<<

arXiv:2008.11456v2 [physics.optics] 20 Oct 2020 ihlrevle ftesotnosemission spontaneous the lasers of nanoscale values simulate to large used with been have be- stochas- methods Recently, noise tic quantum [11–15]. important associated increasingly the comes up mode(s), cavity ending large rela- photons the a emitted in a and spontaneously With photons the intra-cavity of statistics: fraction of photon number and small noise tively photon the is research. prominent of a and currently micro- subject are biochem- Therefore, properties and their [7–10]. and chemical nanolasers detection like as and uses, [6], sensing promis- other interconnects ical many them optical for make as on-chip devices, well for the candidates of ing with footprint combined small features, laser high and These the for consumption useful [2–5]. power be speeds low for can modulation allows This it as [1]. devices, light mode(s) emitted cavity emission spontaneously the light of into coupling for the modes increase available and of number the reduce htnpoaiiydniyfntosa aigsystem lasing a intra-cavity as the functions consider density to it probability use photon numerous a we to with Here leading approach method, applications. new versatile the and with made avoided robust be assumptions can numerical [18] in certain or- [16– Additionally, several approaches previous by to 18]. times compared magnitude computation dynamics, of the laser ders the decrease describe can to used and biochem- be and can chemistry of [19–22], fields istry the in used cavity. often the proach, in the emitters with and associated photons noise the the includ- is, of by that discreteness of well noise; statistics shot surprisingly only and captured ing noise be the that can found nanolasers was it and [16–18], ∗ [email protected] n ao rao neeti ead onanolasers to regards in interest of area major One can scale nanometer and micro- the on cavities Optical nti okw ilso htaohrsohsi ap- stochastic another that show will we work this In ffiin tcatcsmlto frt qain n photon and equations rate of stochastic Efficient htaetpclyascae ihtepeec fsrn in strong of presence the analyti with the associated both typically that shown are is p that it intra-cavity Fi and examine exactly. equations, analytics to rate the altered used follow results is numerical FRM the The that numerical circumventing events. T also of redu thousandfold efficient. ordering while a more algorithms, to significantly earlier up gives to be (FRM), statis to Method and shown Reaction numbers is First photon algorithms steady-state the to of lead micr of methods behaviour Both steady-state and dynamics the simulating ae nart qainmdlfrsnl-oetolvllas two-level single-mode for model equation rate a on Based .INTRODUCTION I. 2 aohtn-Cne o aohtnc,TcnclUniversi Technical Nanophotonics, for Center - NanoPhoton 1 eateto htnc niern,TcnclUniversit Technical Engineering, Photonics of Department mlC Andr´e, C. Emil rtd ld 4A K20 g.Lnb,Denmark Lyngby, Kgs. DK-2800 Denmark 345A, Lyngby, Plads Ørsteds Kgs. DK-2800 345A, Plads Ørsteds 1 eprMørk, Jesper β -factor ,2 1, n aiymd,a hw ceaial nFig. in schematically shown as cavity-mode, ae.Fnly ewl umrz n ics u results our discuss and Section summarize in will emission we Finally, stimulated and asym- rates. Section effective spontaneous an the in introducing between and of metry effects laser the the examine will of new distributions the photon Using Section equations. in rate algorithm, laser the stochasti- simulating for cally algorithm then efficient and more [18] another, in laser introduce used the algorithm of simulation overview stochastic brief the Section in a In consider give we equations. and rate that paper, model the laser of most basic the introduce inter-will collective way of a effects as equations the correlations. rate emitter describe the qualitatively effectively in to introducing emission we of of and rates prospects altered emission, the coherent investigate to will thermal from transitions oewt h aeculn tegh oi sconve- is populations it level so total the strength, with coupling work cavity to same the nient to the couple with emitters all mode that simplicity for sume o h xie n rudlvl,rsetvl.Snewe have Since we respectively. emitters, levels, two-level ground consider and excited the for h xie tt.Tedpnec of dependence The state. excited the eshv aeo osit o-aigmdsgvnby given modes non-lasing into loss of rate a have ters a ea aeprpoo ie by given photon per rate decay a has elehneet ial,teeitr r upda a at of pumped emitter are Pur- per emitters includes rate the and Finally, [18], enhancement. in cell discussed is parameters device h aepreitro msinit h aigmd is mode lasing the into emission of denoted emitter per rate The aiymd,wl edntdby the denoted of be population will the mode, i.e. cavity number, photon intra-cavity 0 I AE OE N AEEQUATIONS RATE AND MODEL LASER II. ecnie oe o aoae ossigof consisting nanolaser a for model a consider We h ae ssrcue sflos nSection In follows: as structured is paper The itnttolvleitr neatn ihasingle a with interacting emitters two-level distinct n atj Wubs Martijn and al,teFMi ple oasto slightly of set a to applied is FRM the nally, susrgrigtm-nrmn ieand size time-increment regarding issues a n ueia eut xii features exhibit results numerical and cal e-mte orltosi nanolasers. in correlations ter-emitter -adnnlsr r ecie ndetail. in described are nanolasers and o- γ r r,toagrtm o stochastically for algorithms two ers, to ncmuaintm compared time computation in ction iscaatrsi flsr,btone but lasers, of characteristic tics VI iigattldcyrate decay total a giving , oo itiuin,adi sfound is it and distributions, hoton i loih,kona Gillespie’s as known algorithm, his . fDenmark, of y γ yo Denmark, of ty p IV . ,2, 1, ttsiso nanolasers of statistics ewl xmn h intra-cavity the examine will we ∗ III ewl eiwaddiscuss and review will we n p h aiymode cavity The . n γ γ e r γ t + c = nmtra and material on n h emit- the and , n γ g r = + 1 n eas- We . e γ n bg and 0 The . V II from γ we we bg n g . 2

