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

Stochastic nonlinear model of the dynamics of actively Q-switched lasers

Open Access Open Access

Abstract

In this paper, we present a novel stochastic and spatially lumped multi-mode model to describe the nonlinear dynamics of actively Q-switched lasers and random perturbations due to amplified spontaneous emission. This model will serve as a basis for the design of (nonlinear) control and estimation strategies and thus a high value is set on its computational efficiency. Therefore, a common traveling-wave model is chosen as a starting point and a number of model-order reduction steps are performed. As a result, a set of nonlinear ordinary differential equations for the dynamic behavior of the laser during a switching cycle is obtained. A semi-analytic solution of these differential equations yields expressions for the population inversion after a switching cycle and for the output energy, which are then used to formulate a nonlinear discrete-time model for the pulse-to-pulse dynamics. Simulation studies including models with different levels of complexity and first experimental results demonstrate the feasibility of the proposed approach.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Nanosecond laser pulses are widely used in areas such as materials processing [1], spectroscopy [2], plasma research [3] and medical applications [4], where high pulse energies are required, and also in lidar systems [5]. The most established technology for achieving such pulses is active Q-switching. Since high pulse repetition rates are desirable for many applications [6], current research efforts are focussed on building actively Q-switched lasers which can operate at up to 1 MHz. At such frequencies and usual pump powers, large variations in output pulse energies which, in most cases, are detrimental to the intended application can ultimately damage optical elements inside the cavity. These energy variations are primarily the result of two distinct effects. Firstly, spontaneous emission is a stochastic process with known properties [7] which influences the shape of a pulse as well as its total energy ([8], Figs. 12 and 13). Secondly, it is already known from regenerative amplifiers [9] that the pulse-to-pulse dynamics of pumping and inversion depletion exhibits instability regions where pulse energies are elements of nontrivial limit cycles: strong pulses which entail substantial population depletion and therefore strongly weaken the subsequent pulse’s gain are followed by weak pulses which allow the populations to recover etc. In the case of actively Q-switched lasers, simulations ([10], Fig. 13) and experiments [11] have hinted at similar instabilities which ultimately yield (deterministically) chaotic behavior. For Yb-doped fiber lasers, the pulse-to-pulse instability regions at higher switching frequencies were studied comprehensively in [12], along with other phenomena such as stimulated Brillouin scattering and premature pulse build-ups powered by lower or stronger levels of stochastic seed from amplified spontaneous emission.

Hardware-based techniques have been applied to mitigate stochastic [13] and dynamic effects [14] but limit the flexibility of the system. These limitations could be avoided by employing active feedback control strategies as in [9], which in turn requires accurate modeling of the system’s behavior. A review of existing mathematical models of actively Q-switched lasers can be found in [15]. These models always involve some form of rate equation(s) for the lasing population densities and hence the optical gain of the active medium. A second set of equations then describes the evolution of optical energy within the cavity depending on the current gain and losses. On a more detailed level, we will differentiate mathematical models along three mutually independent key characteristics:

  • 1. Spatially distributed or spatially lumped: In the former approach, population densities and optical powers are assumed spatially dependent (especially along the axis of beam propagation), while the latter disregards power transport phenomena and concentrates populations and powers into one point, see, e.g., [16]. Studies have shown [15,17] that spatially distributed modeling achieves superior accuracy especially for long active media.
  • 2. Multi-mode or single-mode: Multi-mode models take into account the spectrally varying and potentially competing gain/loss characteristics of different longitudinal optical cavity modes as pointed out e.g. in [18]. Several techniques have been invented [19] to select a single longitudinal mode and avoid pulse energy variations resulting from mode-hopping in a mode competition. Single-mode models assume a single effective gain and loss and represent radiation by one longitudinal mode, preventing the simulation of spectral selection with the benefit of shorter overall simulation times.
  • 3. Stochastic or deterministic: Contributions of spontaneous emission to the optical power are essential especially for the modeling of self-seeded lasers as they define the initial intracavity power level. If said contributions are represented by stochastic processes (deterministic quantities), we will call a model stochastic (deterministic). To the authors’ knowledge, the stochastic nature of spontaneous emission has so far not been systematically considered in the literature on Q-switched lasers.
The current state of the art is the deterministic and spatially distributed multi-mode approach of traveling-wave modeling as used recently, e.g., in [2022]. Traveling-wave models feature spatially dependent rate equations as well as power transport equations for every longitudinal mode. The complex behavior resulting from the coupled nonlinear partial differential equations is useful for simulation studies, but renders an analysis of the pulse-to-pulse dynamics difficult. However, its superior accuracy to other existing models also makes traveling-wave modeling a good starting point for this paper. Using significant simplifications, Degnan [23] started off with a deterministic and spatially lumped single-mode model and arrived at a transcendental relation between the population inversions before and after a pulse. This is the first and, to the authors’ knowledge, only explicit expression for pulse-to-pulse dynamics so far. Unfortunately, the assumption in this paper that the energy build-up always goes to the threshold (the point where the power gains equal the losses and no more amplification occurs) is no longer valid at higher pulse repetition rates, which makes this work not applicable to our problem. Therefore, we first present a first principles-based model of actively Q-switched lasers which employs a stochastic generalization of traveling-wave modeling where spontaneous emission is represented by a Bose-Einstein distributed noise term. For the inhomogeneous gain, this paper uses the exemplary case of Nd:YAG active media. Taking the general model as a reference, a number of successive simplification steps are then performed to derive a similarly accurate, but spatially lumped continuous-time stochastic multi-mode model and a discrete-time model which explicitly describes the pulse-to-pulse dynamics. Such a discrete-time model is beneficial for three reasons: Firstly, instead of computationally demanding and time-consuming numerical simulations for the complete Q-switching cycle, a single (semi-)analytic expression is given for the population inversion and output pulse energy, respectively. Secondly, the model and its derivation throw light on the question how existing models from the literature are connected, as we will discuss in Section 3.4. Finally, the main motivation for this model is the fact that it serves as a theoretical basis for the design of control and estimation algorithms to actively stabilize the pulse-to-pulse dynamics and to suppress stochastic influences on the intracavity power. Thus, Section 2. first establishes the general framework and the fundamental equations of the model. The step-by-step simplifications are then performed in Section 3. In Section 4, we show simulation results which illustrate the stochastic and dynamic contributions leading to energy fluctuations. Based on these simulation results, agreement with first measurements and theoretical results on pulse statistics and modeling errors introduced by the various simplification steps are assessed in Section 4. as well. The paper concludes with a summary of the presented work and an outlook in Section 5.

2. General model

2.1 Rate and transport equations

Contrary to regenerative amplifiers, the cavities of Q-switched lasers are constantly filled with continuous-wave optical energy, while the pulse is created via the Q-switching process and is usually of greater duration than the cavity round-trip time. Throughout this work, we will make the usual assumption that no two laser modes of different wavelengths are mutually coherent and thus that no interference phenomena such as mode-locking occur. Therefore, radiation will be represented by powers instead of electric fields. Our starting point will be the (multi-mode) traveling-wave equations, where we use the example of a forward pumped Nd:YAG crystal within a linear (Fabry-Perot) cavity. A sketch of the system is drawn in Fig. 1. The spatial coordinate inside the cavity along which light propagates is defined as $z$-direction, and the active medium with population densities $N_i(z,t)$ resides within the interval $[0,L]$. Population densities and optical powers will always be assumed to be transversally uniform, which is why there will be no spatial dependences other than on $z$. At $z = z_{PC}$, a Pockels cell with a reflection coefficient $R(t)$ performs the Q-switching. All optical components apart from the active medium and the Q-switch determine the boundary conditions at either end of the cavity, which are specified in Section 2.2. For Nd:YAG active media, the most interesting lasing transitions occur at wavelengths $\lambda _A$ = 946 nm and $\lambda _B$ = 1064 nm. Their technological relevance is due to a small quantum defect and hence high theoretical efficiency in the former and a large emission cross-section in the latter [24], which renders the 1064 nm transition the dominant one. Pumping can be achieved at $\lambda _p$ = 809 nm from the ground state to an elevated state, which quickly decays to the upper state for both lasing transitions considered. Thus, our full model of the active medium effectively consists of five population densities as visualised in Fig. 2. The rate equations for the population densities $N_i(z,t)$ with the mentioned underlying energy level scheme and a total of $M$ longitudinal cavity modes read [15]

