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

Phase diffusion in gain-switched semiconductor lasers for quantum random number generation

Open Access Open Access

Abstract

In this paper, we report a theoretical and experimental study of the phase diffusion in a gain-switched single-mode semiconductor laser. Our modelling includes nonlinear dependencies of the carrier recombination rate and the semiconductor material gain on the carrier number. Also a comparison with the results obtained with linear models is performed. We focus on the below threshold operation, that is, the regime where most of the phase diffusion occurs in quantum random number generators based on the above mentioned system. Using the extracted parameters for our laser and simple analytical expressions for the laser linewidth, we show that logarithmic material gain and cubic carrier recombination dependence on the carrier number must be considered for a good agreement between experiments and theory. In this way we show that consideration of these nonlinearities is essential for obtaining a good quantitative description of the phase diffusion when using rate equations modelling.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Random number generation (RNG) is widely used in many applications including cryptographycally secured communications, Monte Carlo simulation, industrial testing, quantitative finance, gambling, etc. [1,2]. Both pseudorandom number generators (PRNGs) based on deterministic algorithms, and random number generators based on hardware physical devices, have been used for the previous applications. Examples of physical processes used in these generators include: radioactive decay, Zener noise, Johnson’s noise [1,2], chaos noise [2,3] and quantum phenomena [1]. Quantum random number generators (QRNGs) are a particular case of hardware physical RNGs in which the data are obtained from quantum events. The main advantage of QRNGs is that their randomness is inherent to quantum mechanics making quantum systems a perfect source of entropy for random number generation [2].

Most of the existing QRNGs are based on quantum optics [2]. Light from single-photon sources, semiconductor lasers or light emitting diodes are used in these generators. QRNGs based on single-photon detection methods include: two-path splitting of single photons [4], generators measuring the time of arrival of photons [5], photon counting generators [6], and attenuated pulse generators [7]. The properties of some of these methods have been experimentally compared in [8]. Multiphoton QRNGs have also been proposed and experimentally and theoretically demonstrated. For instance, generators based on quantum vacuum fluctuations [9], on amplified spontaneous emission (ASE) signals [10], and on phase noise in continuous wave [1113] and in pulsed semiconductor lasers [1417] including full integration on an InP platform [18], have been demonstrated.

In these last generators the current applied to a single-mode semiconductor laser is periodically modulated from a well below threshold value to a value above threshold, in such a way that gain-switching operation appears, typically at Gbps rates [1419]. While the laser is below threshold the optical phase becomes random due to the spontaneous emission noise. The laser emits then a series of pulses with a random phase. Phase fluctuations are converted into amplitude fluctuations by using an unbalanced Mach-Zehnder interferometer, with a delay matching the pulse repetition period. Detection and filtering of the amplitude provides the generation of random values with an arcsine distribution. Fast generation rates, up to 43 Gbps quantum random bit generation, have been experimentally shown [15]. There are several advantages in using QRNGs based on pulsed gain-switching of single-mode laser diodes. They are made of commercially available components and the high signal level permits the use of standard photodetectors. They are simple, fast, robust, have low cost, and operate with flexible clock frequencies.

The description and understanding of the physical processes underlying the operation of a QRNG are essential for their applications. For instance in cryptography, randomness is very often used for the generation of a secret key. QRNGs based on gain-switched laser diodes are appropriate for Quantum Key Distribution (QKD) systems because high speed random bit generation for state preparation is required [17,20]. For these type of QRNGs it is essential to have a good quantitative description of the phase diffusion process in which their fluctuations are characterized and measured.

Theoretical and experimental studies of the fluctuations of laser light began in the decade of the sixties [2125]. Study of fluctuations of the light emitted by semiconductor lasers has also received a lot of attention [2631]. A good theoretical description, deduced from first principles and valid for below and above threshold operation, is based on the Fokker-Planck equation or alternativelly on stochastic rate equations of Langevin’s type [23,32]. Good quantitative agreement between experiments and theory can only be achieved when the complete set of parameters of the stochastic rate equations is known for the specific laser diode. Having this knowledge permits a good quantitative description of the dependence of phase diffusion on the laser and modulation parameters.

Modelling of phase fluctuations in gain-switched semiconductor lasers for QRNG has been performed by numerical integration of the laser stochastic rate equations [15,17,18,33,34]. All these models have considered a linear dependence of the carrier recombination rate, the spontaneous emission rate, and the semiconductor material gain on the carrier density. Very recently, numerical simulations using a model in which the spontaneous emission rate has a quadratic dependence have been performed and compared with results from models using linear terms for that rate [35]. These numerical simulations use the complete set of parameters that have been extracted for a single-mode discrete mode laser (DML) [35,36]. Using these parameters a very good quantitative agreement between experiments and theory for describing optical frequency comb formation has been shown for a very wide range of gain-switching conditions [36,37].

In this paper we report an experimental and theoretical study of the phase diffusion in the gain-switched DML characterized in [35,36]. Our theoretical modelling includes nonlinear dependencies of the carrier recombination rate and the semiconductor material gain on the carrier number. We compare with the results obtained with linear models in order to know if these linear approximations are able to quantitatively describe the phase diffusion process. We focus our attention in the below threshold operation that is the regime where most of the phase diffusion occurs in QRNG. Using the simple analytical expressions for the laser linewidth derived from the Schawlow-Townes law, and the extracted parameters for our laser, we show that nonlinear dependence of both, carrier recombination rate and semiconductor material gain must be considered for a good quantitative description of the phase diffusion. In particular, good agreement between experiments and theory is obtained when a 2-parameter logarithmic material gain and a 3-parameter cubic carrier recombination are considered. We show that going beyond the usual linear dependencies is important to quantitatively describe the measured nonlinear dependence of the spectral linewidth, and hence of the phase diffusion coefficient, as a function of the bias current.

Our paper is organized as follows. In section 2, we present our theoretical model. Section 3 is devoted to the calculation of simple expressions of the spectral width and their comparison with our experimental results. In section 4, we analyze theoretically and experimentally the phase diffusion for below and above threshold operation. Finally, in section 5 we discuss the validity of the linear approximations and summarize our results.

