two distinct configurations allow for the emission of a photon into the laser mode. Fig. 3a shows the laser transition in the presence of a valence electron in the state ∣1〉 or, in other words, the recombination of an s-exciton 1XS → 0X (the exciton laser transition). Similarly the transition with the second electron in ∣4〉 shown in Fig. 3b amounts to a transition from an sp-biexciton to a p-exciton 2Xsp → 1Xp (the biexciton laser transition). Since the pair of s-states have similar wave functions in good approximation, the s-exciton is not only globally but also locally neutral. The same is true for the p-exciton and therefore there is no direct Coulomb interaction between them. This shows that the Hartree interaction is not lifting the degeneracy between the exciton and biexciton transitions in Fig. 3 and, in the absence of other effects, they would both be resonant with the cavity mode.

 figure: Fig. 2.

Fig. 2. System configurations corresponding to two electrons in the four confined QD states of Fig. 1. The top row shows the state occupation for electrons in the conduction and valence-band picture while the bottom row corresponds to electrons (filled circles) and holes (open circles) in the electron-hole picture. The ground states (0X) is given by two valence electrons in the states ∣1〉, ∣2〉. Single excitons with electrons and holes in the lowest QD state (s-state) and excited QD state (p-state) are represented by 1XS and 1Xp, respectively. The sp-biexciton is labeled as 2Xsp. Selection rules typically identify 0XS and 0XP as optically dark configurations.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Transitions coupled to the laser mode: (a) exciton and (b) biexciton recombination.

Download Full Size | PDF

On the other hand, for same spin carriers, one has to take into account the exchange Coulomb interaction as well. Indeed, the transition of Fig. 3b involves the loss of the s-p exchange energy in both bands, while the one in Fig. 3a does not. This would take in principle the 2Xsp → 1Xp transition out of resonance with the laser mode (tuned to 1XS → 0X). In practice though the resonance condition is not sharp, but is a result of the interplay between the size of the detuning due to exchange (typically a few meVs [20, 21]) and the finite linewidths of the levels involved. Recent investigations showed that detuned QD transitions can still couple to high-quality microcavity modes [22].

Regarding the exchange Coulomb interaction of QD carriers as well as the finite linewidth of optical transitions, an important role is played by excited carriers in the WL. When in sufficient number, they efficiently screen the Coulomb interaction and thus reduce the strength of the exchange. Also, scattering processes involving these WL carriers (in addition to the carrier-phonon interaction) lead to finite lifetimes for the QD states. The WL screening and dephasing effects require the presence of an energetically close continuum of states, into which a population of carriers can be created, e.g. by direct WL pumping. This is not the case in atomic systems, but is quite common in QDs. One encounters here a third major difference between the two types of emitters, QDs and atoms.

With these facts in mind one should not exclude the biexcitonic configuration from the model, leaving the QD properties and the excitation condition to decide whether only one transition of Fig. 3 or, due to the excitation-induced screening and dephasing, both of them are efficiently coupled to the cavity mode. But this is not only the case for the Jaynes-Cummings coupling to the photon subsystem, each of the other processes considered (pumping, relaxation) may act upon several pairs of configurations. For instance the pumping gives rise to both the 0X → 1Xp and the 1Xs → 2Xsp transition. In a recent publication [23] coupling of several configurations by the same processes was also discussed. In the single-QD laser model which we define below, such possibilities are naturally taken into account.

We consider the four-level, single spin QD system described above. The formalism can be generalized in a straightforward manner to include both spins, p-level degeneracy, higher QD states a.s.o. The fermionic creation (annihilation) operators corresponding to the four states are denoted ai (ai), i = 1…4. The cavity mode, assumed resonant with the 1Xs → 0X transition, is described by the bosonic operators b and b. The equation of motion for the QD-photon density matrix ρ(t) in the interaction picture reads (h̄ = 1)

Tρ=i[HJC+HCoul,ρ]
+(i,j)γij2(2aiajρajaiajaiaiajρρajaiaiaj)
+κ2(2bbρbb).

The first line in Eq. (1) is the von Neumann evolution generator corresponding to the Jaynes-Cummings and the Coulomb interaction Hamiltonian. The former is given by

HJC=g[ba2a3+ba3a2],

where g is the coupling constant between the laser mode and the s-levels. The Coulomb Hamiltonian for this QD system [21] contains direct and exchange terms. For reasons discussed above only the exchange part

HCoulX=EspX[n3n4+(1n1)(1n2)]

contributes to the dynamics of the density matrix (see Appendix I), and gives rise to a detuning of 2EXsp in the biexciton transition. EXsp is the sp-exchange energy and ni = ai ai are the occupation number operators.