$${ \frac{\partial {N_5(z,t)}}{\partial {t}} } ={-}\gamma_{54} N_5(z,t) + \gamma_{45} N_4(z,t) + \frac{ \lambda_{p} \sigma_{p}}{h c A_p} P_p(z,t) \left(N_1(z,t) - N_5(z,t)\right)$$
$$\begin{aligned} &{ \frac{\partial {N_4(z,t)}}{\partial {t}} } = \gamma_{54} N_5(z,t) - \gamma_{45} N_4(z,t) - (\gamma_{42} + \gamma_{43}) N_4(z,t) - \sum_{j=1}^{M} \frac{ \lambda_j}{h c A_s} \times\\ &\times \left(P^{+}(z,t,\lambda_j) + P^{-}(z,t,\lambda_j)\right)\Big(\sigma_A(\lambda_j) \left(N_4(z,t) - N_2(z,t)\right) +\sigma_B(\lambda_j) \left(N_4(z,t) - N_3(z,t)\right) \Big) \end{aligned}$$
$$\begin{aligned} & { \frac{\partial {N_3(z,t)}}{\partial {t}} } ={-}\gamma_{31} N_3(z,t) + \gamma_{13} N_1(z,t) + \gamma_{43} N_4(z,t)\\ & + \sum_{j=1}^{M} \frac{ \lambda_j}{h c A_s} \left(P^{+}(z,t,\lambda_j) + P^{-}(z,t,\lambda_j)\right)\Big(\sigma_B(\lambda_j) \left(N_4(z,t) - N_3(z,t)\right) \Big) \end{aligned}$$
$$\begin{aligned} &{ \frac{\partial {N_2(z,t)}}{\partial {t}} } ={-}\gamma_{21} N_2(z,t) + \gamma_{12} N_1(z,t) + \gamma_{42} N_4(z,t)\\ & + \sum_{j=1}^{M} \frac{ \lambda_j}{h c A_s} \left(P^{+}(z,t,\lambda_j) + P^{-}(z,t,\lambda_j)\right)\Big(\sigma_A(\lambda_j) \left(N_4(z,t) - N_2(z,t)\right) \Big) , \end{aligned}$$
with relaxation rates $\gamma _{ij}$ for the populations and absorption and induced emission terms due to the pump power $P_p(z,t)$ as well as forward- and backward-running lasing powers $P^{\pm }(z,t,\lambda _j)$. Wavelengths $\lambda _j$ belong to a discrete spectrum of cavity modes. The cross-sectional areas of the pump and the signal beams are given by $A_p$ and $A_s$, respectively. The absorption/emission cross-sections (which we set equal for simplicity) are denoted by $\sigma _A(\lambda _j)$ and $\sigma _B(\lambda _j)$ for the transition wavelengths $\lambda _A$ and $\lambda _B$, respectively. They determine the spectral lineshapes of the two transitions (which in general may overlap). $N_{dop}$ is the crystal’s total Neodymium dopand density, and $N_1(z,t)$ can be calculated from the other population densities via the algebraic equation $N_1(z,t) = N_{dop} - N_5(z,t) - N_4(z,t) - N_3(z,t) - N_2(z,t)$. For setups such as cladding-pumped lasers, one also needs to include appropriate overlapping factors $\Gamma$, see, e.g., [15].

 figure: Fig. 1.

Fig. 1. A sketch of the cavity, active medium (red area), Q-switch (Pockels cell, yellow area), forward pump power (blue) and forward- and backwards-running signal powers (green). The coupling of the two beam propagation directions as well as all other optical components are accounted for via the boundary conditions B.C. (green dots).

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. The five energy occupation levels that constitute our model and their transitions: pumping (blue) and two lasing transitions (red, green). The black arrows stand for nonradiative thermal transitions.

Download Full Size | PDF

Because active media usually have very similar refractive indices at comparable wavelengths $\lambda _A$, $\lambda _B$, and $\lambda _p$, the group velocities of radiation in this spectral range are all approximately equal to a common value $v$. The power transport equations can then be written as ([25], Cha. 8)

$$\frac{1}{v}{ \frac{\partial {P_p(z,t)}}{\partial {t}} } + { \frac{\partial {P_p(z,t)}}{\partial {z}} } = \sigma_{p} P_p(z,t) \left( N_5(z,t)-N_1(z,t) \right)-\alpha_p P_p(z,t)$$
$$\begin{aligned}& \frac{1}{v}{ \frac{\partial {P^{{\pm}}(z,t,\lambda_j)}}{\partial {t}} } \pm { \frac{\partial {P^{{\pm}}(z,t,\lambda_j)}}{\partial {z}} } = P^{{\pm}}(z,t,\lambda_j)\Big( \sigma_A(\lambda_j) \left( N_4(z,t)-N_2(z,t) \right) + \\ & \sigma_B(\lambda_j) \left( N_4(z,t)-N_3(z,t) \right) \Big) -\alpha(\lambda_j) P^{{\pm}}(z,t,\lambda_j) +\alpha_{RS} P^{{\mp}}(z,t,\lambda_j) + \frac{2hc^{2}}{\lambda_j^{3}} \Delta \lambda_j \frac{\Delta\Omega}{4\pi} \zeta(z,t,\lambda_j) , \end{aligned}$$
where the first terms on the right-hand sides stand for absorption and induced emission due to interaction with the population densities $N_i(z,t)$. The second terms represent effective losses, for example diffraction, with the coefficients $\alpha _p$ and $\alpha (\lambda _j)$, respectively, and $\alpha _{RS}$ is the Rayleigh backscattering coefficient (capture factor included). The final term on the right-hand side of Eq. (2b) accounts for spontaneous emission, where $2hc^{2}/\lambda _j^{3} \Delta \lambda _j = 2h\nu _j \Delta \nu$ is the contributed power of spontaneously emitted photons of 2 independent polarisations belonging to a mode $\lambda _j$ with a spectral spacing $\Delta \lambda _j$ between modes into a solid angle $\Delta \Omega$ where photons are kept inside the cavity [26]. Moreover, $\zeta (z,t,\lambda _j)$ is random in nature and describes the number of photons of wavelength $\lambda _j$ spontaneously emitted in the region $[z,z+\textrm{d}z)$ at time $t$ per length increment $\textrm{d}z$. The term will be further discussed in Section 2.3.

Note that this paper does not focus on modeling all possible effects occurring in Nd:YAG lasers, but rather on demonstrating, based on an accurate traveling-wave model, how to derive a simple set of equations (see Section 3) which describes the stochastically and dynamically influenced build-up of energy within the cavity. For a more detailed description, effects such as thermalization of state manifolds or excited state absorption can be additionally accounted for. In Section 3.2, we perform a thermalization analysis for the quasi-stationary case, while more elaborate approaches can be used similar to [27]. To model excited state absorption, as was done in [28], one can either introduce additional populations which are then pumped from the metastable state(s) or simply adapt the absorption loss coefficient $\alpha _p$ in Eq. (2a) to simulate the increase in threshold pump power and decrease in slope efficiency caused by excited state absorption.

2.2 Cavity boundary conditions

As can be seen in Fig. 1, optical elements (whose effective round-trip cavity efficiency is $\eta (\lambda _j)$) bring along that the pump and beam powers are subject to boundary conditions (B.C.)

$$P_p(0,t) = P_p(t)$$
$$P^{+}(0,t,\lambda_j) = \sqrt{\eta(\lambda_j)} P^{-}(0,t,\lambda_j)$$
$$P^{-}(L_{cav},t,\lambda_j) = \sqrt{\eta(\lambda_j)} P^{+}(L_{cav},t,\lambda_j) \ .$$
For $R(t)\neq 1$, the Pockels cell introduces a discontinuity at $z_{PC}$ given by
$$P^{-}(z_{PC}-\textrm{d}z ,t+ \textrm{d}z / v,\lambda_j) = R(t) P^{-}(z_{PC},t,\lambda_j) ,$$
with the resulting output power
$$P_{out}(t,\lambda_j) = \left(1-R(t)\right) P^{-}(z_{PC},t,\lambda_j) \ .$$
From this and the Q-switching frequency $f_{switch}$, we get the $m$-th output pulse energy as
$$E_{m} = \sum_{j=1}^{M} \int_{(m-1)/f_{switch}}^{m/f_{switch}} P_{out}(t,\lambda_j) \ \textrm{d}t \ .$$