lasing [11–14]. The steady-state photon statistics of a

 laser are typically measured by the first two statistical moments of the photon number, or the mean value hnpi 2 2 2 and the variance h∆npi = hnpi − hnpi . They are of- ten summarised using the zero-delay second-order pho- ton auto-correlation function g(2)(0) [26, 27]. This can be expressed through the relative intensity noise (RIN) [26] as FIG. 1. Sketch of the two-level laser model. 2 h∆npi (2) hnp(np − 1)i 1 Event type Average rate RIN = , g (0) = = 1+RIN− . aµ hn i2 hn i2 hn i Stimulated Em. a γ n (t)n (t) p p p st r e p (4) Spontaneous Em. a γ n (t) sp r e Lasing is then characterised by a transition from thermal Absorption a γ (n0 − n (t))n (t) a r e p (LED) light, for which g(2)(0) = 2, to coherent (laser) Cavity decay ac γcnp(t) light, for which g(2)(0) ≈ 1. Pumping ap γp(n0 − ne(t)) Background decay abg γbg ne(t)

TABLE I. Average rates of event types in the laser rate equa- III. STOCHASTIC SIMULATION tions. In [18] it was shown that stochastic based In this work we will focus on so-called Class B lasers, on the algorithm presented in [17] agreed very well with where the dephasing rate γ of the emitters is much larger analytical results obtained from a small-signal analysis of 2 the rate equations that is accurate even down to a few than the photon loss rate γc, so that the emitter polar- ization can be adiabatically eliminated from the analysis emitters (n0 ≥ 10). In this section we will start by re- [23]. For such devices, the laser dynamics may be de- viewing the specific method used for the stochastic laser scribed through rate equations for the photon number dynamics simulations in [18], and then we will introduce n and the number of excited emitters n [2, 18, 24, 25]: a different method to produce the same results with sig- p e nificantly increased efficiency. n˙ p = γr(2ne − n0)np + γrne − γcnp + Fp, (1)

n˙ e = γp(n − ne) − γr(2ne − n )np − γtne + Fe. (2) 0 0 A. Fixed Time-Increment Stochastic Algorithm Here Fp and Fe are conventional zero-mean, delta- correlated stochastic Langevin terms, described explicitly The basic idea behind the stochastic simulations is to in [18, 26]. They take into account the shot noise in the assume that the laser rate equations (1)-(2) describe a particle numbers due to the discrete nature of photons collection of discrete particles, namely photons and ex- and excited emitters. The laser rate equations (1)-(2) cited emitters, the numbers of which fluctuate due to give a simple and intuitively pleasing description of laser several processes: loss of particles (cavity or background dynamics, since each term can be easily associated with a losses), exchange of particles (emission or absorption) or process occurring in the laser, as summarized in Table I. addition of new particles (pumping). Individual events Note that the pumping term explicitly takes into account occur randomly, but their average rates are given by the the effect of pump blocking [18]. The spontaneous emis- terms on the right-hand sides of eqs. (1)-(2), as summa- sion β-factor is defined as the fraction of spontaneously rized in Table I. emitted photons which end up in the cavity mode, i.e. In Appendix A, we give a more detailed derivation of γrne γr γr the simulation algorithm used in [18], but essentially it β = = = . (3) can be boiled down to the iteration over many time incre- γtne γt γr + γbg ments dt, where one asks the question “how many events The rate equations (1)-(2) can be solved analytically of type µ happened during dt?” for each event type: stim- in steady state, yielding expressions for the mean pho- ulated emission, absorption, spontaneous emission, cav- ton number hnpi and number hnei of excited emitters as ity loss, background loss and pumping. In general, this is functions of the pump rate, γpn0. These steady-state so- not an easy question to answer, since each event that oc- lutions can exhibit characteristic features of laser opera- curs will immediately change the current number of pho- tion, namely emitter population clamping at large pump tons and/or excited emitters; and this in turn affects the rates and a sudden jump in the photon number as a func- rates aµ and the probability of subsequent events. How- tion of the pump rate. These features are characteristic ever, in [18] a few simplifying assumptions were made, of a crossing the lasing threshold, but for cer- which make it possible to approximate the answer: It tain parameter values, typically associated with β → 1, was assumed that there is some order in which event these features become less distinct, and instead the pho- types happen, in the sense that all the events of a par- ton statistics are used to characterise the transition to ticular type occur together, before all events of the next 3 type; it was assumed that one occurrence of an event does For instance, if there are ne(t) excited emitters at time not affect the probability of any other events of the same t and one is de-excited due to spontaneous emission into type occurring; and it was assumed that the binomial the lasing mode, then the number of excited emitters distributions for the number of occurring events can be available for stimulated emission should in principle be approximated by Poisson distributions. These assump- ne(t) − 1. However, if the stimulated emission event oc- tions are all reasonable when the value of dt is sufficiently curred first, then the situation is opposite. This ordering small. The algorithm is similar to the so-called τ-leap issue may only play a small role in the outcome of the method, described in detail in [22, 28]. simulations as long as the time increment dt is chosen Explicitly, the algorithm can be written as follows: sufficiently small, but this again leads to the question of how small is sufficient. All these challenges can be Fixed Time-Increment Method for Laser Rate ameliorated by use of another stochastic simulation algo- Equations (FTI) rithm, which will be described in more detail below. 1. Initialize: Set system parameters, set initial num- ber of photons np(t = 0) and excited emitters ne(t = 0).