2. Theoretical model

Gain-switched single-mode semiconductor laser dynamics can be modelled by using a set of stochastic rate-equations that read (in Ito’s sense) [26,35,36,38]

$$\frac{d P}{d t} = \left[ \frac{G(N)}{1+\epsilon P} -\frac{1}{\tau_p} \right] P + \beta B N^2 + \sqrt{2 \beta BP}N F_p (t)$$
$$\frac{d \phi}{d t} = \frac{\alpha}{2} \left[ G(N) -\frac{1}{\tau_p} \right] + \sqrt{\frac{\beta B}{2P} }N F_\phi (t)$$
$$ \frac{d N}{d t} = \frac{I}{e} - (AN+BN^2+CN^3)- \frac{G(N)P}{1+\epsilon P} $$
where $P(t)$ is the number of photons inside the laser, $\phi (t)$ is the optical phase, and $N(t)$ is the number of carriers in the active region. The parameters appearing in these equations are the following: $G(N)$ is the unsaturated gain, $\epsilon$ is the non-linear gain coefficient, $\tau _p$ is the photon lifetime, $\beta$ is the fraction of spontaneous emission coupled into the lasing mode, $\alpha$ is the linewidth enhancement factor, $I$ is the injected current, and $e$ is the electron charge, respectively. In these equations we have considered a carrier recombination rate, $R(N)$, given by
$$R(N)=AN+BN^{2}+CN^{3},$$
where $A, B$ and $C$ are the non-radiative, spontaneous, and Auger recombination coefficients, respectively. We consider two types of unsaturated gain, $G(N)$: the linear gain, $G_\textrm {lin}(N)=G_N(N-N_t)$ (where $G_N$ is the differential gain and $N_t$ is the carrier number at transparency), and the logarithmic gain, $G_\textrm {log}(N)=G_0\ln (N/N_0)$. The Langevin terms $F_P(t)$ and $F_\phi (t)$ in Eqs. (1)–(2), represent fluctuations due to spontaneous emission, with the following correlation properties, $< F_i(t) F_j(t^\prime )> = \delta _{ij} \delta (t-t^\prime )$, where $\delta (t)$ is the Dirac delta function and $\delta _{ij}$ the Kronecker delta function with the subindexes $i$ and $j$ referring to the variables $P$ and $\phi$.

Numerical integration of the previous stochastic rate equations present instabilities when the photon number is very small because $P$ can become negative when the values of the current are smaller than the threshold current, $I_{th}$, a situation always found when using these lasers for QRNG. This problem can be solved by integrating the equivalent equations for the complex electrical field [35]. These equations read

$$\frac{d E}{d t} = \left[ \left(\frac{1}{1+\epsilon \mid E\mid^2}+i\alpha\right)G(N) -\frac{1+i\alpha}{\tau_p} \right] \frac{E}{2} + \sqrt{\beta B}N \xi (t)$$
$$\frac{d N}{d t} = \frac{I}{e} -(AN+BN^2+CN^3) - \frac{G(N)\mid E\mid^2}{1+\epsilon \mid E\mid^2}$$
where $E(t)=E_1(t)+iE_2(t)$ is the complex electrical field and $\xi (t)=\xi _1 (t)+i\xi _2(t)$ is the complex Gaussian white noise with zero average and correlation given by $< \xi (t) \xi ^*(t^\prime )> = \delta (t-t^\prime )$ that represents the spontaneous emission noise. This model corresponds to our initial equations because the application of the rules for the change of variables in the Ito’s calculus [32] to $P=\mid E\mid ^2=E_1^2+E_2^2$ and $\phi = \arctan {(E_2/E_1)}$ in Eqs. (5)–(6) gives Eqs. (1)–(3). There are no instabilities because $P$ does not appear in the square root factor that multiplies the noise term. We will numerically solve Eq. (5) and Eq. (6) by using the Euler-Maruyama algorithm [23,39] with an integration time step of 0.001 ps. We will use the numerical values of the parameters that have been extracted for a discrete mode laser (DML) [3537]. This device is a single longitudinal mode semiconductor laser emitting close to 1550 nm wavelength. For modelling this device we have usually considered $G(N)=G_\textrm {lin}(N)$. Simulation and experimental results have shown not only qualitative but also a remarkable quantitative agreement for a very wide range of gain-switching conditions [3537]. Although some of the formulas that we will use are well known and can be found in standard laser diode textbooks [26,27] we have preferred to include them in the paper to set the notation and to make a coherent and self-contained description in order to perform a clear comparison between different models.

Analysis of phase diffusion in gain-switched laser diodes has been done by using a rate-equation model in which the carrier recombination and the unsaturated gain are assumed to be linear [15,17,18,33,34] not only for QRNG but also for quantum key distribution using the same laser systems [20]. These models are based on a linearization of $R(N)$ around the carrier number at threshold, $N_{th}$ [26,40], where $N_{th}=N_t+1/(G_N\tau _p)$. This linear expansion reads $R(N)\sim R(N_{th})+(N-N_{th})/\tau _n$, where $\tau _n$ is the differential carrier lifetime at threshold that is given by $\tau _n^{-1}=\bigl (\frac {dR}{dN}\bigr )_{N_{th}}$. Substituting this expansion in Eq. (3) and using that the threshold current is given by $I_{th}=eR(N_{th})$ we obtain

$$\frac{d N}{d t} = \frac{I-I_{th}}{e} -\frac{N-N_{th}}{\tau_n}- \frac{G_N (N-N_t)P}{1+\epsilon P}$$
that is valid only when $N\sim N_{th}$. The carrier lifetime at threshold, $\tau _e$, is given by $\tau _e=N_{th}/R(N_{th})=eN_{th}/I_{th}$. It is easy to find $\tau _e$ and $\tau _n$ in terms of the parameters of the model
$$\tau_e=\frac{1}{A+BN_{th}+CN_{th}^{2}},$$
$$\tau_n=\frac{1}{\frac{1}{\tau_e}+BN_{th}+2CN_{th}^{2}},$$
where we have used the definition of $\tau _n$ and Eq. (4). Substitution of Eq. (9) in Eq. (7) gives
$$\frac{d N}{d t} = \frac{I}{e} -\frac{N}{\tau_e}- (N-N_{th})(BN_{th}^2+2CN_{th}^2)-\frac{G_N (N-N_t)P}{1+\epsilon P}$$
that can be approximated by the usual rate equation [15,17,18,33,34]
$$\frac{d N}{d t} = \frac{I}{e} -\frac{N}{\tau_e}- \frac{G_N (N-N_t)P}{1+\epsilon P}$$
only when $N\sim N_{th}$. Equation (11) gives good quantitative results if the condition $N\sim N_{th}$ is satisfied, that is when $I\geq I_{th}$ or when $I$ is slightly below $I_{th}$. This means that when a square wave modulation of the current is used ($I(t)=I_{on}$ during $T/2$, and $I(t)=I_{off}$ during the rest of the period, $T$) $I_{off}$ must be close to $I_{th}$. Using $I_{off}$ considerably smaller than $I_{th}$ requires the use of Eq. (3) for a good quantitative agreement between simulations and experiments, so the consideration of the nonlinear carrier recombination given by Eq. (4) is essential for achieving that agreement. This is particularly important for describing the stochastic evolution of the phase in QRNG based on laser gain-switching because most of the phase diffussion occurs when the laser is switched off, that is when the current is $I_{off}$, which value is typically well below $I_{th}$. This discussion will be performed at the beginning of Section 5.