2.3 Stochastic seeding via spontaneous emission

First, we define the total scattering cross-section

$$\sigma(\lambda_j):=\sigma_A(\lambda_j) + \sigma_B(\lambda_j) \ .$$
We will use it for the ensemble mean of $\zeta (z,t,\lambda _j)$,
$$\mathbb{E}(\zeta(z,t,\lambda_j)) = \sigma(\lambda_j) N_4(z,t) ,$$
which corresponds to the inverse mean free path. It can be shown quantum-mechanically for a highly simplified model [29], where for instance, all longitudinal modes experience identical gains, that amplified spontaneous emission in $K$ independent modes belonging to one bosonic excitation state follows the $K$-fold degenerate Bose-Einstein distribution, which for large $K$ approaches a Poisson distribution and for low $K$ is significantly broader [30]. Since we simulate every cavity mode individually, we can set $K=1$ and therefore propose that the probability of $n$ photons of wavelength $\lambda _j$ being spontaneously emitted within $[0,z]$ at time $t$ follows a single-mode Bose-Einstein distribution
$$\mathbb{P}[\mathfrak{C}(z,t,\lambda_j) = n] = \left( \frac{\mu(z,t,\lambda_j)}{1 + \mu(z,t,\lambda_j)} \right)^{n} \frac{1}{1 + \mu(z,t,\lambda_j)} ,$$
with mean
$$\mu(z,t,\lambda_j) := \sigma(\lambda_j)\int_0^{z} N_4(z',t) \ \textrm{d}z' \ .$$
Hence, the number of photon counts,
$$\mathfrak{C}(z,t,\lambda_j) := \int_0^{z} \zeta(z',t,\lambda_j) \ \textrm{d}z' ,$$
is an inhomogeneous jump process following the probability distribution Eq. (8).

The dominant mode $\lambda _j$ of a pulse arises from a competition among modes around the maximum of $\sigma _{A}(\lambda _j)$ or $\sigma _{B}(\lambda _j)$ due to the greatest amount of spontaneously emitted photons in the early phase of the energy build-up. Bragg gratings and other wavelength-selective elements inside the cavity can shape $\eta (\lambda _j)$ in the boundary conditions Eqs. (3a3c) such that the setup exhibits a few or many significant modes. As a consequence, the experimental output pulses will be approximately Bose-Einstein or Poisson distributed, as was shown in [30].

3. Model simplification

3.1 Transport equation solution and total rate equations

For control-related tasks, numerically solving the coupled rate Eqs.  (1a1d) and power transport Eqs.  (2a,2b) equations with the boundary conditions Eqs.  (3a4a) is not feasible as systematically investigating the pulse-to-pulse dynamics based solely on simulations is difficult. To eliminate spatial dependences, we introduce the total populations

$$N_i^{tot}(t):=\int_0^{L} N_i(z,t) \ \textrm{d}z$$
and make a number of simplifications which allow to express the dynamical behavior of the laser in terms of these new variables.

We start the model simplification by choosing frames $t \rightarrow t \mp z/v$ in Eqs.  (2a,2b) which move with a point of the pump or signal beam. Additionally, we make the following two assumptions:

  • (A1.) When the forward-running (the backward-running) laser beam travels through the active medium starting at $t_k$, the population densities and the lasing powers are assumed to be temporally constant, i.e. $N_i(z,t) = N_i(z,t_k)$ and $P^{\mp }(z,t) = P^{\mp }(z,t_k)$ for $t \in [t_k,t_k+L/v)$ in (2). Moreover, for the Rayleigh backscattering contribution $\alpha _{RS}P^{\mp }(z,t,\lambda _j)$, the dependence of $P^{\mp }$ on $z$ will be neglected, i.e. $P^{-}(z,t,\lambda _j)=P^{-}(L,t,\lambda _j)$ and $P^{+}(z,t,\lambda _j)=P^{+}(0,t,\lambda _j)$. The transport phenomena can therefore be described completely in terms of the forward- and backward-moving frames $t \rightarrow t \mp z/v$.
  • (A2.) Photons that are spontaneously emitted or backscattered at $z \in [0,L]$ join the forward-running (the backward-running) beam at $z=L$ ($z=0$). In the supplemental document, we provide the formulae representing this assumption.

With initial conditions given by the incident powers $P_p(0,t_k) = P_p(t_k)$ and $P^{\pm }(\frac {L\mp L}{2},t_k,\lambda _j)$ for the pump and signal beams before transmission, respectively, and separating each two channels’ frequencies by one inverse cavity round-trip time, i.e. $\Delta \nu = 1/t_{RT}$, we show in the supplemental document that the power transport Eq. (2) can be solved analytically provided that the assumptions (A1.) and (A2.) hold. The solutions, i.e. the outgoing powers corresponding to any incoming ones, read

$$P_p(L,t_k+L/v) = \exp \Big( \sigma_{p} (N_5^{tot}(t_k) - N_1^{tot}(t_k)) -\alpha_pL \Big) P_p(0,t_k)$$
$$\begin{aligned} &P^{{\pm}}\left(\frac{L\pm L}{2},t_k+L/v,\lambda_j\right) = 2 \frac{hc}{\lambda_j t_{RT}} \frac{\Delta \Omega}{4\pi} \mathfrak{C}(t_k,\lambda_j) + \alpha_{RS}L P^{{\mp}}\left(\frac{L\pm L}{2},t_k,\lambda_j\right) +\\ & \ \ \ \ \ \ \ \ \ \ \ \ P^{{\pm}}\left(\frac{L\mp L}{2},t_k,\lambda_j\right) \exp \Big( \sigma_A(\lambda_j) \left(N_4^{tot}(t_k) - N_2^{tot}(t_k)\right) +\\ & \ \ \ \ \ \ \ \ \ \ \ \ \sigma_B(\lambda_j) \left(N_4^{tot}(t_k) - N_3^{tot}(t_k)\right) -\alpha(\lambda_j)L \Big) , \end{aligned}$$
where the spontaneously emitted photon count $\mathfrak {C}(t_k,\lambda _j) := \mathfrak {C}(L,t_k,\lambda _j)$ is a random number with distribution (8) and mean $\mu (L,t_k,\lambda _j)$ according to Eq. (9), i.e. the value the counting process has reached between $z=0$ and $z=L$. As a next step, we want to find rate equations for the total populations defined by Eq. (11). To this end, we make use of the following assumptions:
  • (A3.) When calculating the depletion of population inversion due to lasing powers or its increase due to the pump power, dependence on $z$ of the loss terms will be neglected, i.e. $\alpha (\lambda _j)P^{+}(z,t,\lambda _j) = \alpha (\lambda _j)P^{+}(0,t,\lambda _j)$ and $\alpha (\lambda _j)P^{-}(z,t,\lambda _j) = \alpha (\lambda _j)P^{-}(L,t,\lambda _j)$ as well as $\alpha _p P_p(z,t) = \alpha _p P_p(0,t)$.
  • (A4.) For the induced emission contributions to the lower populations $N_3$ and $N_2$ in Eqs. (1a1d), spectral overlap of the lineshapes $\sigma _A(\lambda )$ and $\sigma _B(\lambda )$ can be disregarded.

Assumption (A4.) may be implemented using lineshape separator functions $g_A(\lambda )$ and $g_B(\lambda )$ which are either 1 or 0 and switch support at some separation wavelength $\lambda _{AB}$. The effect of lineshape separators can be seen in Fig. 3. With assumptions (A1.)–(A4.), we show in the supplemental document that the total rate equations read as