The second line in Eq. (1) describes dissipation processes derived from the interaction with different reservoirs, treated in the Born-Markov limit [24]. The sum runs over the following pairs of indices (i,j) = (1,2), (3,4), (2,3), (1,4) and the coefficients γij are each the relaxation rates for the ∣j〉 → ∣i〉 process. Additionally, in the sum we include the pumping term as an “inverse relaxation” ∣1〉 → ∣4〉, with the rate γ 41 = P. This simulates the situation where carriers are initially generated by optical excitation in the WL, and then are captured in the QD. With Coulomb scattering processes as the main capture mechanism, the pumping rate P is a function of the incoherent WL population. The discussion about the relationship between pump rate, WL population and resulting excitation-induced dephasing are found in Sections 2 and 3 of the appendix. Finally, the third line of Eq. (1) describes cavity losses at a rate κ in a similar way.

The Lindblad terms in Eq. (1) look similar to those used in atomic models [17, 25, 26, 27], with the important difference that here they are formulated using the fermionic operators ai aj, instead of the operators σij = ∣i〉〈j∣ acting on configurations. This leads to a different structure of the density-matrix evolution operator, as seen by comparing σjiσij = σjj of the atomic models with aj aiai = aj aj - aj ai ai aj , which is aj aj only in the case of a system with a single electron (ai aj = 0). With more than one electron, the difference amounts to the fact, described above, that a given physical process (relaxation, Jaynes-Cummings interaction) is involved in transitions between several pairs of configurations. In this sense the atomic models are formally equivalent to the present one with only one carrier [17], with the caveat that the levels considered there are configurations and not one-particle states.

To illustrate the implications of the discussed interplay of multiple configurations in single-QD lasers, the Liouville-von Neumann equation for the discussed system is solved numerically without further approximations. We present the input/output characteristics along with the photon autocorrelation function for different physical situations by considering typical QD parameters. To include the influence of a WL continuum, we account for the excitation-dependent broadening of the transition lines and screening of the Coulomb interaction, which are determined as functions of the pump rate as discussed in Appendix III. Additionally, we also consider the case without the impact of the WL. This more closely resembles the situation in an atomic system, where the energetic positions of the transitions are not excitation-density dependent. In the following examples, the equation of motion for the density matrix is solved numerically in time until a steady state is reached.

By tracing over the electronic degrees of freedom, our calculations provide the density matrix for the laser mode that contains the information about the statistical properties of the emitted photons. In experiments, frequently the photon autocorrelation function at zero delay time g (2)(0) = (〈n 2〉-〈n〉)/〈n2 with the photon number operator n = b b is used to characterize statistical properties. Single-photon sources are associated with anti-bunching behavior and g (2)(0) < 1, typical for a sub-poissonian photon statistics, indicates that immediately after a first photon emission, the likelihood for a second one is reduced. This is expected to be a common feature for single emitters, which need a certain re-excitation time after an emission event. Thermal emission with bunching behavior is characterized by g (2)(0) = 2, while a Poisson distribution of photons corresponds to g (2)(0) = 1.

In Fig. 4a,b the steady-state mean photon number 〈n〉 and the second-order photon correlation function g (2)(0) are shown versus pump rate P for parameters similar to the experiments of [16]. For P > 5 × 10-2/ps our results are reasonably close to the experimental findings. The s-shaped photon output curve in this pumping range resembles the behavior of solutions of the conventional rate equation for the multi-emitter laser. The characteristic kink appearing at the onset of the stimulated emission is related in these solutions to the value of the spontaneous emission coupling factor [28]. This similarity is surprising in the present context because, for single-emitter lasers, one expects saturation effects above the threshold region. The simplest argument is that for a fully inverted laser, a further increased pump rate does not enlarge the photon production. Moreover, in [17, 5] a quenching of the laser output has been discussed due to pump-induced dephasing in a situation where the pump process directly involves one of the laser levels.

Here the appearance of a kink has a different interpretation. In highly excited QDs, one expects excitation-dependent scattering rates into the laser levels. This effect introduces excitation-dependent dephasing which on the one hand lowers the rates for optical transitions and on the other hand relaxes the resonance condition for these transitions by line-broadening. In our model, these features are included via excitation-dependent rates γ 12 and γ 34. Additionally, in our multi-configuration model, with increasing pump rate first the probability for the 1XS configuration increases, but at elevated pump rates its probability diminishes at the expense of the configuration 2Xsp. This can be seen from the contributions of the exciton and biexciton transitions to the total photon production, which can be separated as described in Appendix IV and are shown in Fig. 4a. A kink in the input/output curve is obtained when the biexciton transition takes over.

 figure: Fig. 4.