3. Spectral width below threshold: comparison between theory and experiments

In this section we first focus on the calculation of the width of the optical spectrum. In the second part of this section we compare with experiments performed with a discrete mode semiconductor laser (DML). The spectral linewidth below threshold is given by the usual Schawlow-Townes law that is accurate for the amplified spontaneous emission (ASE) regime [26]. This law reads

$$\Delta\nu = \frac{R_{sp}(\bar{N})}{2\pi \bar{P}}= \frac{\beta B\bar{N}^2}{2\pi \bar{P}}$$
where $\Delta \nu$ is the FWHM of the optical spectrum, $R_{sp}(N)$ is the rate of the spontaneous emission coupled into the lasing mode, and $\bar {P}$ and $\bar {N}$ are the averaged steady-state values of the photon and carrier numbers, respectively [26]. In order to calculate these values we make $dP/dt=dN/dt=0$ in Eq. (1) and Eq. (3) and take the average to obtain
$$ 0 = \left( G(\bar{N}) -\frac{1}{\tau_p} \right) \bar{P} + \beta B \bar{N}^{2}$$
$$ 0 = \frac{I}{e} - (A\bar{N}+B\bar{N}^{2}+C\bar{N}^{3}) $$
where we have used that the last term of Eq. (3) can be neglected and $\epsilon \bar {P} << 1$ because $I$ is well below $I_{th}$. Using Eq. (13) in Eq. (12) we obtain $\Delta \nu = \frac {1}{2\pi }\left ( \frac {1}{\tau _p} - G(\bar {N})\right )$ that can be written in terms of the gain function by noting that $1/\tau _p=G(N_{th})$:
$$\Delta\nu = \frac{1}{2\pi}\left( G(N_{th}) - G(\bar{N})\right).$$

Very simple expressions are obtained for the two considered unsaturated gain functions, $G_\textrm {lin}(N)$, and $G_\textrm {log}(N)$. These read

$$\Delta\nu_\textrm{{lin}} = \frac{G_N}{2\pi}\bigl( N_{th}-\bar{N}\bigr)$$
and
$$\Delta\nu_\textrm{{log}} = \frac{G_0}{2\pi}\ln \Bigl( \frac{N_{th}}{\bar{N}}\Bigr)$$
for the linear and logarithmic gain, respectively. $\bar {N}$ is a function of $I$, $\bar {N}=\bar {N}(I)$, that is obtained by solving the third order Eq. (14), and $N_{th}=\bar {N}(I_{th})$. The real root of Eq. (14) is given by:
$$\bar{N}(I) = \left( -\frac{q}{2}+\sqrt{D}\right)^{1/3}-\left( \frac{q}{2}+\sqrt{D}\right)^{1/3}-\frac{B}{3C}$$
where $q=(2B^3-9ABC)/(27C^3)-I/(eC)$, $D=(p/3)^3+(q/2)^2$, and $p=(A-B^2/(3C))/C$.

Our experiments are performed using a single-mode DML. Details on the device, measurement procedure and extraction of the parameters of the model are detailed in [3537]. The parameters have been extracted using $G_\textrm {lin}(N)$ and their values are the following: $G_N=1.48\times 10^4$s$^{-1}$, $N_t=1.93\times 10^7$, $\epsilon = 7.73\times 10^{-8}$, $\tau _p = 2.17$ ps, $\alpha =3$, $\beta =5.3\times 10^{-6}$, $A=2.8\times 10^8$ s$^{-1}$, $B=9.8$ s$^{-1}$, and $C=3.84\times 10^{-7}$ s$^{-1}$ [3537]. We note that measurements for extracting parameters have been performed for bias current above threshold, a condition for which $G_\textrm {lin}(N)$ works well. Measurements have been performed fixing the temperature at 25$^o$C. For this value $I_{th}$ is 14.14 mA, the emission wavelength is 1546.985 nm at $I$=30 mA, and the wavelength separation between consecutive longitudinal modes is 1.28 nm. $I_{th}$ is a parameter in our model that is used for calculating $N_{th}$ from $N_{th}=\bar {N}(I_{th})=5.045\times 10^7$. This value coincides to that obtained by using $G_\textrm {lin}(N)$, that is $N_t+1/(G_N\tau _p)$. We also note that the values of $\tau _n$ and $\tau _e$ are rather different: using Eq. (8) and Eq. (9) we obtain that $\tau _n=$0.24 ns and $\tau _e$= 0.57 ns.

The laser linewidth (FWHM) as a function of $I$ when $I<I_{th}$ is shown with squares in Fig. 1. Data close to threshold have been obtained with a high resolution (0.08 pm) optical spectrum analyzer. Data corresponding to currents below 13 mA have been obtained with another optical spectrum analyzer with worse wavelength resolution (0.06 nm). As the current decreases, $\Delta \nu$ has the nonlinear increase typically found in ASE operation. We first try to describe that increase by using Eq. (16), derived for $G_\textrm {lin}(N)$. Nonlinear behaviour is also expected for $\Delta \nu _\textrm {lin}$ because $\bar {N}$ does not depend linearly on $I$, as seen from Eq. (18). Figure 1 also shows the results obtained with Eq. (16). $\Delta \nu _\textrm {lin}$ describes well the experimental results close to $I_{th}$ as expected since $G_\textrm {lin}(N)$ works well close to threshold [40]. However it greatly underestimates the linewidth as $I$ decreases: $\Delta \nu _\textrm {lin}$ is 32 and 50 $\%$ smaller than the experimental values when $I=0.7 I_{th}$ and $I=0.5I_{th}$, respectively.

 figure: Fig. 1.

Fig. 1. Measured linewidth as a function of $I$ for bias currents below threshold. Theoretical results obtained with Eq. (16) and with Eq. (17) are plotted with black and red solid lines, respectively.