$$\begin{aligned} &{ \frac{\textrm{d} {N_5^{tot}(t)}}{\textrm{d} {t}} } ={-}\gamma_{54} N_5^{tot}(t) + \gamma_{45} N_4^{tot}(t) +\\ & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \frac{\lambda_{p}}{h c A_p} \bigg(1 - \exp \Big( \sigma_{p} (N_5^{tot}(t) - N_1^{tot}(t)) -\alpha_pL \Big) -\alpha_pL \bigg) P_p(t) \end{aligned}$$
$${ \frac{\textrm{d} {N_4^{tot}(t)}}{\textrm{d} {t}} } ={-} (\gamma_{45} + \gamma_{43} + \gamma_{42}) N_4^{tot}(t) + \gamma_{54} N_5^{tot}(t) - \sum_{j=1}^{M} G(t,\lambda_j)$$
$${ \frac{\textrm{d} {N_3^{tot}(t)}}{\textrm{d} {t}} } ={-} \gamma_{31} N_3^{tot}(t) + \gamma_{13} N_1^{tot}(t) + \gamma_{43} N_4^{tot}(t) + \sum_{j=1}^{M} g_B(\lambda_j) G(t,\lambda_j)$$
$${ \frac{\textrm{d} {N_2^{tot}(t)}}{\textrm{d} {t}} } ={-} \gamma_{21} N_2^{tot}(t) + \gamma_{12} N_1^{tot}(t) + \gamma_{42} N_4^{tot}(t) + \sum_{j=1}^{M} g_A(\lambda_j) G(t,\lambda_j)$$
$$\begin{aligned} &G(t,\lambda_j):= \frac{\lambda_j}{h c A_s} \left(P^{+}(0,t,\lambda_j)+P^{-}(L,t,\lambda_j) \right) \times\\ & \times \bigg( \exp \Big( \sigma_A(\lambda_j) \left(N_4^{tot}(t) - N_2^{tot}(t)\right) + \sigma_B(\lambda_j) \left(N_4^{tot}(t) - N_3^{tot}(t)\right) -\alpha(\lambda_j)L \Big) -1 + \alpha(\lambda_j)L \bigg) , \end{aligned}$$
with the algebraic equation $N_1^{tot}(t) = L N_{dop} - N_5^{tot}(t) - N_4^{tot}(t)- N_3^{tot}(t)- N_2^{tot}(t)$ .

 figure: Fig. 3.

Fig. 3. The overlap between lineshapes of scattering cross-sections $\sigma _A(\lambda )$ and $\sigma _B(\lambda )$ is eliminated by lineshape separators $g_A(\lambda )$ and $g_B(\lambda )$.

Download Full Size | PDF

3.2 State reduction via slow-fast dynamics

The model so far considers 5 populations of which only 4 are independent due to the algebraic condition involving $N_{dop}$. If the relaxation rate $\gamma _{ik}$ of an excitation state population $N_i$ is very large, the state is depopulated nearly instantly to $N_k$. In that case, which is usually a good approximation for low energy gaps $\Delta E_{ik}$ in the order of phonon energies, we can apply singular perturbation theory to reduce the order of the model. Assuming that the active medium is at thermal equilibrium (thus introducing a thermalization factor $B_{45} := \exp \left (\frac {-\Delta E_{45}}{k_BT}\right )$) [31] and employing a linear state transform, we show in the supplemental document that the total rate Eqs. (13a13d) consist of a slow and a fast dynamic subsystem. The limit of infinitely fast relaxation yields a quasi-stationary model for the parameters

$$\gamma_{31} = a_{31} \gamma_{54}, \ \gamma_{21} = a_{21} \gamma_{54}, \ a_{31} > 0, \ a_{21} > 0, \ \gamma_{54} \rightarrow \infty ,$$
which results in a slow dynamics that can be expressed solely in terms of the total population
$$N(t) := N_4^{tot}(t) \ .$$
In light of Eq. (13e), a reasonable choice of the spatially lumped intracavity power can be given by the power directly applying to the active medium, which is given by
$$P_j(t) := P^{+}(0,t,\lambda_j) + P^{-}(L,t,\lambda_j) \ .$$
In the limit Eq. (14) with Eqs. (15), (16) and (6) and defining $b:=1/(1+B_{45})$ as well as $\gamma :=\gamma _{42} + \gamma _{43}$, we show in the supplemental document that one can obtain from Eqs. (13a13d) the reduced total rate equation
$$\begin{aligned} { \frac{\textrm{d} { N(t)}}{\textrm{d} {t}} } & ={-} b \gamma N(t) - b \sum_{j=1}^{M}\frac{\lambda_j}{h c A_s} \bigg( \exp \Big(\sigma(\lambda_j) N(t) -\alpha(\lambda_j)L \Big) -1 + \alpha(\lambda_j)L \bigg) P_j(t)\\ & + b \frac{\lambda_{p}}{h c A_p} \bigg( 1- \exp \Big( \sigma_{p} \left((2 B_{45} + 1) N(t) -L N_{dop} \right) -\alpha_pL \Big) -\alpha_pL \bigg) P_p(t)\ . \end{aligned}$$

3.3 Continuous-time nonlinear model (CTNM)

Since we are mainly concerned with the evolution of the intracavity power rather than the transport effects inside the cavity, we will average the transmission through the active medium, the cavity losses and the Q-switch influence over half a cavity round-trip for a linear cavity, as was done in [23]. For this, the power transport solution Eq. (12b) and the boundary conditions Eqs. (3b,3c,4a) are utilized and for simplicity, $\sqrt {R(t)}$ is applied to both forward- and backward-running beams instead of $R(t)$ to the backward beam. We also assume that amplification and losses also apply to spontaneously emitted or backscattered photons. Then, half a cavity round-trip, i.e. a time increment of $\Delta t := t_{RT}/2 = L_{cav}/v$, affects the powers according to

$$\begin{aligned} &P^{{\pm}}\left(\frac{L\mp L}{2},t_{k+1},\lambda_j\right) =\sqrt{ \eta(\lambda_j) R(t_k)} \exp\left(\sigma(\lambda_j) N(t_k)-\alpha(\lambda_j)L\right) \times\\ & \times \left( P^{{\mp}}\left(\frac{L\pm L}{2},t_k,\lambda_j\right) + 2 \frac{hc}{\lambda_j t_{RT}} \frac{\Delta\Omega}{4\pi} \mathfrak{C}(t_k,\lambda_j) +\alpha_{RS} LP^{{\mp}}\left(\frac{L\pm L}{2},t_k,\lambda_j\right) \right) \ . \end{aligned}$$
The time derivative of Eq. (16) can be approximated by
$${ \frac{\textrm{d} {P_j(t)}}{\textrm{d} {t}} } \approx \frac{P_j(t)}{\Delta t} \left( \ln P_j(t_{k+1}) - \ln P_j(t_{k}) \right) ,$$
which with Eqs.  (16) and (18) and assuming that spontaneous emission and Rayleigh backscattering are small effects (i.e. $\ln (1+x) \approx x$) yields an ordinary differential equation which is derived in the supplemental document and reads as
$${ \frac{\textrm{d} {P_j(t)}}{\textrm{d} {t}} } = \frac{2\sigma(\lambda_j)}{t_{RT}} N(t) P_j(t) - \frac{P_j(t)}{\tau(t,\lambda_j)} + \frac{ 2}{t_{RT}} \Big( \frac{ 4 hc}{\lambda_j t_{RT}} \frac{\Delta\Omega}{4\pi} \mathfrak{C}(t,\lambda_j) +\alpha_{RS}L P_j(t) \Big) ,$$
where we have defined the loss rate
$$1/\tau(t,\lambda_j) := \frac{1}{t_{RT}} \left( 2\alpha(\lambda_j) L - \ln\left( \eta(\lambda_j) R(t)\right) \right) \ .$$
Again, $\mathfrak {C}(t,\lambda _j)$ is a Bose-Einstein noise term with expectation value $\mathbb {E}( \mathfrak {C}(t,\lambda _j)) = \sigma (\lambda _j) N(t)$ that can be understood in a Langevin sense. More formally, one could write Eq. (20) as a stochastic differential equation involving a renewal-reward process. We show in the supplemental document how the special case of a compound Poisson process with exponentially distributed waiting times and Poisson distributed jumps can be used to model the emission of photons at random times while producing equivalent means and variances after a time increment of $t_{RT}/2$ as the Bose-Einstein distribution from Eq. (8). Neglecting time delays and Rayleigh backscattering during half a round-trip, we show in the supplemental document that the output power, originally given by Eq. (4b), can be obtained from the spatially lumped intracavity power by
$$P_{out}(t,\lambda_j) = \frac{\left( 1-R(t) \right)}{R(t)} \left( \frac{P_j(t)\exp\left( \sigma(\lambda_j)N(t)-\alpha(\lambda_j)L \right) + 2 \frac{hc}{\lambda_j t_{RT}} \frac{\Delta\Omega}{4\pi} \mathfrak{C}(t,\lambda_j)}{\frac{1}{\sqrt{\eta(\lambda_j)}R(t)} + \exp\left( \sigma(\lambda_j)N(t)-\alpha(\lambda_j)L \right) } \right),$$
while the output energy is still given by Eq. (5).