2. Calculate rates aµ for all event types µ according B. Gillespie’s First Reaction Method for the Laser to Table I using current particle numbers. Rate Equations 3. Determine the number of each type of event µ oc- curing during dt by random draws from Poisson To obtain a faster algorithm, which has the added ben- distributions with parameters aµdt. efit of avoiding ordering issues as well as the need to define a time-step size, we can change the fundamental 4. Update n and n accordingly, set t → t + dt. p e question asked at each iteration to ”How long time until 5. If the maximal simulated time has been reached, the next event occurs?” and ”Which type of event will end. Otherwise, go to step 2. occur?”. In this way, the size of the time increment at each iteration will differ, and only one event will happen Solving the laser rate equations with this algorithm during each time increment. Changing the basic point of leads to equidistant for the number of intra- view in this way corresponds to applying the algorithm cavity photons and excited emitters, (np(ti),ne(ti)), and known as Gillespie’s First Reaction Method (FRM) [19], statistical convergence is ensured once reduction of the which is well-known in the chemistry and biochemistry time increment dt no longer affects the steady-state mean communities [22]. It is used in numerical calculations re- photon number hnpi, excited-emitter number hnei, and garding chemical reactions involving several species of 2 the photon variance h∆npi. In practice, the size of dt molecules with finite populations, assuming that each can be chosen at a given pump rate as some fraction type of reaction that can occur is a stochastic Markov dtfrac of the smallest reciprocal rate appearing in the rate process. The rate equations (1)-(2) describe an analogous equations (1)-(2), computed using the analytical steady- type of system, where the photons and excited emitters state values for the particle numbers. For instance, if the are analogous to the particles, and the events of emis- mean steady-state rate of spontaneous emission is the sion, absorption, loss and pumping are analogous to the largest among the steady-state rates in eqs. (1)-(2), then reactions that change the particle populations. dt = dtfrac/γrhnei. This ensures that most iterations in the low-pump limit involve at most one event of any Changing the operational question of the simulation in this way means that instead of generating six integers cor- type, and convergence is typically obtained for dtfrac on the order of 10−2 or less. For all FTI simulations in this responding to the change in particle numbers due to the −2 six different event types, we are interested in generating work we have used dtfrac = 10 . The mean photon and emitter numbers obtained using the FTI show the be- two numbers: One corresponding to the time until the haviour expected for lasers, and the steady-state photon next event, and one corresponding to the type of event. statistics, as computed using eqs. (4), also exhibit the As shown in [19], one way of doing this involves assum- characteristic transition from chaotic to coherent light. ing for each of the different event types that it happens This is shown in Fig. 3 of [18]. before any of the others, and then generating a tentative While the algorithm is stable and very easy to imple- time τµ until this event occurs. Once all the tentative ment, it also has a few downsides. First, the computa- times τµ have been determined, we can choose the short- tionally expensive act of drawing six Poisson-distributed est, τµ0 , as the actual time until the first event (or re- random numbers at each iteration leads to long compu- action) occurs, and the corresponding type of event, µ0, tation times, since many iterations are needed to obtain as the event that occurs. The tentative time τµ can be reliable photon statistics. Second, the assumption that efficiently generated by drawing a random number from all events of one type happen independently of all other an with parameter aµ, which is event types introduces some ambiguity in the ordering proved in Appendix A and in [19]. of the events which happen during each time-increment: Specifically, the algorithm can be stated as follows: 4

FIG. 2. (a)-(c) Steady-state analytical [full lines] and numerical [markers] results for the mean photon number, relative intensity 11 −1 noise and photon auto-correlation versus pump rate. For all plots, the cavity decay rate is γc = 10 s and the total decay 10 −1 −3 4 −1 2 rate is γt = 10 s . The β-factor and number of emitters, (β, n0), are varied between (10 , 2 · 10 ) [blue], (10 , 2 · 10 ) [red] and (1, 2 · 101) [black]. For β = 10−3 [blue], low-pump results have been discarded because of statistical uncertainty. (d) Computation time using the FTI [dashed lines] and FRM [full lines] algorithms for different lengths of simulated time at the pump rates indicated in (a). Computation times have been normalized to the lowest computation time, which is 0.01 s, obtained using MatLab on a PC with a 2.7 GHz quad-core processor and 16 GB DDR4 RAM. The parameters are the same as those used for the β = 0.1 plots in (a)-(c).