Download Full Size | PDF

In order to try to improve the theoretical description of the linewidth we now consider the logarithmic dependence of $G(N)$, $G_\textrm {log}(N)=G_0\ln (N/N_0)$. However we do not know the values of the $G_0$ and $N_0$ parameters. A very simple way of obtaining $G_0$ is to use Eq. (17) to fit the experimental results because $\Delta \nu _\textrm {log}$ only depends on $G_0$. We consider then the value of $G_0$ for which we obtain equal values in experiments and theory at an intermediate value of Fig. 1 ($I=9.75$ mA$\sim 0.7 I_{th}$), that is $G_0=1.05\times 10^{12} s^{-1}$. The value of $N_0$ can then be obtained from $G_\textrm {log}(N_{th})=1/\tau _p$, that is $N_0=N_{th}\exp (-1/(G_0\tau _p))=3.25\times 10^7$. These values are similar to those found in the literature. For instance typical $g_0$ and $n_t$ in [26] are 1800 cm $^{-1}$ and 2.2 $\times 10^{18}$ cm $^{-3}$, respectively. If we calculate the corresponding $g_0$ and $n_t$ for our device we get: $g_0=G_0/(\Gamma v_g)=2040$ cm $^{-1}$ and $n_t=N_0/V_{act}=2.13\times 10^{18}$ cm $^{-3}$, where we have used the confinement factor, $\Gamma$, group velocity, $v_g$, and active volume, $V_{act}$, corresponding to our device [36].

The description of the linewidth significantly improves if $G_\textrm {log}(N)$ is used. Figure 1 shows that $\Delta \nu _\textrm {log}$ describes much better than $\Delta \nu _\textrm {lin}$ the measured linewidth: $\Delta \nu _\textrm {log}$ is only 18 $\%$ smaller than the experimental values when $I=0.5I_{th}$. These results show that considering the logarithmic dependence of the unsaturated material gain on the carrier number is necessary for a good description of the linewidth below threshold.

Another confirmation of the good results obtained with the logarithmic gain is by plotting the optical spectrum calculated from the numerical simulation of our model. We numerically integrate Eqs. (5)–(6) with $G_\textrm {log}(N)$ using a 2.5 ps sampling time and a 81.92 ns temporal window. The optical spectrum is calculated by averaging over 1000 temporal windows. We compare in Fig. 2 the theoretical optical spectrum obtained with the corresponding experimental one obtained for a well-below threshold current (9 mA). A good agreement between experiment and theory is found.

 figure: Fig. 2.

Fig. 2. Optical spectrum for a constant bias current of 9 mA. Experimental and simulated results are plotted with black and red solid lines, respectively. Zero frequency corresponds to a 1546.54 nm wavelength.

Download Full Size | PDF

4. Phase diffusion

Studies on phase diffusion in gain-switched laser diodes for QRNG have been based on the numerical simulation of stochastic rate equations. Most of these analysis consider multiplicative noise for the three variables: $P, \phi$, and $N$, in a similar way to Eqs. (1)–(3). It has been shown that noise in the carrier equation is not important for the statistical properties of the phase [30,31,34]. Two different dependencies of the spontaneous emission rate on $N$, linear and quadratic, have been considered giving rise to different multiplicative noise terms in equations for $P$ and $\phi$. Comparison between results obtained with both dependencies has been recently performed [35] to show that the quadratic dependence is the best choice, in agreement with [26]. Since these numerical studies are aimed to study phase diffusion in QRNG the bias current is modulated, typically with a squared wave dependence like that described in Section 2.

Analytical studies of phase diffusion have not considered modulation of the bias current. They just consider a constant value of $I$ and additive noises in which the steady-state average values of the variables appear in the terms that multiply the Langevin terms, $F_P(t)$ and $F_\phi (t)$ [27,29]. In fact, these terms have been derived from first principles for a system where the matter and the radiation have reached equilibrium [22]. Taking this into account we consider the following equations

$$\frac{d P}{d t} = \left[ \frac{G(N)}{1+\epsilon P} -\frac{1}{\tau_p} \right] P + \beta B N^2 + \sqrt{2 \beta B\bar{P}}\bar{N} F_p (t)$$
$$ \frac{d \phi}{d t} = \frac{\alpha}{2} \left[ G(N) -\frac{1}{\tau_p} \right] + \sqrt{\frac{\beta B}{2\bar{P}} }\bar{N} F_\phi (t)$$
together with Eq. (3) with constant $I$ in order to analyze the phase diffusion. We first consider the phase statistics when the current is below threshold. When the steady-state corresponding to $I<I_{th}$ has been reached the phase diffuses in such a way that its variance, $\sigma _\phi ^2$, increases linearly with $t$ with diffusion coefficient, $D_\phi$ [29,32], that is
$$\sigma_\phi^2=2D_\phi t=\frac{\beta B\bar{N}^2}{2\bar{P}}t$$

Taking into account Eq. (12) and that $\bar {N}$ and $\bar {P}$ obtained from Eqs. (19)–(20) and Eq. (3) are equal to those obtained from Eqs. (1)–(3) we get that

$$\sigma_\phi^2=\pi \Delta\nu t$$
where $\Delta \nu$ is given in general by Eq. (15), and specifically by Eq. (17) because it is the best approximation when $I<I_{th}$.

We now consider the phase statistics when $I>I_{th}$. The spectral linewidth above threshold is given by the "modified" Schawlow-Townes law [26,29]

$$\Delta\nu_>{=} \frac{R_{sp}(N_{th})}{4\pi \bar{P}}(1+\alpha^2)= \frac{\beta BN_{th}^2}{4\pi \bar{P}}(1+\alpha^2)$$
that can be written in terms of $I$ in the following way [40]:
$$\Delta\nu_>{=} \frac{e\beta BN_{th}^2(1+\alpha^2)}{4\pi \tau_p (I-I_{th})}$$

When the steady-state corresponding to $I>I_{th}$ has been reached the phase diffuses in such a way that its variance, $\sigma _\phi ^2$, increases linearly with $t$ with diffusion coefficient, $D_\phi$ [27,29], that is

$$\sigma_\phi^2=2D_\phi t=\frac{\beta BN_{th}^2}{2\bar{P}}(1+\alpha^2)t$$

Taking into account Eq. (23) we get that

$$\sigma_\phi^2=2\pi \Delta\nu_> t$$
where $\Delta \nu _>$ is given by Eq. (24).