Equation (20) and the reduced total rate Eq. (17) provide a very compact description of the system dynamics during a switching cycle and constitute what we will call the continuous-time nonlinear model (CTNM) of actively Q-switched lasers. In order to obtain the population and power after a switching cycle (and hence the pulse-to-pulse behavior of the system) from the CTNM, one would have to solve the coupled, nonlinear differential Eqs. (17) and (20), which is not possible without further approximations, as we will discuss in Section 3.5.

3.4 Comparison to existing models

It is an interesting observation that our modeling effort so far bridges the gap between spatially distributed traveling-wave modeling (where we started our investigation) and the more traditional point models which are spatially lumped. From the spatially lumped total rate Eq. (17), the rate equation used in traditional point models can be derived by isolating the pump and relaxation terms, additionally assuming losslessness and non-pumping-saturation, i.e. $\alpha _p =0$ and $LN_{dop} \gg N(t)$, whose solution yields Eq. (5) in [32], and assuming low single-transmission gains, i.e. $\exp (\sigma (\lambda _j) N(t) -\alpha L) \approx 1+(\sigma (\lambda _j) N(t) -\alpha L)$. This indicates that the extra terms in Eq. (17) are effects of the spatially distributed nature of the amplification process of the intracavity power as well as of the absorption of the pump power neglected by traditional point models. These modifications have a significant effect on the model’s behavior during a switching cycle and affect the pulse-to-pulse dynamics, especially for low to medium switching frequencies. Furthermore, the power equation used along with a point model rate equation in Degnan’s model [23], which has been used extensively for experimental laser optimisation [32], can be retrieved from Eq. (20) by neglecting the last terms accounting for spontaneous emission and backscattering and disregarding multi-mode behavior. To the best of the authors’ knowledge, the model proposed in this paper is the first one on Q-switched lasers where the stochastic nature of spontaneous emission is incorporated.

3.5 Pulse-to-pulse dynamics

As pointed out in Section 1, the output pulse energies of an actively Q-switched laser are subject to fluctuations due to stochastic and dynamic effects. The former mainly come from amplified spontaneous emission, while the latter are a result of unstable pulse-to-pulse dynamics due to a coupling of subsequent pulses. Since Q-switching (e.g. a Pockels cell represented by $R(t)$) operates in a cyclic manner, the pulse-to-pulse dynamics can be described by a discrete-time dynamic system where each switching cycle is a discrete event. If $t_m$ is the time immediately after the $(m-1)$-th pulse has been coupled out, one switching cycle evolves according to $t_{m+1} = t_m + 1/f_{switch}$. We use the $m$-th pulse energy $E_{m}$ given by Eqs. (5) and (21) and the abbreviation $N(t_m):=N_m$. While most generally, $N_m$ and $P_{j}(t_m)$ would both be dynamic variables, we can argue that $P_{j}(t_m) \approx 0$ for all $m$ and $j$ since all intracavity power has just been coupled out at those times, which reduces the dimension of the dynamical system from $1+M$ to $1$. With future control applications in mind, typical feedback quantities include the measured pulse energy, i.e. $E_m$ is the considered output quantity. As input quantity, we can choose the intracavity power $P_j(t_m+t_{pump}) =: u_{j,m}$ after the pumping phase and at the beginning of the $m$-th energy build-up. This input quantity is either given by spontaneous emission (uncontrolled case) or by a defined level from external seeding or controlled prelasing [13] (controlled case). Thus, defining the pulse-to-pulse and output maps by $f$ and $h$, respectively, the discrete dynamics are given by

$$N_{m+1} = f(N_m,u_{j,m})$$
$$E_m = h(N_m,u_{j,m}) \ .$$
These mappings are either found from numerical solutions of the general model discussed in Section 2, numerical solutions of Eqs. (17) and (20) with Eqs. (21) and (5) or an approximate analytic solution. Thus, in order to arrive at explicit expressions for Eq. (22), we will separate the switching cycle into its essential sub-processes and approximately solve Eqs. (17) and (20) for each sub-process, incorporating additional fitting parameters. The three essential sub-processes of a switching cycle are:
  • 1. pumping and population relaxation for a duration of $t_{pump}$,
  • 2. energy build-up during the duty cycle $t_{bu} = \textrm{DC}/f_{switch}$ (DC is the relative duty cycle), and
  • 3. coupling-out of the pulse with pulsewidth $t_{coup} \approx t_{close}$, where $t_{close}$ is the closing time of the cavity.
The composition of these solutions then constitutes our discrete-time dynamic model of the pulse-to-pulse dynamics. This semi-analytic procedure can be done in different ways, of which the authors explored a few. We present a detailed derivation in the supplemental document whose results are relatively simple and can be fitted effectively to adjust the dynamics to the observed behavior. What renders the procedure especially difficult is a combination of exponential growth underlying the energy build-up and nonlinearities in both Eqs. (17) and (20). The nonlinearities stemming from spatially distributed amplification and absorption are essential to accurately describe the dynamics, especially for low to medium switching frequencies. Therefore, linearisation or omission of loss terms immediately leads to substantial modeling errors. In the supplemental document, we show that one can find explicit expressions for the pulse-to-pulse dynamics of the form
$$N_{m+1} = f(N_m,u_{j,m}) = f_{3,N} \left( \mathbf{f}_2 \left( \left[ \begin{array}{c} f_1(N_m)\\\sum_{j=1}^{M} u_{j,m} / M \end{array} \right] \right) \right)$$
$$E_{m} = h(N_m,u_{j,m}) = f_{3,E} \left( \mathbf{f}_2 \left( \left[ \begin{array}{c} f_1(N_m)\\\sum_{j=1}^{M} u_{j,m} / M \end{array} \right] \right) \right) ,$$
where $f_1$ represents pumping, $\mathbf {f}_2$ the energy build-up and $f_{3,N}$ and $f_{3,E}$ the effect of coupling-out on inversion and pulse energy, respectively. Such explicit expressions provide an efficient tool for analyzing the resulting laser dynamics as well as designing real-time capable control and estimation algorithms.

4. Simulation results

In the following, we will show simulation results which demonstrate the interplay of stochastic and dynamic effects, agreement with first measurement results and the accuracy of the model simplifications proposed in the previous section.

The supplemental document contains the modeling parameters used in these simulations.

4.1 Pulse statistics and bifurcations

Apart from noise on the intracavity and output powers during a switching cycle, spontaneous emission also causes variations in the output energies even when the system is dynamically stable. To study the interplay of stochastic and dynamic effects in our general model discussed in Section 2. and given by (1–5), the duty cycle (DC) of the Pockels cell is varied at a fixed switching frequency ($f_{switch} =$ 1 MHz) and pump power ($P_p =$ 22 W). By plotting the output energy histograms and deterministic output pulses for comparison, a bifurcation diagram (Fig. 4) is obtained which displays how period-doubling bifurcations and increasing losses due to a longer build-up affect the energy distribution. Since a cavity design with large bandwidth was assumed, we have a large number of contributing modes and hence expect roughly Poissonian output energy statistics, as was pointed out in [30], which is indeed the case, cf. Fig. 5(b)). On top of the stochastic variations, period-doubling bifurcations at relative duty cycles of around 34% and 42% can be observed. From the third period-doubling onwards, we get deterministic chaos with an intermittent window of a 3-limit cycle between 53 and 56 %. The mean output energy diminishes at such build-up times since the amplification phase has no significant gain left and losses dominate instead.

 figure: Fig. 4.

Fig. 4. Distribution of output pulse energies of the full stochastic model (1–5) and its deterministic counterpart for varying relative duty cycles (DC). As the DC increases, losses lower the mean output energy and period-doubling bifurcations occur.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Comparison of the output energies $E_m$ of Eqs. (17),(20),(21),(5) to measured data and theoretical predictions [29]. a) shows the simulated steady-state energies $E_s$ and average energies from an experimental Yb:CaF2 laser. In b), the distribution of experimentally measured pulse energies is compared to simulations of an Nd:YAG laser. The bottom row compares the distribution of simulated normalized photon numbers Eq. (24) obtained from output energies of an Nd:YAG laser to c) a Poisson ($M=270$) and d) a Bose-Einstein distribution with $M=1$.