Fig. 4. (a) Steady-state mean photon number 〈n〉 vs. pump rate P (solid line) together with its contributions from the exciton (dashed line) and biexciton (dotted line) transitions according to Fig. 3. (b) Second-order photon correlation function g (2)(0) versus pump rate P. (c) Excitation-induced dephasing (dashed line) and reduction of the Coulomb exchange energy EXsp due to WL carrier screening (solid line). (d) Photon emission spectrum for selected pump rates. Lines are vertically displaced. For all curves the light-matter coupling constant g = 0.9 ps-1, cavity losses at rate κ = 0.1 ps-1 and spontaneous emission rates γ 23 = γ 14 = 0.01 ps-1 are used.

Download Full Size | PDF

For weak pump rates P < 10-1/ps, when the excitonic recombination dominates the emission, we obtain the expected anti-bunching behavior g (2)(0) < 1 in Fig. 4b. In the transition region between excitonic and biexcitonic emission, a clear photon-bunching behavior with 1 < g (2) (0) < 2 is observed. This might seem surprising for a single-emitter laser, and was tentatively attributed in the literature to background contributions [16]. However, within the present model, with the increasing contribution of the biexcitonic configuration, shown by the dotted line in Fig. 4a, successive photon-pair emission becomes possible without a repumping of the system. Note that in our system the subsequent emission by the sp-biexciton and the s-exciton requires a carrier relaxation from p- to s-shells. Even though the relaxation times are not constant, but increase with the excitation density, see Fig. 4a, they are short compared with the cavity lifetime of 10ps, enabling storage of the emitted photons and the development of cavity photon bunching. The contribution of the biexcitonic transition to photon emission into the laser mode is facilitated for larger pump rates by the increased dephasing of the optical transitions that broadens the emission lines, and the reduction of the exchange splitting of the exciton and biexciton lines, as depicted in Fig. 4c. Also our calculations exhibit a transition from the strong to weak-coupling regime. For a weak pump rate P = 10-2/ps the emission spectrum in Fig. 4d shows two main peaks due to Rabi-splitting of the degenerate cavity and exciton line, weak side peaks due to the next step of the Jaynes-Cummings ladder and additional peaks from the sp-biexciton transitions. The latter ones are positioned at twice the Coulomb exchange energy EXsp Excitation-induced dephasing leads to a transition to the weak-coupling regime. We note that for pump rates P > 5 × 10-2/ps, where our input-output curve shows similar behavior as the results in [16], the strong coupling regime gradually fades out, in an experiment of finite resolution possibly even earlier.

 figure: Fig. 5.

Fig. 5. (a) Steady-state mean photon number 〈n〉 vs. pump rate P. From bottom to top, the cavity losses are reduced using κ = 0.1 ps-1 , 0.05 ps-1, 0.025 ps-1, 0.0125 ps-1. For a cavity mode wavelength of 915 nm this amounts to a cavity quality Q ~ 21 × 103,42 × 103, 84 × 103, 168 ×103. Other parameters are the same as in Fig. 4. (b) Corresponding results for the second-order photon correlation function g (2)(0) in the stationary regime vs. pump rate P.

Download Full Size | PDF

In the following, we discuss the impact of modified system parameters and properties in order to obtain a more general picture. Starting from the above example, the influence of the cavity quality and the WL-influence is analyzed. Additionally, we compare our QD results to the limiting case of an “atomic model”.

Fig. 5 reveals, that for reduced cavity losses (increasing cavity-quality factor Q = ω/κ defined with the cavity mode frequency ω) the mean photon number increases and the kink in the transition region is smoother. More remarkably, anti-bunching (for small P) and bunching (for intermediate P) are reduced. The steady state photon statistics depends not only on the QD emission rate, but also on the photon storage properties of the cavity. Therefore, if antibunching is desired, a compromise with respect to the cavity quality has to be made. For high pump rates, in all cases the lasing-regime with stimulated emission is reached, as reflected in g (2)(0) = 1.