The phase diffusion coefficient $D_\phi$ is given by

$$D_\phi = \begin{cases} \pi\Delta\nu /2 & \textrm{if}\,\, I<I_{th} \\ \pi\Delta\nu_> & \textrm{if}\,\, I>I_{th}. \end{cases}$$

In Fig. 3 we show $D_\phi$ as a function of the bias current, for values below and above $I_{th}$. In this figure we have considered the logarithmic gain dependence, that is $\Delta \nu$ given by Eq. (17). $D_\phi$ decreases with $I$ in such a way that around threshold it is dramatically reduced: 3 orders of magnitude in a narrow current range (from 13 to 16 mA). The agreement between experimental and theoretical results below threshold is good. We have not measured values above threshold because of the resolution limit of our equipment (10 MHz) although we should expect also good agreement between theory and experiments, as was shown in [40]. A gap in the theoretical results is apparent from Fig. 3. This corresponds to the transition around $I_{th}$ because Eq. (24) is not valid if $I=I_{th}$. However we note that the results obtained from that equation have been plotted for $I\geq$14.16 mA=1.001 $I_{th}$, a value very close to threshold.

 figure: Fig. 3.

Fig. 3. Phase diffusion coefficient as a function of the bias current. Measured and theoretical results are plotted with squares and solid line, respectively

Download Full Size | PDF

In Fig. 4 we show the experimental and theoretical values of the standard deviation of the phase after 1 ns (in $2\pi$ units), $\sigma _\phi (1\, \textrm {ns})/(2\pi )=\sqrt {D_\phi /2}/\pi$, as a function of $I$. $\sigma _\phi (1 \,\textrm {ns})$ is reduced several orders of magnitude when the current is increased above threshold: for instance it goes from 4$\pi$ at $I$=7 mA to 0.018$\pi$ at $I=$ 30 mA, that is a reduction in a factor of 222. Our analytical and experimental results are in agreement with those obtained from numerical simulation [33,41]. The system analyzed in this work is a true quantum entropy source when the standard deviation of the phase at the end of the period is greater than 2$\pi$ ($\sigma _\phi > 2\pi$) [34,42]. The good quantitative agreement between experiment and theory permits to assure that the previous condition is fulfilled after 1 ns time evolution if $I<11.6 mA\sim 0.8 I_{th}$.

 figure: Fig. 4.

Fig. 4. Standard deviation of the phase after 1 ns (in 2$\pi$ units) as a function of the bias current. Measured and theoretical results are plotted with squares and solid line, respectively

Download Full Size | PDF

5. Discussion and conclusions

Up to now we have discussed the importance of considering the logarithmic gain dependence in order to obtain a realistic quantitative description of the phase diffusion when the laser is below threshold, that is the regime where most of that diffusion occurs in QRNGs using laser diodes. This description has been made using nonlinear carrier recombination terms. We now discuss the validity of considering a linear carrier recombination, approximation that has been usually followed for modelling phase diffusion for QRNG and QKD using gain-switched semiconductor lasers [15,17,18,20,33,34]. In this approximation Eq. (11) describes the change of the carrier number. Using this equation the steady state value of $N$ corresponding to the below threshold operation depends linearly on $I$: $\bar {N}=\tau _e I/e$. Substitution of this expression in Eq. (16) and Eq. (17) gives

$$\Delta\nu_\textrm{{lin}}^\prime = \frac{G_N\tau_e}{2\pi e}\bigl( I_{th}-I\bigr)$$
and
$$\Delta\nu_\textrm{{log}}^\prime = \frac{G_0}{2\pi}\ln \Bigl( \frac{I_{th}}{I}\Bigr),$$
respectively. The linewidths calculated with a linearised $R(N)$ have been denoted with primes in order to distinguish them from the linewidths calculated in section 3. Figure 5 shows the comparison between the linewidths obtained with nonlinear (already shown in Fig. 1) and linearised $R(N)$. Also the measured values have been included in the figure. Linewidths obtained from models with linearised $R(N)$ and a linear gain dependence [15,17,18,20,33,34], $\Delta \nu _\textrm {{lin}}^\prime$, do not describe well the experimental results because Eq. (28) gives a linear dependence on $I$ that it is not observed in the experimental results. The linewidth obtained with linearised $R(N)$ and a logarithmic gain dependence, $\Delta \nu _\textrm {{log}}^\prime$, does not describe the experimental results either. Therefore the best description, $\Delta \nu _\textrm {{log}}$, is given by the nonlinear $R(N)$ and $G_\textrm {log}(N)$.

 figure: Fig. 5.

Fig. 5. Measured linewidth as a function of $I$ for bias currents below threshold. Theoretical results obtained with Eq. (28) and with Eq. (29) are plotted with black and red dashed lines, respectively. Theoretical results obtained with Eq. (16) and with Eq. (17) and nonlinear $R(N)$ have been also plotted with black and red solid lines, respectively.

Download Full Size | PDF

The equations that have been used to obtain the analytical expression of the phase diffusion coefficient, Eqs. (19) and (20), are not the same than Eqs. (1) and (2). Spontaneous emission noise terms in Eqs. (1) and (2) are good approximations to those of Eqs. (19) and (20), that correspond to the exact form that was deduced from first principles for a system where the matter and the radiation have reached equilibrium [22,31]. When the matter-radiation equilibrium has not yet been achieved, the exact form of the spontaneous emission strength is unknown [31]. In situations where a high-frequency modulation of the bias current is applied, like those found in QRNG, the laser is in a transient regime and the assumption of the validity of the spontaneous emission strength in Eqs. (1) and (2) cannot be justified in a rigorous way [31]. Nevertheless, noise terms like those in Eqs. (1) and (2) have been used for studying the statistics during the transient regime of laser diodes showing good agreement between experiments and theory [30,38,4350]. A different set of rate equations in which additive noises have strengths given by the time dependent averages of the variables have been recently proposed [34]. Statistics obtained with these types of modelling could be different because of the different nature of noise terms. This comparison will be the subject of future work. However, a first estimation of the frequencies for which we would expect similar results when using Eqs. (1)–(2) and Eqs. (19)–(20) can be done by evaluating the time it takes the system to reach the steady state conditions after switching-off the current. Under typical modulation conditions used in QRNG this time is several hundreds of ps [35], so similar results would be expected if the modulation frequency is around 1 GHz or smaller. Extension of our analytical calculations to faster modulation frequencies can be done by modifying in an appropriate way the theory developed in [49].

Typical rate equations, like those used in this work, assume certain approximations. For instance $B$ is taken as a constant value although it is a parameter that depends on the carrier density [26,51]. It means that considering a single value for $B$, whether it is obtained with above or below threshold estimations, will not be enough to achieve perfect agreement between theoretical and experimental results. Therefore, using rate equations for analyzing the gain switched semiconductor laser for QRNG, in which the evolution above and below threshold is described with a single model, implies the consideration of these types of approximations.