Download Full Size | PDF

4.2 Validity of the CTNM

The continuous-time nonlinear model (CTNM) given by Eqs. (17),(20) together with Eqs. (21),(5) simplifies the spatially distributed full model (1–5) by averaging all processes within the cavity over half a round-trip such that ordinary differential equations remain to describe the whole system dynamics. We will show in Section 4.3 (cf. Fig. 6) that this introduces only small model errors as the full model and the CTNM agree well both in stable and unstable regimes of operation. Figure 5 demonstrates that the CTNM can reproduce experimental data in terms of dependence on $P_p$ and the DC as well as theoretical predictions regarding the photon statistics of output pulses. The measured average pulse energies shown in a) stem from a double-pass diode-pumped Yb:CaF2 laser system at 200 kHz switching frequency and build-up times $t_{bu} = \textrm{DC}/f_{switch}$ of 1.4 $\mu$s and 1.5 $\mu$s, respectively. The corresponding simulation data represent steady-state output energies $E_s$ of (22) and were obtained using Eqs. (17),(20),(21),(5) without stochastic effects, with technical parameters adapted to the experiments and physical parameters similar to those from the literature (for reference, see the supplemental document for the Nd:YAG laser simulation parameters). One can see that the model fits the measured data very well both for different pump powers and for different build-up times.

 figure: Fig. 6.

Fig. 6. Collection of simulation results of full model ((1–5), blue), continuous-time nonlinear model (CTNM, Eqs. (17),(20),(21),(5), red), single-mode CTNM (yellow) and the discrete-time pulse-to-pulse model ((23), green). The two columns a) and b) correspond to a dynamically stable and unstable scenario. The rows display 1) total population inversions, 2) pulse energiese and 3),4) normalised pulse-to-pulse dynamic maps for populations and output energies.

Download Full Size | PDF

To compare the stochastic distributions of output pulse energies, we focused on a pulse-to-pulse-dynamically stable, non-saturating regime where dynamics effects do not interfere with the stochastics of amplified spontaneous emission. The bottom row of Fig. 5 shows the distribution of the normalized photon numbers

$$n := \frac{E_m \lambda}{M hc }$$
obtained from stochastic pulse energies $E_m$ given by Eq. (5), similarly to how it was done in [30] with experimental data. The histograms come from simulations of an Nd:YAG system at 1 MHz switching frequency and 20% DC exhibiting stable pulse-to-pulse dynamics, where the cavity supports around 270 contributing longitudinal modes in c) and a single mode in d). For comparison, the red lines in c) and d) of Fig. 5 represent a Poisson and a Bose-Einstein distribution and hence the theoretical predictions from [29], respectively. It should also be noted here that the CTNM and the full model exhibit equivalent pulse energy statistics. It can be seen that the simulated energy histograms and the theoretically predicted distributions, which were used in Section 2.3 to design our model of spontaneous emission, show excellent agreement in both cases.

In b), a simulation of an Nd:YAG system at 1 MHz switching frequency and 6% DC is compared to experimental pulse energies. Here, our simulation exhibits slightly larger stochastic variations than the experimental observation. This is plausible since the theoretical prediction made significant approximations, such as assuming stochastic independence between the photon numbers of lasing modes to arrive at the $M$-fold degenerate Bose-Einstein distribution, which are expected to broaden the pulse spectrum.

4.3 Comparison of models

Since we have shown in Section 4.2 that the CTNM Eqs. (17),(20),(21),(5) is able to reproduce the slope efficiencies of experimental lasers and the photon statistics of output pulses at least qualitatively, we will now assess whether the CTNM differs significantly from the full model (1–5) and how further simplifications compromise the model’s accuracy by analysing a collection of simulation results shown in Fig. 6. One simplification lies in reducing the system to single-mode operation, neglecting all other cavity modes in Eqs. (17),(20) besides the dominating one (i.e. the mode subject to the highest gain). Another simplification of the CTNM is its approximate analytic solution, i.e. the discrete-time model (23) of the pulse-to-pulse dynamics. Figure 6 contains two exemplary cases for a system at $f_{switch}$ = 1MHz repetition rate and a pump power $P_p$ = 22.5 W exhibiting stable (a: 8% DC) and unstable (b: 35% DC) pulse-to-pulse dynamics. They are simulated using the full model (1–5), the CTNM Eqs. (17),(20),(21),(5), a single-mode version of the CTNM and the discrete-time pulse-to-pulse model (23) discussed in Section 3.5. Here, $N_s$, $u_s$ and $E_s$ stand for the steady-state population, uncontrolled intracavity power and pulse energy of the pulse-to-pulse dynamics (22). Row 3) of Fig. 6 demonstrates that increasing the DC value alters the shape of the pulse-to-pulse dynamics curve $f(N_m,u_s)/N_s$. Increasing one of the other two bifurcation parameters $f_{switch}$ or $P_p$ moves the steady state (position of the grey arrow) of the pulse-to-pulse dynamics further to the right of the dynamical map $f(N_m,u_s)$ (to higher $N_m$) such that eventually the absolute value of the slope becomes larger than one and the dynamics hence unstable. In Fig. 6, we can see that the single-mode approximation (yellow) introduces major modeling errors as all photons are emitted into the channel with maximum gain. This in turn leads to a premature energy build-up with enormous depletion and power losses. One can further see – most distinctly in the insets of Fig. 6(1a) and (1b) – that the population inversions of the full model and the CTNM differ only by about 1 percent. The discrete-time model matches these dynamics relatively well, but in the instability region of the pulse-to-pulse dynamics, the limit cycles for populations in Fig. 6(1b) as well as pulse energies in Fig. 6(2b) have smaller amplitudes. In terms of pulse energies $E_m$, row 2) again displays a good agreement between the CTNM and the full model. The normalised pulse-to-pulse dynamic maps in rows 3) and 4) of Fig. 6 for the full model, the CTNM and the discrete-time model stay in close proximity. However, it should be emphasized that the discrete-time model fits the dynamics well up to around 1.1 times the steady-state population $N_s$ in rows 3) and 4). At larger inversion, too much population is depleted in the discrete-time model, and the output map $h(N_m,u_s)$ overshoots significantly. This is due to the fact that losses can only be incorporated via operator splitting, as we showed in the supplemental document. In practice, the overshoot is not problematic at all as such high inversions are never reached if the cavity is being opened and closed periodically, but an analysis on a broad domain of inversion values is still useful for assessing the global dynamical behavior of the models.

5. Conclusion and outlook

In this work, a computationally efficient mathematical model for an actively Q-switched laser system including the effects of stochastic spontaneous emission is derived as a basis for the design of advanced control and estimation strategies. A spatially distributed multi-mode traveling-wave model, which is commonly used in the literature, serves as a starting point. Based on this model, the power transport equations were simplified by making physically feasible assumptions and the singular perturbation theory was applied to systematically reduce the number of relevant populations. Then all the processes within the cavity were averaged over half a cavity round-trip. As a result, a set of coupled, nonlinear ordinary differential equations were obtained, which describe the evolution of the inversion and each cavity laser mode at any point in time. Moreover, in order to separately study dynamic and stochastic effects on energy fluctuations, a discrete-time model of the pulse-to-pulse dynamics was derived by solving the differential equations in a semi-analytic way. Simulation results show that the model simplification steps only introduce small errors in the range of sub-percent, except for the discrete-time model at large populations, which does not constitute an issue for the practical application. In particular, the systematic incorporation of multi-mode behavior and spatially distributed amplification, which distinguishes the spatially lumped continuous-time nonlinear model presented in this work from traditional point models known from the literature, provides superior modeling accuracy with respect to the full model at moderate computational costs. Thus, the continuous-time nonlinear model accurately describes the behavior of actively Q-switched lasers during a switching cycle, while the presented discrete-time model provides a simple relation between the initial populations and the corresponding output energies of two different cycles. The simulation results show that stochastic seeding leads to amplified spontaneous emission noise whose pulse energy statistics correspond to theoretical results from the literature which the stochastic modeling was based on. A pulse energy distribution from simulation is compared to an experimentally obtained one. The simulated slope efficiencies of output energies for various build-up times agree with first measurement results presented in the paper.

Moreover, as it was expected from experimental observations and other papers from the literature, under certain conditions the nonlinear pulse-to-pulse dynamics exhibits period-doubling bifurcations and deterministic chaos. The model presented in this paper aims at laying the foundation to systematically design control and estimation strategies to achieve robustly stable Q-switched lasers at high repetition rates. In particular, the pulse-to-pulse dynamics can be stabilised or otherwise modified via active feedback methods and prelasing strategies can be employed to suppress stochastic seeding due to spontaneous emission.