A different situation is found when the excitation-induced effects of the WL carriers are not considered. This would require a very large energy separation of the QD states from delocalized states, or other QD systems like semiconductor microcrystalites. We describe this situation by using pump-independent values for the scattering/dephasing rates γ 12 and γ 34 and for the exciton-biexciton Coulomb exchange splitting 2EXsp. The corresponding red lines in Fig. 6 exhibit a strong quenching of the mean photon number. In contrast to the systems of Refs. [17, 5], this is not induced by dephasing due to the pump process (as a result of the pump process acting directly on one of the system configurations contributing to the laser process) but rather by bringing the system into the biexcitonic configuration that inefficiently couples to the cavity mode (when the detuning is too large and line broadening is weak). From the red to the blue lines in Fig. 6 the Coulomb exchange energy EXsp is increased by a factor of two. This further reduces the biexciton contribution to the recombination and the quenching of the mean photon number is more pronounced. The decoupling of the QD from the WL induces a stronger antibunching, and the transition into the bunching-regime occurs at higher pump rates.

 figure: Fig. 6.

Fig. 6. (a) Steady-state mean photon number 〈n〉, (b) second-order photon correlation function g (2)(0), (c) exciton (solid line) and biexciton (dashed line) contribution to the mean photon number, versus pump rate P. The black lines correspond to the results of Fig. 4. For the red lines the Coulomb energy and the dephasing are kept constant: EXsp = 2.68 meV, γ 12 = γ 34 = 0.05 meV. In comparison to the red curves, EXsp = 5.36 meV is used for the blue lines. The green curves correspond to the “atomic model”, using σij instead of ai aj. Apart from the Coulomb interaction, which is assumed to be already included in the atomic con-figuration energies (therefore EXsp = 0), the parameters are the same as for the red curves.

Download Full Size | PDF

As the true atomic limit we refer to the situation where off-resonant transitions are completely removed from the model and the remaining ones (between a reduced number of configurations) are described by σij, operators. Replacing ai aj by σij, in our case leads to a four-state atomic model that was considered, e.g. in [17]. The results are shown in the green curve of Fig. 6. In the low-pumping regime, the mean photon number coincides with the QD-models. Differences occur with rising importance of the biexciton contribution (dashed lines in Fig. 6c), which is not considered in the atomic model. A striking difference to the QD-model is the pronounced antibunching behavior throughout the whole range of pump rates shown in Fig. 6b.

Returning to our starting point, we conclude that a single QD is truly a semiconductor system and in its nature subject to many-particle effects. A static configuration picture, which can be used in the description of atomic systems, is in general not justified for QDs in the presence of carriers in a WL. We find that the demonstrated transition from photon anti-bunching to bunching with increasing pump rate as well as the characteristic shape of the input/output curve depend on interaction-induced dephasing and screening effects. For parameters corresponding to present experiments the transition from strong to weak coupling is identified in a regime where spontaneous emission still dominates over stimulated emission and the average cavity photon number is below one.

Appendix

1. Treatment of the Coulomb interaction

With initial conditions involving only diagonal elements of the density matrix, it turns out that only a few off-diagonal elements appear in the evolution. These are created by the action of the Jaynes-Cummings Hamiltonian, and the pair of states involved differ only by the change of the s-state occupancies and by a corresponding change in the photon number, as depicted in Fig. 3. This circumstance simplifies considerably the action of the Coulomb Hamiltonian. With a limited number of QD states, and apart from single-particle energy renormalizations, the Hamiltonian describing the Coulomb interaction between QD configurations reduces to the direct and exchange terms HCoul = HDCoul + HXCoul. The exchange part is given in Eq. (2), and the direct part is [21]

HCoulD=12i,j=s,pEi,jD(nienih)(njenjh)

with EDi,j the direct (Hartree) energies and nei, nhi the electron and hole number operator respectively. This particular form is a consequence of having equal electron and hole wavefunction for both the s- and the p-levels, and therefore the Coulomb matrix elements EDi,j,EXi,j do not depend on the band index. For an equal number of electrons and holes nei = nhi,i = s, p the Hartree term does not contribute, and this is the expression of the local neutrality of the exciton.

It is seen that both the direct and the exchange Hamiltonian are diagonal in the occupation number representation. Therefore their contributions in the Liouville-von Neumann equation are proportional to the Coulomb energy difference between the initial and the final configuration. This can be non-zero only for off-diagonal elements. There is no difference introduced by the direct part, because the states involved have the same number of electrons and holes in both the s- and the p-states. The only such effect comes from the exchange, and it appears in the density matrix element between the 2Xsp biexciton and the Xp exciton. The corresponding value is given by 2EXsp.