Experimental results corresponding to parameter extraction of our laser were obtained in [35,36] but a few of them were assumed, in particular, $B$. The value of $B$ was chosen based on reported values of high speed 1.55 micron lasers. We chose the value $B=9.8 s^{-1}$ that corresponds to $B_d=1.5 \times 10^{-10} cm^3s^{-1}$ [52] (where $B_d$ is the corresponding recombination parameter written in equations for carrier and photon densities). The range of variation of $B_d$ is not very large. For instance, $B_d$ varies from $0.6 \times 10^{-10} cm^3s^{-1}$ to $2.4 \times 10^{-10} cm^3s^{-1}$, depending on different material composition, carrier density and temperature [26,51]. Our chosen value of $B$ is in the middle of the previous range. We can study the implications of assuming this value by evaluating how the theoretical linewidth changes from the value we calculated in Fig. 1 with respect to the values obtained at the extremes of the previous range. We have calculated $\Delta \nu _\textrm {log}$ for $B=3.92 s^{-1}, 9.8 s^{-1}$, and $15.68 s^{-1}$, that corresponds to $B_d=0.6\times 10^{-10} cm^3s^{-1}, 1.5\times 10^{-10} cm^3s^{-1}$ , and $2.4\times 10^{-10} cm^3s^{-1}$, respectively. The largest changes of the linewidth are obtained for the smallest values of the current and are only 4% and 9% for $B=3.92 s^{-1}$, and $15.68 s^{-1}$, respectively. So the implications of our assumption for the experimental validation performed in this work are small.

Our results have been obtained only with one laser. It would be desirable to have more experiments to support the generalization of our results. We think that our results could hold for other types of single-mode edge emitting lasers (DBR, DFB$\ldots$) because their dynamics and the way in which spontaneous emission noise affects that dynamics are similar. In fact, stochastic rate equations modelling is a common tool to describe the statistics of the emitted light in all those devices. Some recent results [53] have compared the gain-switching performance of the laser analyzed in this manuscript (Discrete Mode Laser) with DFB lasers, indicating that they are very similar: the driving conditions rather than the device properties determine the coherence of the emitted signal when generating optical frequency combs. Extending these experimental comparisons to the case of gain-switching for QRNG would be interesting in order to support the generalization of our results.

Summarizing, we have analyzed theoretically and experimentally the phase diffusion problem in a gain-switched single-mode semiconductor laser. We have shown the importance of a good knowledge of the values of the laser parameters and the need of considering logarithmic material gain and nonlinear recombination of carriers for achieving a good agreement between theory and experiments. This is essential for a good quantitative description of QRNGs based on gain-switched laser diodes.

Funding

Ministerio de Ciencia, Innovación y Universidades (MCIN/AEI/10.13039/501100011033, RTI2018-094118-B-C22).

Acknowledgments

A. Quirce acknowledges financial support from Beatriz Galindo program, Ministerio de Ciencia, Innovación y Universidades (Spain). Angel Valle would like to thank Marcos Valle Miñón for his help in the calculations.

Disclosures

The authors declare no conflicts of interest.

Data availability

The data that support the plots within this letter and other findings of this study are available from the corresponding authors upon reasonable request.

References

1. M. Stipcevic and C. K. Koç, “True random number generators,” in Open Problems in Mathematics and Computational Science (Springer, 2014), pp. 275–315.

2. M. Herrero-Collantes and J. C. Garcia-Escartin, “Quantum random number generators,” Rev. Mod. Phys. 89(1), 015004 (2017). [CrossRef]  

3. M. Sciamanna and K. A. Shore, “Physics and applications of laser diode chaos,” Nat. Photonics 9(3), 151–162 (2015). [CrossRef]  

4. T. Jennewein, U. Achleitner, G. Weihs, H. Weinfurter, and A. Zeilinger, “A fast and compact quantum random number generator,” Rev. Sci. Instrum. 71(4), 1675–1680 (2000). [CrossRef]  

5. M. Stipčević and B. M. Rogina, “Quantum random number generator based on photonic emission in semiconductors,” Rev. Sci. Instrum. 78(4), 045104 (2007). [CrossRef]  

6. H. Fürst, H. Weier, S. Nauerth, D. G. Marangon, C. Kurtsiefer, and H. Weinfurter, “High speed optical quantum random number generation,” Opt. Express 18(12), 13029–13037 (2010). [CrossRef]  

7. W. Wei and H. Guo, “Bias-free true random-number generator,” Opt. Lett. 34(12), 1876–1878 (2009). [CrossRef]  

8. T. Durt, C. Belmonte, L.-P. Lamoureux, K. Panajotov, F. Van den Berghe, and H. Thienpont, “Fast quantum-optical random-number generators,” Phys. Rev. A 87(2), 022339 (2013). [CrossRef]  

9. Y. Shen, L. Tian, and H. Zou, “Practical quantum random number generator based on measuring the shot noise of vacuum states,” Phys. Rev. A 81(6), 063814 (2010). [CrossRef]  

10. A. Argyris, E. Pikasis, S. Deligiannidis, and D. Syvridis, “Sub-tb/s physical random bit generators based on direct detection of amplified spontaneous emission signals,” J. Lightwave Technol. 30(9), 1329–1334 (2012). [CrossRef]  

11. H. Guo, W. Tang, Y. Liu, and W. Wei, “Truly random number generation based on measurement of phase noise of a laser,” Phys. Rev. E 81(5), 051137 (2010). [CrossRef]  

12. B. Qi, Y.-M. Chi, H.-K. Lo, and L. Qian, “High-speed quantum random number generation by measuring phase noise of a single-mode laser,” Opt. Lett. 35(3), 312–314 (2010). [CrossRef]  

13. F. Xu, B. Qi, X. Ma, H. Xu, H. Zheng, and H.-K. Lo, “Ultrafast quantum random number generation based on quantum phase fluctuations,” Opt. Express 20(11), 12366–12377 (2012). [CrossRef]  

14. M. Jofre, M. Curty, F. Steinlechner, G. Anzolin, J. Torres, M. Mitchell, and V. Pruneri, “True random numbers from amplified quantum vacuum,” Opt. Express 19(21), 20665–20672 (2011). [CrossRef]  

15. C. Abellán, W. Amaya, M. Jofre, M. Curty, A. Acín, J. Capmany, V. Pruneri, and M. Mitchell, “Ultra-fast quantum randomness generation by accelerated phase diffusion in a pulsed laser diode,” Opt. Express 22(2), 1645–1654 (2014). [CrossRef]  