Funding

Österreichische Forschungsförderungsgesellschaft (884176).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper will be made available upon request by the authors.

Supplemental document

See Supplement 1 for supporting content.

References

1. H. Wang, Q. Wen, X. Xu, J. Lu, F. Jiang, and C. Cui, “Ablation characteristics and material removal mechanisms of a single-crystal diamond processed by nanosecond or picosecond lasers,” Opt. Express 29(14), 22714–22731 (2021). [CrossRef]  

2. X. Li, Q. Gao, L. Han, M. Li, Z. Zhu, and B. Li, “The spatial resolution of nanosecond laser-induced plasma spectroscopy in gases,” J. Anal. At. Spectrom. 36(5), 993–998 (2021). [CrossRef]  

3. B. Wu, “High-intensity nanosecond-pulsed laser-induced plasma in air, water, and vacuum: A comparative study of the early-stage evolution using a physics-based predictive model,” Appl. Phys. Lett. 93(10), 101104 (2008). [CrossRef]  

4. H. Chichan, M. Maus, and L. Heindl, “Subthreshold nanosecond laser, from trials to real-life clinical practice: A cohort study,” Clin. Ophthalmol. 15, 1887–1895 (2021). [CrossRef]  

5. J. Glaser, “High power nanosecond pulse laser driver using an gan fet,” in PCIM Europe 2018; International Exhibition and Conference for Power Electronics, Intelligent Motion, Renewable Energy and Energy Management, (2018), pp. 1–8.

6. S. Ma, H. Yu, and H. Zhang, “Efficient high repetition rate electro-optic q-switched laser with an optically active langasite crystal,” Sci. Rep. 6(1), 30517–30524 (2016). [CrossRef]  

7. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995).

8. S. A. Kolpakov, S. Sergeyev, C. Mou, N. T. Gordon, and K. Zhou, “Optimization of erbium-doped actively q-switched fiber laser implemented in symmetric configuration,” IEEE J. Sel. Top. Quantum Electron. 20(5), 329–336 (2014). [CrossRef]  

9. A. Deutschmann, T. Flöry, K. Schrom, V. Stummer, A. Baltuška, and A. Kugi, “Bifurcation suppression in regenerative amplifiers by active feedback methods,” Opt. Express 28(2), 1722–1737 (2020). [CrossRef]  

10. Y. Wang and C.-Q. Xu, “Switching-induced perturbation and influence on actively q-switched fiber lasers,” IEEE J. Quantum Electron. 40(11), 1583–1596 (2004). [CrossRef]  

11. Y. Barmenkov, A. Kir’yanov, and M. Andres, “Nonlinear dynamics of ytterbium-doped fiber laser q-switched using acousto-optical modulator,” Eur. Phys. J. Spec. Top. 223(13), 2775–2788 (2014). [CrossRef]  

12. Y. Barmenkov, A. Kir’yanov, and M. Andres, “Experimental Study of the Nonlinear Dynamics of an Actively Q-Switched Ytterbium-Doped Fiber Laser,” IEEE J. Quantum Electron. 48(11), 1484–1493 (2012). [CrossRef]  

13. T. Lian, S. Wang, D. Cai, B. Li, and Z. Guo, “High repetition rate, high peak power, pulsed single-longitudinal-mode nd:yag laser by self-injection-seeding,” Opt. Commun. 323, 154–161 (2014). [CrossRef]  

14. C. Fries, M. Weitz, C. Theobald, P. v. Löwis of Menar, J. Bartschke, and J. A. L’huillier, “Cavity-dumped yb:yag ceramic in the 20 w, 12 mj range at 6.7 ns operating from 20 hz to 5 khz with fluorescence feedback control,” Appl. Opt. 55(24), 6538–6546 (2016). [CrossRef]  

15. Y. Wang and C.-Q. Xu, “Actively q-switched fiber lasers: Switching dynamics and nonlinear processes,” Prog. Quantum Electron. 31(3-5), 131–216 (2007). [CrossRef]  

16. K. Yang, S. Zhao, G. Li, and H. Zhao, “A new model of laser-diode end-pumped actively q-switched intracavity frequency doubling laser,” IEEE J. Quantum Electron. 40(9), 1252–1257 (2004). [CrossRef]  

17. D. Stone, “Effects of axial nonuniformity in modeling q-switched lasers,” IEEE J. Quantum Electron. 28(10), 1970–1973 (1992). [CrossRef]  

18. M. Wohlmuth, C. Pflaum, K. Altmann, M. Paster, and C. Hahn, “Dynamic multimode analysis of q-switched solid state laser cavities,” Opt. Express 17(20), 17303–17316 (2009). [CrossRef]  

19. X. Zhang, Z. Wang, S. Liu, S. Gou, R. Fan, D. Jin, Z. Bai, and Z. Bai, “Development of single-longitudinal-mode selection technology for solid-state lasers,” Int. J. Opt. 2021, 1–10 (2021). [CrossRef]  

20. J. Lee, I. Hong, and J. H. Lee, “Numerical study on an optimum q-switching profile for complete multipeak suppression in an actively q-switched ytterbium fibre laser,” Laser Phys. Lett. 18(8), 085101 (2021). [CrossRef]  

21. S. Jafari, D. Fathi, and H. Taleb, “Optimal pulse-shaping in actively q-switched ytterbium-doped fiber lasers,” IEEE Access 8, 77716–77724 (2020). [CrossRef]  

22. M. Srivastava, D. Venkitesh, and B. Srinivasan, “Effects of wavelength filtering on pulse dynamics in a tunable, actively q-switched fiber laser,” Opt. Laser Technol. 98, 190–197 (2018). [CrossRef]  

23. J. Degnan, “Theory of the optimally coupled q-switched laser,” IEEE J. Quantum Electron. 25(2), 214–220 (1989). [CrossRef]  

24. D. Krennrich, R. Knappe, B. Henrich, R. E. Wallenstein, and J. A. L’huillier, “A comprehensive study of nd:yag, nd:yalo3, nd:yvo4 and nd:ygdvo4 lasers operating at wavelengths of 0.9 and 1.3 µm. part 1: cw-operation,” Appl. Phys. B 92(2), 165–174 (2008). [CrossRef]  

25. J. Verdeyen, Laser Electronics (Third Edition) (Pearson, 1995).

26. C. Yongqian, Z. Guangzhi, C. Hantian, W. M C. Keming, W. Hailin, A. Kozlov, and Z. Xiao, “Evaluation of amplified spontaneous emission in thin disk lasers using the spectral linewidth,” Opt. Express 27(9), 12110–12125 (2019). [CrossRef]  

27. J. Degnan, D. Coyle, and R. Kay, “Effects of thermalization on q-switched laser properties,” IEEE J. Quantum Electron. 34(5), 887–899 (1998). [CrossRef]  

28. P. R. Morkel and R. I. Laming, “Theoretical modeling of erbium-doped fiber amplifiers with excited-state absorption,” Opt. Lett. 14(19), 1062–1064 (1989). [CrossRef]  

29. A. Mecozzi, “Quantum and semiclassical theory of noise in optical transmission lines employing in-line erbium amplifiers,” J. Opt. Soc. Am. B 17(4), 607–617 (2000). [CrossRef]  

30. J. A. Minguela-Gallardo, Y. O. Barmenkov, A. V. Kir’yanov, and G. Beltrán-Pérez, “Photon statistics of actively q-switched erbium-doped fiber laser,” J. Opt. Soc. Am. B 34(7), 1407–1414 (2017). [CrossRef]  

31. M. W. P. Strandberg and J. R. Shane, “General properties of thermal-relaxation rate equations,” Phys. Rev. B 7(11), 4809–4812 (1973). [CrossRef]  

32. Y.-F. Chen, Y. P. Lan, and S. Tsai, “High-power diode-pumped actively q-switched nd:yag laser at 1123 nm,” Opt. Commun. 234(1-6), 309–313 (2004). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       zip file including supplemental document, bibliography and style files

Data availability

Data underlying the results presented in this paper will be made available upon request by the authors.