2. Pumping method

The purpose of the pump process is to excite electrons and holes in the QD conduction- and valence-band states, respectively. This can be accomplished by resonant optical pumping of the excited QD states (∣1〉 and ∣4〉 in Fig. 1). Experimentally this is challenging, e.g. since the resonance condition is changing with excitation density. More practical in experiments is optical pumping of the delocalized states with higher transition energies. In the past, it has been demonstrated both experimentally and theoretically, that carrier scattering processes rapidly redistribute these carriers, and in particular ultrafast capture of electrons and holes from the delocalized states into the QD states is possible, see [19] and references therein.

By directly solving the Liouville-von Neumann equation (1) in terms of QD carrier configurations and photon number states we avoid ambiguities arising when approximate methods for the treatment of the carrier-photon interaction are applied to systems with few carrier configurations and photons. (For systems with many QDs, the vastly growing number of electronic configurations may justify the application of approximate methods developed in many-particle physics.) In turn, the direct solution of the Liouville-von Neumann equation is only feasible when carriers in delocalized states do not appear explicitly in the system configurations. This relies on the approximation that WL carriers – while influencing in the system dynamics – are not considered in the density matrix of the coupled carrier-photon system by their own dynamical variables. The large degrees of freedom associated with the delocalized states justifies this treatment.

Accordingly we assume that electrons are generated in the state ∣4〉 and removed from the state ∣1〉 of Fig. 1 via the coupling through Fermionic reservoires representing the WL carriers. The pump rate is described by the generation rate of electrons and holes in the states ∣4〉 and ∣1〉, respectively, using the obove discussed Lindblad terms added to the Liouville-von Neumann equation. Furthermore, we suppose that the pump mechanism generates electrons and holes at equal rates. One may argue that, in priciple the capture rates are different for electrons and holes. This can be realized via two separate Lindblad terms with different rates, describing the injection of electrons and holes in the manner of [5]. We have performed calculations also for this pump mechanism and found results very similar to those we have used in this paper.

With the capture mechanism given by Auger processes involving two carriers initially in the WL, we assume P = αN 2, where N is the WL population density. For the calculations in this paper we use α = 10-22cm4ps-1, which corresponds to a carrier density of 1012cm-2 at a pump rate of 100ps-1. In an experiment, this relationship depends on the spot diameter, wave length and alignment of the exciting laser.

3. Interaction-induced dephasing and screening

Various effects like population factors and screening of the Coulomb interaction directly influence the dependence of the dephasing due to carrier scattering processes on the excited carrier density in the WL N. It has been shown in Refs. [29, 30] for bulk semiconductors and QWs and in [31] for QD systems, that a simple fit formula γ 12 = γ 34 = γ 0 + (g 1 N)c provides in many cases a reasonable first estimate of the dephasing rate in terms of the carrier density N.

As another effect, screening the Coulomb interaction due to WL carriers leads directly to a weakening of the exchange interaction. Based on the same theory, we account for this effect via EXsp = a + b e-dN using appropriate parameters. Details of the treatment of these excitation-density dependent effects are given below.

The sum of both discussed effects, the line broadening and the screening of the exchange interaction, finally determines the overlap of the biexcitonic 2Xsp → 1Xp transition with the resonator mode (tuned to the exciton transition 1XS → 0X). The FWHM of the biexciton line is given by γ 12 + γ 34; for typical microresonators with quality factors in the range of 104, the mode linewidth is much smaller. With increasing WL excitation density, the line broadening and the Coulomb screening act in a way that partly compensates the spectral shift introduced by the exchange interaction. Depending on the system parameter, the biexciton line remains (partly) in resonance with the cavity mode. As mentioned before, this is different to the situation in atomic systems, where large energy shifts tune different configurations far away from the resonator mode.

From a quantum-kinetic theory for Coulomb scattering that provides the dominant contributions at elevated carrier density N, the dephasing was determined and values of c = 0.63 and γ 1 = 2.77 × 10-12 (for N in cm-2 and γ 12,γ 34 in meV) have been extracted. While the calculations in Ref. [31] were performed for a temperature of 300K, we repeated this analysis in the low temperature regime. At low carrier densities, the dephasing is strongly temperature dependent due to contributions from longitudinal accustic and optical phonons. Their influence, that only weakly depends on carrier density, is summarized in γ 0, for which we use a typical value of 0.05meV [32].

To account for the screening of the Coulomb interaction in the presence of WL carriers, we use the parameters a=0.88meV, b=1.8meV and d = 0.5 × 10-11 cm2. These are obtained from a fit to calulations along the lines of [33] for the considered low-temperature conditions.