16. Z. Yuan, M. Lucamarini, J. Dynes, B. Fröhlich, A. Plews, and A. Shields, “Robust random number generation using steady-state emission of gain-switched laser diodes,” Appl. Phys. Lett. 104(26), 261112 (2014). [CrossRef]  

17. R. Shakhovoy, V. Sharoglazova, A. Udaltsov, A. Duplinskiy, V. Kurochkin, and Y. Kurochkin, “Influence of chirp, jitter, and relaxation oscillations on probabilistic properties of laser pulse interference,” IEEE J. Quantum Electron. 57, 1 (2021). [CrossRef]  

18. C. Abellan, W. Amaya, D. Domenech, P. Muñnoz, J. Capmany, S. Longhi, M. W. Mitchell, and V. Pruneri, “Quantum entropy source on an inp photonic integrated circuit for random number generation,” Optica 3(9), 989–994 (2016). [CrossRef]  

19. D. G. Marangon, A. Plews, M. Lucamarini, J. F. Dynes, A. W. Sharpe, Z. Yuan, and A. J. Shields, “Long-term test of a fast and compact quantum random number generator,” J. Lightwave Technol. 36(17), 3778–3784 (2018). [CrossRef]  

20. K. Nakata, A. Tomita, M. Fujiwara, K.-i. Yoshino, A. Tajima, A. Okamoto, and K. Ogawa, “Intensity fluctuation of a gain-switched semiconductor laser for quantum key distribution systems,” Opt. Express 25(2), 622–634 (2017). [CrossRef]  

21. M. Lax, “Quantum noise. iv. quantum theory of noise sources,” Phys. Rev. 145(1), 110–129 (1966). [CrossRef]  

22. M. Lax and W. Louisell, “Quantum noise. xii. density-operator treatment of field and population fluctuations,” Phys. Rev. 185(2), 568–591 (1969). [CrossRef]  

23. H. Risken, “Fokker-planck equation,” in The Fokker-Planck Equation (Springer, 1996).

24. C. H. Henry and R. F. Kazarinov, “Quantum noise in photonics,” Rev. Mod. Phys. 68(3), 801–853 (1996). [CrossRef]  

25. F. Arecchi, V. Degiorgio, and B. Querzola, “Time-dependent statistical properties of the laser radiation,” Phys. Rev. Lett. 19(20), 1168–1171 (1967). [CrossRef]  

26. L. A. Coldren, S. W. Corzine, and M. L. Mashanovitch, Diode lasers and photonic integrated circuits, vol. 218 (John Wiley & Sons, 2012).

27. G. P. Agrawal and N. K. Dutta, Semiconductor lasers (Springer Science & Business Media, 2013).

28. P. Spano, S. Piazzolla, and M. Tamburrini, “Phase noise in semiconductor lasers: A theoretical approach,” IEEE J. Quantum Electron. 19(7), 1195–1199 (1983). [CrossRef]  

29. C. Henry, “Phase noise in semiconductor lasers,” J. Lightwave Technol. 4(3), 298–311 (1986). [CrossRef]  

30. S. Balle, P. Colet, and M. San Miguel, “Statistics for the transient response of single-mode semiconductor laser gain switching,” Phys. Rev. A 43(1), 498–506 (1991). [CrossRef]  

31. S. Balle, F. De Pasquale, N. Abraham, and M. San Miguel, “Statistics of the transient frequency modulation in the switch-on of a single-mode semiconductor laser,” Phys. Rev. A 45(3), 1955–1966 (1992). [CrossRef]  

32. C. W. Gardiner, Handbook of stochastic methods, vol. 3 (springer, Berlin, 1985).

33. B. Septriani, O. de Vries, F. Steinlechner, and M. Gräfe, “Parametric study of the phase diffusion process in a gain-switched semiconductor laser for randomness assessment in quantum random number generator,” AIP Adv. 10(5), 055022 (2020). [CrossRef]  

34. R. Shakhovoy, A. Tumachek, N. Andronova, Y. Mironov, and Y. Kurochkin, “Phase randomness in a gain-switched semiconductor laser: stochastic differential equation analysis,” arXiv preprint arXiv:2011.10401, (2020).

35. A. Quirce and A. Valle, “Spontaneous emission rate in gain-switched laser diodes for quantum random number generation,” submitted to Opt. Laser Technol. (2021).

36. A. Rosado, A. Pérez-Serrano, J. M. G. Tijero, A. Valle, L. Pesquera, and I. Esquivias, “Numerical and experimental analysis of optical frequency comb generation in gain-switched semiconductor lasers,” IEEE J. Quantum Electron. 55(6), 1–12 (2019). [CrossRef]  

37. A. Quirce, A. Rosado, J. Díez, A. Valle, A. Pérez-Serrano, J.-M. G. Tijero, L. Pesquera, and I. Esquivias, “Nonlinear dynamics induced by optical injection in optical frequency combs generated by gain-switching of laser diodes,” IEEE Photonics J. 12(4), 1–14 (2020). [CrossRef]  

38. N. Schunk and K. Petermann, “Noise analysis of injection-locked semiconductor injection lasers,” IEEE J. Quantum Electron. 22(5), 642–650 (1986). [CrossRef]  

39. P. E. Kloeden and E. Platen, Stochastic differential equations,” in Numerical Solution of Stochastic Differential Equations (Springer, 1992).

40. P. Pérez, A. Valle, I. Noriega, and L. Pesquera, “Measurement of the intrinsic parameters of single-mode vcsels,” J. Lightwave Technol. 32(8), 1601–1607 (2014). [CrossRef]  

41. C. Abellán Sánchez, “Quantum random number generators for industrial applications,” PhD Thesis (2018).

42. R. Shakhovoy, D. Sych, V. Sharoglazova, A. Udaltsov, A. Fedorov, and Y. Kurochkin, “Quantum noise extraction from the interference of laser pulses in an optical quantum random number generator,” Opt. Express 28(5), 6209–6224 (2020). [CrossRef]  

43. P. Spano, A. D’Ottavi, A. Mecozzi, and B. Daino, “Experimental observation of time jitter in semiconductor laser turn-on,” Appl. Phys. Lett. 52(26), 2203–2204 (1988). [CrossRef]  

44. A. Mecozzi, S. Piazzolla, A. D’Ottavi, and P. Spano, “Passage time statistics in semiconductor laser turn on,” Phys. Rev. A 38(6), 3136–3138 (1988). [CrossRef]  