Gillespie’s First Reaction Method for Laser Rate FTI algorithm. Therefore, there is no need for additional Equations (FRM) checks of convergence in terms of such parameters. In this sense, the FRM is an exact method to stochastically 1. Initialize: Set system parameters, set initial num- simulate the laser dynamics. In addition, exponentially ber of photons np(t = 0) and excited emitters distributed random numbers τµ with parameters aµ can ne(t = 0). be quickly and inexpensively generated using uniformly distributed random numbers r on the unit interval (0, 1) 2. Calculate rates a for all event types µ according µ µ as τ = − log(r )/a [29, 30]. Therefore, step 3 in the to Table I using current particle numbers. µ µ µ FRM involves six draws from a uniform distribution, in-

3. For each µ, generate a tentative time increment τµ stead of the six draws from six different Poisson distri- by a random draw from an exponential distribution butions needed in the FTI algorithm. This drastically with parameter aµ. reduces the computation time of the FRM compared to the FTI, while the end results for the steady-state mean 4. Determine the event type µ0 for which particle numbers and photon statistics are the same for

τµ0 = minµ{τµ}. both algorithms. This is shown in Fig. 2 by reproducing the results of Fig. 3 in [18] through the use of the FRM, 5. Update np and ne according to event type µ0, set and comparing computation times versus simulated times t → t + τµ0 . for the two algorithms. We note that the very small mean 6. If the maximal simulated time has been reached, photon numbers obtained for the blue curves in the low- end. Otherwise go to step 2. pump range of Fig. 2 lead to relatively large statistical uncertainties. Therefore, results below a certain pump An important property of the FRM is that there are no rate have been discarded, like in the corresponding fig- numerical parameters, like the time increment dt in the ure of [18]. 5

It is clear from Fig. 2(d) that the computation time approaching a value slightly different from the character- 2 increases approximately linearly with the simulated time istic Poissonian h∆npi = hnpi. Equivalently, the Fano 2 for both the FTI and FRM algorithms, as expected. factor F = h∆npi/hnpi does not reach a value of 1 at However, as illustrated, the computation time at a fixed high pump rates. From the analytical expression in [18] simulated time also increases for increasing pump rates. for the photon variance we can show that at large values This is because the time delays between events become of the mean photon number we have smaller at higher pump rates, as the larger particle num- 2 bers lead to larger rates of the different events of emission, h∆npi F = ≈ 1+ c, for hnpi ≫ 1, (5) absorption and loss. Smaller time increments imply more hnpi iterations needed to obtain the same simulated time, and hence larger computation times. In all cases, it is clear where that the FRM algorithm leads to significantly shorter 1+ n /n computation times at all pump rates: Indeed, improve- c = th 0 . (6) ments of several orders of magnitude can be obtained by 2(n0/nth − 1) using the FRM algorithm instead of the FTI algorithm. Here n ≡ γ /γ is the semi-classical threshold value of We see that the computation time for the FRM algo- th c r rithm in the low-pump case is approximately constant for the emitter population inversion. The result in eq. (5) is illustrated in Fig. 3(c), where we have plotted the Fano the shortest simulated times. This seeming lower bound factor versus the pump rate, with the constant c = 0.75 on the computation time is related to the time for ini- tialization rather than having to do with the specific al- for both values of β. Physically speaking, this deviation from exact Poissonian statistics has its roots in the laser gorithm. level-scheme: Similar, non-Poissonian photon variances Since the FRM significantly reduces computation for lasers above threshold have been reported in earlier times, new features of the laser dynamics can easily be theoretical and experimental work [31–33], where two-, studied, which would otherwise be extremely time con- three- and four-level lasers are considered. In particu- suming. One such feature is the evolution of the pho- lar it is argued in [25, 31, 32] that the two-level emitter ton number distributions within the laser cavity as the scheme we consider here should generally lead to super- system transitions from chaotic, LED-like behaviour to Poissonian statistics above threshold due to depletion of coherent, laser-like behaviour. This will be examined in the emitter ground states. These effects are what we see the next section. in our analytical and numerical results. We note that the effects occur both when using the FTI and the FRM algorithms. IV. PHOTON DISTRIBUTIONS With the FRM algorithm it is possible to obtain pho- ton distributions which match the analytical predictions Once the laser settles into steady-state operation, the almost perfectly, while the computation time is kept rel- mean values for observables no longer change, but the atively short – less than 2 hours on a commercially avail- particle numbers still fluctuate in time. By examining the able laptop. This demonstrates once more the extreme simulated time series of the intra-cavity photon number efficiency and exactness of the FRM applied to the laser in steady state, we can obtain the discrete photon prob- rate equations. In the next section we will use this to ability density function (PDF) by evaluating how much show how small alterations in the laser rate equations of the total simulation time is spent with 1 photon in can lead to results which exhibit features that are typi- the cavity, how much is spent with 2 photons in the cav- cally obtained from much more intricate laser models. ity, etc. In Fig. 3 the photon PDF obtained from the stochastic simulations is shown for three different values of the pump rate using parameters corresponding to the V. BREAKING THE SYMMETRY BETWEEN cases of β = 1 and β = 10−3 from Fig. 2. In both cases SPONTANEOUS AND STIMULATED EMISSION there is a clear transition from thermal statistics giving RATES a Bose-Einstein distribution (d,g), through an intermedi- ate distribution near threshold (e,h), to near-Poissonian In the traditional rate-equation description of the laser statistics (f,i). dynamics, the rate per emitter of spontaneous emission As shown in Fig. 3(b), the photon auto-correlation into the cavity mode is the same as the rate per emitter g(2)(0) seems to indicate that Poissonian statistics are per photon of stimulated emission into the cavity mode, obtained at large pump rates for both β = 1 and consistent with the Einstein relations [26]. This can be β = 10−3, the simulated values being g(2)(0) = 1.0015 seen in eqs. (1)-(2), where the terms related to sponta- and g(2)(0) = 1.0001, respectively. However, we see neous and stimulated emission and absorption are all pro- in (f,i) that at high pump rates, the photon PDF ap- portional to the rate of radiative decay, γr. However, it proaches a Gaussian distribution which is wider than a is possible to imagine that certain physical effects, which with the same mean value. Math- are not accounted for in the traditional derivation of the ematically, we can explain this by the photon variance rate equations, could give rise to an effective difference 6