4. Discussion of output quenching mechanisms

In the steady state, the mean photon number is directly proportional to the real part of the photon-assisted polarization (b a 2 a 3〉, which contains two terms

ba2a3=ba2a3a1a1+ba2a3a4a4,

corresponding to the electron not taking part in the transition being either in the state ∣1〉 (exciton transition shown in Fig. 3a) or in ∣4〉 (biexciton transition shown in Fig. 3b), respectively. Both terms are shown in Fig. 4a and by the two sets of curves in Fig. 6c. The first term is dephased by the pumping, because its configuration allows for the pumping transition (Fig. 3a). Thus, it decreases with increasing pump rate P. The other term is not dephased and reaches saturation. The second effect is that the pump does not only create the 1Xp exciton state, but also depletes the photon-producing 1Xs state in favour of the biexciton 2Xsp, which is off-resonant. This is a new quenching mechanism, relying on the fact that the pump process acts simultaneously between different configurations, which, in general, are not both resonant with the cavity mode. The magnitude of both effects is seen in Fig. 6c.

Acknowledgments

We thank Michael Lorke for discussions and for calculating the dephasing and screening dependence used in the appendix. This work has been financially supported through the Deutsche Forschungsgemeinschaft and with a grant for CPU time at the Forschungszentrum Jülich

Footnotes

1Note that in the description of atomic emitters the system configurations are considered, see e.g. Fig. 1 in [17], and not the single particle states as in Fig. 1 of this paper.
2In [5] a two-level QD system was investigated, where pumping was facilitated by direct coupling of the QD laser transitions to electron and hole reservoirs via a tunneling process. This inevitably leads to quenching of the laser output for increasing pump rates as discussed in the cited paper. The used four configurations of Ref [5] are not suitable to describe optical excitation.

References and links

1. D. Meschede, H. Walter, and G. Müller, “One-atom maser,” Phys. Rev. Lett. 54, 551–554 (1985). [CrossRef]   [PubMed]  

2. K. An, J. J. Childs, R. R. Dasari, and M. S. Feld, “Microlaser: A laser with one atome in an optical resonator,” Phys. Rev. Lett. 73, 3375–3378 (1994). [CrossRef]   [PubMed]  

3. J. McKeever, A. Boca, A. D. Boozer, J. R. Buck, and H. J. Kimble, “Experimental realization of a one-atom laser in the regime of strong coupling,” Nature 425, 268–271 (2003). [CrossRef]   [PubMed]  

4. H. Carmichael and L. A. Orozco, “Quantum optics: Single atom lases orderly light,” Nature 425, 246–247 (2003). [CrossRef]   [PubMed]  

5. O. Benson and Y. Yamamoto, “Master-equation model of a single-quantum-dot microsphere laser,” Phys. Rev. A 59, 4756–4763 (1999). [CrossRef]  

6. P. Michler, A. Kiraz, C. Becher, W. V. Schoenfeld, P. M. Petroff, L. Zhang, E. Hu, and A. Imamoglu, “A quantum dot single-photon turnstile device,” Science , 290, 2282–2285 (2000). [CrossRef]   [PubMed]  

7. M. Pelton, C. Santori, J. Vuckovic, B. Zhang, G. S. Solomon, J. Plant, and Y. Yamamoto, “Efficient source of single photons: A single quantum dot in a micropost microcavity,” Phys. Rev. Lett. 89, 233602 (2002). [CrossRef]   [PubMed]  

8. J. P. Reithmaier, G. Sek, A. Löffler, C. Hofmann, S. Kuhn, S. Reitzenstein, L. V. Keldysh, V. D. Kulakovskii, T. L. Reinecke, and A. Forchel, “Strong coupling in a single quantum dot - semiconductor microcavity system,” Nature 432, 197–200 (2004). [CrossRef]   [PubMed]  

9. T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper, C. Ell, O. B. Shchekin, and D. G. Deppe, “Vacuum rabi splitting with a single quantum dot in a photonic crystal nanocavity,” Nature 432, 200–203 (2004). [CrossRef]   [PubMed]  

10. J. Hendrickson, B. C. Richards, J. Sweet, S. Mosor, C. Christenson, D. Lam, G. Khitrova, H. M. Gibbs, T. Yoshie, A. Scherer, O. B. Shchekin, and D. G. Deppe, “Quantum dot photonic-crystal-slab nanocavities: Quality factors and lasing,” Phys. Rev. B 72, 193303 (2005). [CrossRef]  

11. S. Reitzenstein, A. Bazhenov, A. Gorbunov, C. Hofmann, S. Münch, A. Löffler, M. Kamp, J. P. Reithmaier, V. D. Kulakovskii, and A. Forchel, “Lasing in high-q quantum-dot micropillar cavities,” Appl. Phys. Lett. 89, 051107 (2006). [CrossRef]  