45. A. D’Ottavi, A. Mecozzi, P. Spano, and S. Piazzolla, “Time jitter in multimode fabry–perot laser diodes,” Appl. Phys. Lett. 53(24), 2362–2364 (1988). [CrossRef]  

46. A. Mecozzi, P. Spano, and A. Sapia, “Quasi-deterministic theory of semiconductor-laser gain switching,” Opt. Lett. 15(19), 1067–1069 (1990). [CrossRef]  

47. P. Spano, A. Mecozzi, and A. Sapia, “Statistical distribution of trajectories in the time-intensity plane during semiconductor-laser gain switching,” Phys. Rev. Lett. 64(25), 3003–3006 (1990). [CrossRef]  

48. A. Czylwik and W. Eberle, “Transient intensity noise of semiconductor lasers: Experiments and comparison with theory,” IEEE J. Quantum Electron. 26(2), 225–230 (1990). [CrossRef]  

49. A. Valle, M. Rodriguez, and C. R. Mirasso, “Analytical calculation of timing jitter in single-mode semiconductor lasers under fast periodic modulation,” Opt. Lett. 17(21), 1523–1525 (1992). [CrossRef]  

50. A. Mecozzi, A. Sapia, P. Spano, and G. P. Agrawal, “Transient multimode dynamics in nearly single-mode lasers,” IEEE J. Quantum Electron. 27(3), 332–343 (1991). [CrossRef]  

51. J. Minch, S. H. Park, T. Keating, and S. L. Chuang, “Theory and experiment of In Ga As P and In Ga Al As long-wavelength strained quantum-well lasers,” IEEE J. Quantum Electron. 35(5), 771–782 (1999). [CrossRef]  

52. S. Mohrdiek, H. Burkhard, F. Steinhagen, H. Hillmer, R. Losch, W. Schlapp, and R. Gobel, “10-Gb/s standard fiber transmission using directly modulated 1.55-μm quantum-well DFB lasers,” IEEE Photonics Technol. Lett. 7(11), 1357–1359 (1995). [CrossRef]  

53. A. Rosado, A. Pérez-Serrano, J.-M. G. Tijero, A. Valle, L. Pesquera, and I. Esquivias, “Experimental study of optical frequency comb generation in gain-switched semiconductor lasers,” Opt. Laser Technol. 108, 542–550 (2018). [CrossRef]  

Data availability

The data that support the plots within this letter and other findings of this study are available from the corresponding authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. Measured linewidth as a function of $I$ for bias currents below threshold. Theoretical results obtained with Eq. (16) and with Eq. (17) are plotted with black and red solid lines, respectively.
Fig. 2.
Fig. 2. Optical spectrum for a constant bias current of 9 mA. Experimental and simulated results are plotted with black and red solid lines, respectively. Zero frequency corresponds to a 1546.54 nm wavelength.
Fig. 3.
Fig. 3. Phase diffusion coefficient as a function of the bias current. Measured and theoretical results are plotted with squares and solid line, respectively
Fig. 4.
Fig. 4. Standard deviation of the phase after 1 ns (in 2$\pi$ units) as a function of the bias current. Measured and theoretical results are plotted with squares and solid line, respectively
Fig. 5.
Fig. 5. Measured linewidth as a function of $I$ for bias currents below threshold. Theoretical results obtained with Eq. (28) and with Eq. (29) are plotted with black and red dashed lines, respectively. Theoretical results obtained with Eq. (16) and with Eq. (17) and nonlinear $R(N)$ have been also plotted with black and red solid lines, respectively.

Equations (29)

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

d P d t = [ G ( N ) 1 + ϵ P 1 τ p ] P + β B N 2 + 2 β B P N F p ( t )
d ϕ d t = α 2 [ G ( N ) 1 τ p ] + β B 2 P N F ϕ ( t )
d N d t = I e ( A N + B N 2 + C N 3 ) G ( N ) P 1 + ϵ P
R ( N ) = A N + B N 2 + C N 3 ,
d E d t = [ ( 1 1 + ϵ E 2 + i α ) G ( N ) 1 + i α τ p ] E 2 + β B N ξ ( t )
d N d t = I e ( A N + B N 2 + C N 3 ) G ( N ) E 2 1 + ϵ E 2
d N d t = I I t h e N N t h τ n G N ( N N t ) P 1 + ϵ P
τ e = 1 A + B N t h + C N t h 2 ,
τ n = 1 1 τ e + B N t h + 2 C N t h 2 ,
d N d t = I e N τ e ( N N t h ) ( B N t h 2 + 2 C N t h 2 ) G N ( N N t ) P 1 + ϵ P
d N d t = I e N τ e G N ( N N t ) P 1 + ϵ P
Δ ν = R s p ( N ¯ ) 2 π P ¯ = β B N ¯ 2 2 π P ¯
0 = ( G ( N ¯ ) 1 τ p ) P ¯ + β B N ¯ 2
0 = I e ( A N ¯ + B N ¯ 2 + C N ¯ 3 )
Δ ν = 1 2 π ( G ( N t h ) G ( N ¯ ) ) .
Δ ν {lin} = G N 2 π ( N t h N ¯ )
Δ ν {log} = G 0 2 π ln ( N t h N ¯ )
N ¯ ( I ) = ( q 2 + D ) 1 / 3 ( q 2 + D ) 1 / 3 B 3 C
d P d t = [ G ( N ) 1 + ϵ P 1 τ p ] P + β B N 2 + 2 β B P ¯ N ¯ F p ( t )
d ϕ d t = α 2 [ G ( N ) 1 τ p ] + β B 2 P ¯ N ¯ F ϕ ( t )
σ ϕ 2 = 2 D ϕ t = β B N ¯ 2 2 P ¯ t
σ ϕ 2 = π Δ ν t
Δ ν > = R s p ( N t h ) 4 π P ¯ ( 1 + α 2 ) = β B N t h 2 4 π P ¯ ( 1 + α 2 )
Δ ν > = e β B N t h 2 ( 1 + α 2 ) 4 π τ p ( I I t h )
σ ϕ 2 = 2 D ϕ t = β B N t h 2 2 P ¯ ( 1 + α 2 ) t
σ ϕ 2 = 2 π Δ ν > t
D ϕ = { π Δ ν / 2 if I < I t h π Δ ν > if I > I t h .
Δ ν {lin} = G N τ e 2 π e ( I t h I )
Δ ν {log} = G 0 2 π ln ( I t h I ) ,
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.