FIG. 3. Mean photon number (a), photon auto-correlation (b) and Fano-factor (c) versus pump rate for a laser with β = 1 − [black curves] and β = 10 3 [blue curves]. (d)-(f) Normalised photon number distribution for the laser with β = 1 at the pump − values indicated with black crosses in (a)-(c). (g)-(i) Normalised photon number distribution for the laser with β = 10 3 at the pump values indicated with blue circles in (a)-(c). The parameters used to make these plots are the same as those used to produce the curves in Fig. 2. In (d,g) the orange curves indicate Bose-Einstein distributions with mean values determined by the mean photon numbers hnpi. In (f,i) the orange dashed curves indicate Poisson distributions with parameters given by the mean photon number hnpi, and the red curves indicate Gaussian distributions with means and variances given by the mean 2 photon number hnpi and photon number variance h∆npi, respectively. between the rates of spontaneous and stimulated emis- the same small-signal analysis applied in [18] with the sion. For instance, collective effects due to correlations new radiative rates, we can also obtain corresponding 2 building up between emitters can affect the dynamics and ξ-dependent analytical expressions for h∆npi, RIN and steady-state behaviour of lasers, as has been shown in g(2)(0). Incorporating the asymmetry in the stochastic recent theoretical and experimental works [13, 34–39]. simulations using the FRM is simple, since we only need Some of these phenomena could potentially be captured to change the rates aµ to include ξ. In Fig. 4, we have by an asymmetry between the rates of spontaneous and plotted the steady-state mean intra-cavity photon num- stimulated emission, which is what we will investigate ber and photon auto-correlation as functions of the pump next. rate for three different values of ξ, illustrating the effects To examine the effects of introducing a difference be- of increasing or decreasing the spontaneous emission rate tween the spontaneous and stimulated radiative events, relative to the stimulated emission rate. We see several we can replace the common rate γr in eqs. (1)-(2) by effects: For small pump rates, the reduction of the rate sp st two separate rates, γr and γr , for the spontaneous and of spontaneous emission leads to a lower mean number stimulated events, respectively. The ratio of photons in the cavity as well as an increased value of the photon auto-correlation function to super-thermal ξ = γsp/γst (7) r r values. Conversely, increasing the spontaneous emission quantifies the difference between the rates of spontaneous rate leads to a larger mean intra-cavity photon number and stimulated emission (and absorption), and the rate and sub-thermal photon statistics in the low-pump limit. equations (1)-(2) can be rewritten using this quantity. In both cases, the results of the stochastic simulations We can solve the rate equations including ξ in the same follow the analytical results very well. In fact, we find way as we did for the traditional rate equations (1)- that there is a simple relationship between the value of ξ (2), both analytically and numerically. If we perform and the value of the photon auto-correlation function at 7

FIG. 4. Photon number (a,b) and photon auto-correlation function (c,d) as functions of the pump rate for different values of sp st st st the ratio ξ = γr /γr . All plots are produced using n0 = 20000. In (a,c) γr and hence nth = γc/γr = 10000 are kept fixed, sp sp −3 st while γr changes with ξ. In (b,d) γr and hence β = 10 are kept fixed, while γr and hence nth change with ξ. low pump rates, namely in a larger g(2)(0). The increased rate of stimulated emis- sion is also the reason for a larger mean photon number (2) g (0) → 1+1/ξ, for γp → 0. (8) in the high-pump limit seen in Fig. 4(b). The converse situation with a larger rate of spontaneous emission / These effects occur in part because the alteration of the lower rate of stimulated emission and absorption can be ratio ξ from 1 leads to changes in one of the derived pa- understood in a similar way. sp sp st rameters β = γr /(γr + γbg) or nth = γc/γr . Note that From recent experiments and theories regarding the well-known result g(2)(0) = 2 at low pump rates is re- nanolasers with collective correlations between the emit- stored in the case where the spontaneous and stimulated ters, it is known that collective effects in small lasers can sp st emission rates are equal, ξ = γr /γr = 1. result in a reduced mean number of photons in the cav- From Fig. 4 we see that decreasing the rate of sponta- ity and super-thermal photon statistics in the low-pump neous emission or increasing the rate of stimulated emis- limit [13, 38, 39]. Fig. 4 shows that these phenomena sion lead to the same qualitative effect, namely a reduced also appear from the laser rate equations if one breaks mean photon number and super-thermal photon statis- the symmetry between the spontaneous and stimulated tics. Correspondingly, increasing the rate of spontaneous emission rates dictated by the Einstein relations. The emission or decreasing the rate of stimulated emission effects are typically described theoretically by the emit- both lead to a larger mean photon number and sub- ter cross-correlations causing so-called excitation trap- thermal photon statistics. We can understand these ef- ping at small pump rates [38, 39], meaning excitations fects qualitatively as follows: Less spontaneous emission in the system preferably reside in the emitters rather will lead to fewer photons in the cavity in the low-pump than the cavity mode. This effect, which is related to regime where spontaneous emission dominates. This subradiance [40, 41], in turn gives rise to bursts of pho- means that there will be more excited emitters which can ton emission, similarly to what we described above in be prompted to emit through stimulated emission by the conjunction with the reduced rate of spontaneous emis- few photons in the cavity, which in turn leads to a higher sion. To see that such bursts of photons occur in our probability of several photons being present at the same simulations for ξ < 1, we have plotted in Fig. 5 the time, hence a larger auto-correlation. Equivalently, an time-resolved photon number for two values of ξ. As increased rate of stimulated events (stimulated emission, shown in Fig. 5(a,b), we examine the case where the absorption) means that any emitted photons are more steady-state mean photon number is the same, namely likely to be reabsorbed, giving fewer photons on average, hnpi = 0.2, while the photon distributions and statis- while each photon has a larger probability to set off a tics are either thermal or super-thermal. From (c,d) it stimulated emission from any excited emitter, resulting is clear that the case where ξ < 1, i.e. the rate of spon- 8