12. S. Strauf, K. Hennessy, M. T. Rakher, Y.-S. Choi, A. Badolato, L. C. Andreani, E. L. Hu, P. M. Petroff, and D. Brouwmeester, “Self-tuned quantum dot gain in photonic crystal lasers,” Phys. Rev. Lett. 96, 127404 (2006). [CrossRef]   [PubMed]  

13. S. M. Ulrich, C. Gies, J. Wiersig, S. Reitzenstein, C. Hofmann, A. Löffler, A. Forchel, F. Jahnke, and P. Michler, “Photon statistics of semiconductor microcavity lasers,” Phys. Rev. Lett. 98, 043906 (2007). [CrossRef]   [PubMed]  

14. Z. G. Xie, S. Götzinger, W. Fang, H. Cao, and G. S. Solomon, “Influence of a single quantum dot state on the characteristics of a microdisk laser,” Phys. Rev. Lett. 98, 117401 (2007). [CrossRef]   [PubMed]  

15. S. Reitzenstein, C. Bockler, A. Bazhenov, A. Gorbunov, A. Loffler, M. Kamp, V. D. Kulakovskii, and A. Forchel, “Single quantum dot controlled lasing effects inhigh-q micropillar cavities,” Opt. Express 16, 4848–4857 (2008). [CrossRef]   [PubMed]  

16. M. Nomura, N. Kumagai, S. Iwamoto, Y. Ota, and Y. Arakawa, “Photonic crystal nanocavity laser with a single quantum dot gain,” Opt. Express 17, 15975–15982 (2009). [CrossRef]   [PubMed]  

17. Y. Mu and C. M. Savage, “One-atom lasers,” Phys. Rev. A 46, 5944–5954 (1992). [CrossRef]   [PubMed]  

18. A. Wojs, P. Hawrylak, S. Fafard, and L. Jacak, “Electronic structure and magneto-optics of self-assembled quantum dots,” Phys. Rev. B 54, 5604–5608 (1996). [CrossRef]  

19. H. Kurtze, J. Seebeck, P. Gartner, D. R. Yakovlev, D. Reuter, A. D. Wieck, M. Bayer, and F. Jahnke, “Carrier relaxation dynamics in self-assembled semiconductor quantum dots,” Phys. Rev. B 80, 235319 (2009). [CrossRef]  

20. P. Hawrylak, “Excitonic artificial atoms: Engineering optical properties of quantum dots,” Phys. Rev. B 60, 5597–5608 (1999). [CrossRef]  

21. N. Baer, P. Gartner, and F. Jahnke, “Coulomb effects in semiconductor quantum dots,” Eur. Phys. J. B 42, 231–237 (2004). [CrossRef]  

22. S. Ates, S. M. Ulrich, A. Ulhaq, S. Reitzenstein, A. Löffler, S. Höfling, A. Forchel, and P. Michler, “Non-resonant dotcavity coupling and its potential for resonant single-quantum-dot spectroscopy,” Nat. Photonics 3, 724–728 (2009). [CrossRef]  

23. E. Valle, S. Zippilli, F. P. Laussy, A. Gonzalez-Tudela, G. Morigi, and C. Tejedor, “Two-photon lasing by a single quantum dot in a high- q microcavity,” Phys. Rev. B 81, 035302 (2010). [CrossRef]  

24. H. J. Carmichael, Statistical Methods in Quantum Optics 1, (Springer, Berlin, 1998).

25. T. Pellizzari and H. Ritsch, “Photon statistics of the three-level one-atom laser,” J. Modern Opt. 41, 609–623 (1994). [CrossRef]  

26. A. D. Boozer, A. Boca, J. R. Buck, J. McKeever, and H. J. Kimble, “Comparison of theory and experiment for a one-atom laser in a regime of strong coupling,” Phys. Rev. A 70, 023814 (2004). [CrossRef]  

27. J. I. Perea, D. Porras, and C. Tejedor, “Dynamics of the excitations of a quantum dot in a microcavity,” Phys. Rev. B 70, 115304 (2004). [CrossRef]  

28. P. R. Rice and H. J. Carmichael, “Photon statistics of cavity-QED lasers,” Phys. Rev. A 50, 4318–4329 (1994). [CrossRef]   [PubMed]  

29. W. A. Hügel, M. F. Heinrich, and M. Wegener, “Dephasing due to carrier-carrier scattering in 2d,” Phys. Stat. Sol. (b) 221, 473–476 (2000). [CrossRef]  