FIG. 5. (a) Photon auto-correlation function versus mean photon number for ξ = 1 [full blue line] and ξ = 1/4 [dashed red line]. Circles indicate simulated results with steady-state mean photon number hnpi = 0.2. (b) Simulated photon distribution for the two cases indicated by the circles in (a). (c,d) Time-resolved photon number for the two cases indicated in (a) during 10 ns of simulated time after reaching steady state. taneous emission is reduced, the photons often come in chaotic to coherent emission, and we have shown that the larger bursts, giving rise to the super-thermal statistics numerical results for the photon statistics follow the an- that we observe. This corresponds well with the theo- alytical results exactly. We have also applied the FRM retical descriptions given of the effects of emitter-emitter to altered laser rate equations, in which the symmetry correlations in nanolasers, meaning an effectively reduced between the spontaneous and stimulated emission rates rate of spontaneous emission may provide a simple and was broken. We found that by reducing the rate of spon- intuitive way of further studying and understanding col- taneous emission, or increasing the rate of stimulated lective effects in nanolasers. emission, both the numerical and analytical results show features that are also found in theories and experiments regarding lasers with strong emitter-emitter correlations. VI. SUMMARY AND DISCUSSION This suggests that there may be a way to use a slightly altered version of the well-known laser rate equations to In summary, we have presented Gillespie’s First Reac- describe collective effects in micro- and nanolasers. tion Method (FRM) in detail, and applied it to stochastic The FRM described here increases the efficiency of the simulations of the traditional laser rate equations. We stochastic simulations tremendously compared to meth- have shown that the FRM reduces computation times ods like the FTI, while remaining exact and conceptually by around three orders of magnitude compared to earlier simple. One of the reasons why the exact FRM algorithm algorithms [18], while it also avoids potential numerical is more efficient than the approximate FTI algorithm for issues related to the size of time-increments and the or- the laser rate equations is that there are relatively few dering of events. Thus, the FRM is an efficient and ex- event types and particles involved. Another reason is act algorithm for stochastic simulations of the laser rate the conservative choice of time-increment size used in the equations. We have used the FRM to examine the intra- FTI to obtain accurate results: In the FTI, and in the cavity photon distribution as the laser transitions from similar τ-leap method [22, 28], faster computation times 9 can always be obtained by choosing the time increments Appendix A: Derivation of the FTI and FRM to be larger, but this naturally reduces the accuracy of algorithms the simulation. In the low-pump limit, where the pho- ton statistics are very sensitive to the particle numbers, In this appendix, we will give a slightly more detailed high accuracy is needed, implying small time-increments description of why the FTI and FRM algorithms have and hence long computation times. In the high-pump the specific forms described in the main text. We will limit, where the particle numbers are large, it is possi- describe in some detail the probability theoretical con- ble that a less conservative choice of time-increment sizes siderations that go into the derivation of the algorithms, could reduce the computation times of the FTI algorithm though the full mathematical description is beyond the without significantly affecting the accuracy. Indeed, op- scope of this paper; see instead e.g. [19]. timizing the size of the time-increment may lead to com- As mentioned in Sec. III, our aim is to simulate the putation times for the FTI algorithm which are closer to interaction of a collection of discrete photons and emit- those for the FRM algorithm in the high-pump limit, at ters in a single-mode optical cavity, whose dynamics are a low cost to the accuracy. We leave an investigation of governed by the set of rate equations (1)-(2). To per- this to future work. form these stochastic simulations, we make a fundamen- tal assumption common to the stochastic formulation of Another exact and conceptually simple algorithm for chemical kinetics [19]: There exist a set of constants cµ stochastic simulation is Gillespie’s Direct Method [19, 20, for each type of event µ, which depend only on the physi- 42], which is similar to the FRM, but requires only two cal properties of the system (not particle numbers), such random draws per iteration. This could potentially in- that crease the efficiency of the simulation further, but due to the relatively low number of event types in the rate cµdt = Average probability, to first order in equations, the difference in computation time is expected dt, that one particular combination of to be minor. We performed a few simple simulation tests photons and/or emitters will undergo (A1) using Gillespie’s Direct Method, and for the simulated times needed to obtain reliable statistics, the prelimi- an event of type µ during the next nary results showed practically no difference in compu- time increment dt. tation time compared to the FRM. Both the FRM and the Direct Method have been expanded upon, leading to For instance, the probability that a particular photon in even more efficient algorithms [21, 22]. However, in most the cavity mode will be lost through the cavity mirrors cases the increase in efficiency happens at the cost of during a time increment dt, averaged over all photons the conceptual simplicity. For instance, the Next Reac- in the cavity mode, is γcdt + o(dt), where o(dt) are ex- tion Method introduced in [21] reduces the computation tra terms satisfying o(dt)/dt → 0 as dt → 0; In other words, c = γ . Likewise, the probability that a particu- time by storing and reusing the tentative times τµ and c c lar photon in the cavity mode will cause a specific excited only updating certain rates aµ, but it requires the intro- duction of additional concepts like dependency graphs emitter to emit in a stimulated emission event during dt and indexed priority queues. Since the rate equations for is γrdt + o(dt), so cst = γr. Additionally, if at time t the nanolasers deals with relatively few particles and event numbers of photons and excited emitters are np(t) and types, this added complexity would likely lead to fairly ne(t), respectively, we can define the functions hµ as small improvements in efficiency. However, such algo- hµ = Number of distinct combinations of photons rithms could be implemented in stochastic simulations of the laser rate equations in future work. and/or emitters that can undergo an event (A2) of type µ, given that the current particle Applying the FRM to stochastic simulations of the numbers are (np(t),ne(t)). laser rate equations leads to short computation times, which makes it possible to experiment more easily with For instance, the number of photons which could poten- changes in system parameters and even alterations of the tially leak out of the cavity is the current photon num- laser model. Using the FRM, one could potentially study ber, i.e. hc = np(t); and since each of the np(t) photons other characteristics of the laser as well, like the emission can prompt a stimulated emission event from each of the spectrum and linewidth, and the response to different ne(t) excited emitters, we take hst = ne(t)np(t). Com- types of pumping. Additionally, if the method can be bining the two quantities defined in eqs. (A1) and (A2), suitably modified, it can perhaps be used to further ex- we can define the rates, or propensity functions, aµ as amine the effects of inter-emitter coupling in nanolasers through simulations using an extended set of rate equa- aµdt = hµcµdt tions [43]. In conclusion, the FRM is efficient, robust = Probability, to first order in dt, and versatile, and we hope that this detailed description that an event of type µ will (A3) of the algorithm and how to apply it to the laser rate occur during the next time equations will lead to more new research and insights in the future. increment dt. 10

In [18], the stochastic simulations are carried out by and then we choose the smallest of these tentative times, choosing a fixed time increment dt and using the quanti- τµ0 , as the time until an event actually occurs. The cor- ties defined above to determine how many of each type of responding event type, µ0, is chosen as the event type event µ that happen at every time step: It is suggested that happens. To generate the random τµ’s, we should that the total number of events of type µ happening dur- consider the probabilities at time t for an event of type ing the time increment dt follows a , µ to happen between times t + τ and t + τ + dt. We may Binom(hµ,cµdt), based on the idea that each individ- compute such a probability as the product of the prob- ual event which happens can be seen as a success in a ability that no event of type µ happens between t and Bernoulli trial, i.e. a yes/no-experiment, whose proba- t + τ, given by bility of success is cµdt (Of course this assumes that dt is small enough that cµdt ≤ 1). As an example, the number P0(τ) = exp(−aµτ), (A4) of photons lost through the cavity mirrors between t and t + dt is taken to be a random integer drawn from the and the probability aµdt that an event of type µ does binomial distribution Binom(np(t),γcdt), since the aver- occur during an interval of length dt [19]. In other words, age probability of losing one photon is γcdt and there are the probabilities that we are looking for are of the form np(t) ”tries”. In order to speed up the simulations, a further assumption is made in [18], namely that the time Pµ(τ) dt = aµ exp(−aµτ) dt. (A5) increment dt may be chosen sufficiently small that the bi- To generate the random tentative times τ , we can there- nomial distributions Binom(h ,c dt) may be replaced by µ µ µ fore draw random numbers from the distributions P (τ), Poisson distributions Poiss(h c dt) = Poiss(a dt). This µ µ µ µ which are essentially exponential distributions with pa- removes the upper bound on the integers that may be rameters a . drawn, introducing the risk that e.g. the number drawn µ for how many photons are lost through the cavity mirrors is greater than the number of photons currently in the Appendix B: Funding cavity mode, but by choosing a sufficiently small time increment dt, this risk is minimal. While the replacement of binomial distributions by Villum Fonden (8692); Danish National Research Poisson distributions significantly reduces computation Foundation, grant number DNRF147. times for the algorithm of [18], the simulation must still run for several hours to obtain satisfactory photon statis- tics for all pump values. Implementing a more efficient Appendix C: Disclosures algorithm is therefore highly desirable, and to do this we can change the fundamental viewpoint in the simu- The authors declare no conflicts of interest. lation: Instead of determining how many events of each type occur during a fixed time increment, we can try to determine the length of time until some event occurs Appendix D: Acknowledgments and then determine which type of event occurs. There are several ways of doing this, but Gillespie’s First Re- The authors wish to thank I.E. Protsenko and A.V. action Method is based on the following idea: For each Uskov for stimulating discussions regarding collective ef- µ we can find a tentative time τµ until an event of type fects in nanolasers. J.M. wishes to thank G.L. Lippi for µ would occur, assuming no other event occurs before, helpful discussions regarding stochastic simulations.