30. H. Haug, “Coulomb quantum kinetics for semiconductor femtosecond spectroscopy,” Phys. Stat. Sol. (b) 221, 179–188 (2000). [CrossRef]  

31. M. Lorke, J. Seebeck, T. R. Nielsen, P. Gartner, and F. Jahnke, “Excitation dependence of the homogeneous linewidths in quantum dots,” Phys. Stat. Sol. (c) 3, 2393–2396 (2006). [CrossRef]  

32. M. Bayer and A. Forchel, “Temperature dependence of the exciton homogeneous linewidth in In0.6Ga0.4As/GaAs self-assembled quantum dots,” Phys. Rev. B 65, 041308(R) (2002). [CrossRef]  

33. M. Lorke, T. R. Nielsen, J. Seebeck, P. Gartner, and F. Jahnke, “Influence of carrier-carrier and electron-phonon correlations on optical absorption and gain in quantum-dot systems,” Phys. Rev. B 73, 085324 (2006). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. Localized single-particle states ∣1〉 - ∣4〉 in the QD confinement potential and carrier transitions for an optically pumped QD laser. The pump process P excites holes and electrons in the states ∣1〉 and ∣4〉 while carrier relaxation with rates γ 34 and γ 12 couples the pump levels to the laser levels; γ 23, γ 14 are the laser- and pump-level spontaneous emission rate into non-lasing modes.
Fig. 2.
Fig. 2. System configurations corresponding to two electrons in the four confined QD states of Fig. 1. The top row shows the state occupation for electrons in the conduction and valence-band picture while the bottom row corresponds to electrons (filled circles) and holes (open circles) in the electron-hole picture. The ground states (0X) is given by two valence electrons in the states ∣1〉, ∣2〉. Single excitons with electrons and holes in the lowest QD state (s-state) and excited QD state (p-state) are represented by 1XS and 1Xp , respectively. The sp-biexciton is labeled as 2Xsp . Selection rules typically identify 0XS and 0XP as optically dark configurations.
Fig. 3.
Fig. 3. Transitions coupled to the laser mode: (a) exciton and (b) biexciton recombination.
Fig. 4.
Fig. 4. (a) Steady-state mean photon number 〈n〉 vs. pump rate P (solid line) together with its contributions from the exciton (dashed line) and biexciton (dotted line) transitions according to Fig. 3. (b) Second-order photon correlation function g (2)(0) versus pump rate P. (c) Excitation-induced dephasing (dashed line) and reduction of the Coulomb exchange energy EX sp due to WL carrier screening (solid line). (d) Photon emission spectrum for selected pump rates. Lines are vertically displaced. For all curves the light-matter coupling constant g = 0.9 ps-1, cavity losses at rate κ = 0.1 ps-1 and spontaneous emission rates γ 23 = γ 14 = 0.01 ps-1 are used.
Fig. 5.
Fig. 5. (a) Steady-state mean photon number 〈n〉 vs. pump rate P. From bottom to top, the cavity losses are reduced using κ = 0.1 ps-1 , 0.05 ps-1, 0.025 ps-1, 0.0125 ps-1. For a cavity mode wavelength of 915 nm this amounts to a cavity quality Q ~ 21 × 103,42 × 103, 84 × 103, 168 ×103. Other parameters are the same as in Fig. 4. (b) Corresponding results for the second-order photon correlation function g (2)(0) in the stationary regime vs. pump rate P.
Fig. 6.
Fig. 6. (a) Steady-state mean photon number 〈n〉, (b) second-order photon correlation function g (2)(0), (c) exciton (solid line) and biexciton (dashed line) contribution to the mean photon number, versus pump rate P. The black lines correspond to the results of Fig. 4. For the red lines the Coulomb energy and the dephasing are kept constant: EX sp = 2.68 meV, γ 12 = γ 34 = 0.05 meV. In comparison to the red curves, EX sp = 5.36 meV is used for the blue lines. The green curves correspond to the “atomic model”, using σij instead of ai aj . Apart from the Coulomb interaction, which is assumed to be already included in the atomic con-figuration energies (therefore EX sp = 0), the parameters are the same as for the red curves.

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

Tρ=i[HJC+HCoul,ρ]
+(i,j)γij2 (2aiajρajaiajaiaiajρρajaiaiaj)
+κ2 (2bbρbb) .
HJC=g[ba2a3+ba3a2] ,
HCoulX=EspX [n3n4+(1n1)(1n2)]
HCoulD=12i,j=s,pEi,jD(nienih) (njenjh)
ba2a3=ba2a3a1a1 +ba2a3a4a4,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.