[1] E. M. Purcell, Proceedings of the American Physical So- 13, 4999 (2013). ciety 69, 681 (1946). [9] A. H. Fikouras, M. Schubert, M. Karl, J. D. Kumar, S. J. [2] H. Yokoyama and S. D. Brorson, Journal of Applied Powis, A. Di Falco, and M. C. Gather, Nat. Commun. Physics 66, 4801 (1989). 9, 1 (2018). [3] H. Altug, D. Englund, and J. Vuˇckovi´c, Nat. Phys. 2, [10] R.-M. Ma and R. F. Oulton, Nat. Nanotech. 14, 12 484 (2006). (2019). [4] T. Suhr, N. Gregersen, K. Yvind, and J. Mørk, Opt. [11] C. Gies, J. Wiersig, M. Lorke, and F. Jahnke, Phys. Rev. Express 18, 11230 (2010). A 75, 013803 (2007). [5] C.-Y. A. Ni and S. L. Chuang, Opt. Express 20, 16450 [12] W. W. Chow, F. Jahnke, and C. Gies, Light: Science & (2012). Applications 3, e201 (2014). [6] D. A. Miller, Journal of Lightwave Technology 35, 346 [13] S. Kreinberg, W. W. Chow, J. Wolters, C. Schneider, (2017). C. Gies, F. Jahnke, S. H¨ofling, M. Kamp, and S. Reitzen- [7] M. C. Gather and S. H. Yun, Nat. Photon. 5, 406 (2011). stein, Light: Science & Applications 6, e17030 (2017). [8] G. Shambat, S.-R. Kothapalli, J. Provine, T. Sarmiento, [14] A. Moelbjerg, P. Kaer, M. Lorke, B. Tromborg, and J. Harris, S. S. Gambhir, and J. Vuckovic, Nano Lett. J. Mørk, IEEE Journal of Quantum Electronics 49, 945 11

(2013). 2004). [15] M. Marconi, J. Javaloyes, P. Hamel, F. Raineri, A. Lev- [30] V. Krishnan, Probability and Random Processes (John enson, and A. M. Yacomotti, Phys. Rev. X 8, 011013 Wiley & Sons, 2006). (2018). [31] A. V. Kozlovskii and A. N. Oraevsky, Quantum Electron- [16] A. Lebreton, I. Abram, N. Takemura, M. Kuwata- ics 24, 255 (1994). Gonokami, I. Robert-Philip, and A. Beveratos, New [32] A. V. Kozlovskii, Optics and Spectroscopy 116, 115 Journal of Physics 15, 033039 (2013). (2014). [17] G. P. Puccioni and G. L. Lippi, Opt. Express 23, 2369 [33] N. J. van Druten, Y. Lien, C. Serrat, S. S. R. Oem- (2015). rawsingh, M. P. van Exter, and J. P. Woerdman, Phys. [18] J. Mørk and G. L. Lippi, Appl. Phys. Lett. 112, 141103 Rev. A 62, 053808 (2000). (2018). [34] E. C. Andr´e, I. E. Protsenko, A. V. Uskov, J. Mørk, and [19] D. T. Gillespie, J. Comp. Phys. 22, 403 (1976). M. Wubs, Optics Letters 44, 1415 (2019). [20] D. T. Gillespie, J. Phys. Chem. 81, 2340 (1977). [35] J. G. Bohnet, Z. Chen, J. M. Weiner, D. Meiser, M. J. [21] M. A. Gibson and J. Bruck, J. Phys. Chem. A 104, 1876 Holland, and J. K. Thompson, Nature 484, 78 (2012). (2000). [36] D. Meiser, J. Ye, D. R. Carlson, and M. J. Holland, [22] J. Pahle, Brief. Bioinform. 10, 53 (2009). Phys. Rev. Lett. 102, 163601 (2009). [23] F. T. Arecchi, G. L. Lippi, and J. R. Puccioni, [37] D. Meiser and M. J. Holland, Phys. Rev. A 81, 033847 G. P. Tredicce, Opt. Commun. 51, 308 (1984). (2010). [24] G. Bjork and Y. Yamamoto, IEEE Journal of Quantum [38] H. A. M. Leymann, A. Foerster, F. Jahnke, J. Wiersig, Electronics 27, 2386 (1991). and C. Gies, Phys. Rev. Appl. 4, 044018 (2015). [25] P. R. Rice and H. J. Carmichael, Phys. Rev. A 50, 4318 [39] F. Jahnke, C. Gies, M. Aßmann, M. Bayer, H. A. M. Ley- (1994). mann, A. Foerster, J. Wiersig, C. Schneider, M. Kamp, [26] L. A. Coldren, S. W. Corzine, and M. L. Mashanovitch, and S. H¨ofling, Nat. Commun. 7, 1 (2016). Diode lasers and photonic integrated circuits (John Wi- [40] R. H. Dicke, Physical Review 93, 99 (1954). ley & Sons, 2012). [41] M. Gross and S. Haroche, Phys. Rep. 93, 301 (1982). [27] R. Loudon, The Quantum Theory of Light (Oxford Uni- [42] L. Chusseau, J. Arnaud, and F. Philippe, Optics and versity Press, 2000). Spectroscopy 94, 746 (2003). [28] D. T. Gillespie, J. Chem. Phys. 115, 1716 (2001). [43] I. Protsenko, E. C. Andr´e, A. Uskov, J. Mørk, and [29] C. P. Robert and G. Casella, M. Wubs, (2017), arXiv:1709.08200 [quant-ph]. Monte Carlo Statistical Methods (Springer New York,