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. A sketch of the cavity, active medium (red area), Q-switch (Pockels cell, yellow area), forward pump power (blue) and forward- and backwards-running signal powers (green). The coupling of the two beam propagation directions as well as all other optical components are accounted for via the boundary conditions B.C. (green dots).
Fig. 2.
Fig. 2. The five energy occupation levels that constitute our model and their transitions: pumping (blue) and two lasing transitions (red, green). The black arrows stand for nonradiative thermal transitions.
Fig. 3.
Fig. 3. The overlap between lineshapes of scattering cross-sections $\sigma _A(\lambda )$ and $\sigma _B(\lambda )$ is eliminated by lineshape separators $g_A(\lambda )$ and $g_B(\lambda )$.
Fig. 4.
Fig. 4. Distribution of output pulse energies of the full stochastic model (1–5) and its deterministic counterpart for varying relative duty cycles (DC). As the DC increases, losses lower the mean output energy and period-doubling bifurcations occur.
Fig. 5.
Fig. 5. Comparison of the output energies $E_m$ of Eqs. (17),(20),(21),(5) to measured data and theoretical predictions [29]. a) shows the simulated steady-state energies $E_s$ and average energies from an experimental Yb:CaF2 laser. In b), the distribution of experimentally measured pulse energies is compared to simulations of an Nd:YAG laser. The bottom row compares the distribution of simulated normalized photon numbers Eq. (24) obtained from output energies of an Nd:YAG laser to c) a Poisson ($M=270$) and d) a Bose-Einstein distribution with $M=1$.
Fig. 6.
Fig. 6. Collection of simulation results of full model ((1–5), blue), continuous-time nonlinear model (CTNM, Eqs. (17),(20),(21),(5), red), single-mode CTNM (yellow) and the discrete-time pulse-to-pulse model ((23), green). The two columns a) and b) correspond to a dynamically stable and unstable scenario. The rows display 1) total population inversions, 2) pulse energiese and 3),4) normalised pulse-to-pulse dynamic maps for populations and output energies.

Equations (39)

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

N 5 ( z , t ) t = γ 54 N 5 ( z , t ) + γ 45 N 4 ( z , t ) + λ p σ p h c A p P p ( z , t ) ( N 1 ( z , t ) N 5 ( z , t ) )
N 4 ( z , t ) t = γ 54 N 5 ( z , t ) γ 45 N 4 ( z , t ) ( γ 42 + γ 43 ) N 4 ( z , t ) j = 1 M λ j h c A s × × ( P + ( z , t , λ j ) + P ( z , t , λ j ) ) ( σ A ( λ j ) ( N 4 ( z , t ) N 2 ( z , t ) ) + σ B ( λ j ) ( N 4 ( z , t ) N 3 ( z , t ) ) )
N 3 ( z , t ) t = γ 31 N 3 ( z , t ) + γ 13 N 1 ( z , t ) + γ 43 N 4 ( z , t ) + j = 1 M λ j h c A s ( P + ( z , t , λ j ) + P ( z , t , λ j ) ) ( σ B ( λ j ) ( N 4 ( z , t ) N 3 ( z , t ) ) )
N 2 ( z , t ) t = γ 21 N 2 ( z , t ) + γ 12 N 1 ( z , t ) + γ 42 N 4 ( z , t ) + j = 1 M λ j h c A s ( P + ( z , t , λ j ) + P ( z , t , λ j ) ) ( σ A ( λ j ) ( N 4 ( z , t ) N 2 ( z , t ) ) ) ,
1 v P p ( z , t ) t + P p ( z , t ) z = σ p P p ( z , t ) ( N 5 ( z , t ) N 1 ( z , t ) ) α p P p ( z , t )
1 v P ± ( z , t , λ j ) t ± P ± ( z , t , λ j ) z = P ± ( z , t , λ j ) ( σ A ( λ j ) ( N 4 ( z , t ) N 2 ( z , t ) ) + σ B ( λ j ) ( N 4 ( z , t ) N 3 ( z , t ) ) ) α ( λ j ) P ± ( z , t , λ j ) + α R S P ( z , t , λ j ) + 2 h c 2 λ j 3 Δ λ j Δ Ω 4 π ζ ( z , t , λ j ) ,
P p ( 0 , t ) = P p ( t )
P + ( 0 , t , λ j ) = η ( λ j ) P ( 0 , t , λ j )
P ( L c a v , t , λ j ) = η ( λ j ) P + ( L c a v , t , λ j )   .
P ( z P C d z , t + d z / v , λ j ) = R ( t ) P ( z P C , t , λ j ) ,
P o u t ( t , λ j ) = ( 1 R ( t ) ) P ( z P C , t , λ j )   .
E m = j = 1 M ( m 1 ) / f s w i t c h m / f s w i t c h P o u t ( t , λ j )   d t   .
σ ( λ j ) := σ A ( λ j ) + σ B ( λ j )   .
E ( ζ ( z , t , λ j ) ) = σ ( λ j ) N 4 ( z , t ) ,
P [ C ( z , t , λ j ) = n ] = ( μ ( z , t , λ j ) 1 + μ ( z , t , λ j ) ) n 1 1 + μ ( z , t , λ j ) ,
μ ( z , t , λ j ) := σ ( λ j ) 0 z N 4 ( z , t )   d z   .
C ( z , t , λ j ) := 0 z ζ ( z , t , λ j )   d z ,
N i t o t ( t ) := 0 L N i ( z , t )   d z
P p ( L , t k + L / v ) = exp ( σ p ( N 5 t o t ( t k ) N 1 t o t ( t k ) ) α p L ) P p ( 0 , t k )
P ± ( L ± L 2 , t k + L / v , λ j ) = 2 h c λ j t R T Δ Ω 4 π C ( t k , λ j ) + α R S L P ( L ± L 2 , t k , λ j ) +                         P ± ( L L 2 , t k , λ j ) exp ( σ A ( λ j ) ( N 4 t o t ( t k ) N 2 t o t ( t k ) ) +                         σ B ( λ j ) ( N 4 t o t ( t k ) N 3 t o t ( t k ) ) α ( λ j ) L ) ,
d N 5 t o t ( t ) d t = γ 54 N 5 t o t ( t ) + γ 45 N 4 t o t ( t ) +                                       λ p h c A p ( 1 exp ( σ p ( N 5 t o t ( t ) N 1 t o t ( t ) ) α p L ) α p L ) P p ( t )
d N 4 t o t ( t ) d t = ( γ 45 + γ 43 + γ 42 ) N 4 t o t ( t ) + γ 54 N 5 t o t ( t ) j = 1 M G ( t , λ j )
d N 3 t o t ( t ) d t = γ 31 N 3 t o t ( t ) + γ 13 N 1 t o t ( t ) + γ 43 N 4 t o t ( t ) + j = 1 M g B ( λ j ) G ( t , λ j )
d N 2 t o t ( t ) d t = γ 21 N 2 t o t ( t ) + γ 12 N 1 t o t ( t ) + γ 42 N 4 t o t ( t ) + j = 1 M g A ( λ j ) G ( t , λ j )
G ( t , λ j ) := λ j h c A s ( P + ( 0 , t , λ j ) + P ( L , t , λ j ) ) × × ( exp ( σ A ( λ j ) ( N 4 t o t ( t ) N 2 t o t ( t ) ) + σ B ( λ j ) ( N 4 t o t ( t ) N 3 t o t ( t ) ) α ( λ j ) L ) 1 + α ( λ j ) L ) ,
γ 31 = a 31 γ 54 ,   γ 21 = a 21 γ 54 ,   a 31 > 0 ,   a 21 > 0 ,   γ 54 ,
N ( t ) := N 4 t o t ( t )   .
P j ( t ) := P + ( 0 , t , λ j ) + P ( L , t , λ j )   .
d N ( t ) d t = b γ N ( t ) b j = 1 M λ j h c A s ( exp ( σ ( λ j ) N ( t ) α ( λ j ) L ) 1 + α ( λ j ) L ) P j ( t ) + b λ p h c A p ( 1 exp ( σ p ( ( 2 B 45 + 1 ) N ( t ) L N d o p ) α p L ) α p L ) P p ( t )   .
P ± ( L L 2 , t k + 1 , λ j ) = η ( λ j ) R ( t k ) exp ( σ ( λ j ) N ( t k ) α ( λ j ) L ) × × ( P ( L ± L 2 , t k , λ j ) + 2 h c λ j t R T Δ Ω 4 π C ( t k , λ j ) + α R S L P ( L ± L 2 , t k , λ j ) )   .
d P j ( t ) d t P j ( t ) Δ t ( ln P j ( t k + 1 ) ln P j ( t k ) ) ,
d P j ( t ) d t = 2 σ ( λ j ) t R T N ( t ) P j ( t ) P j ( t ) τ ( t , λ j ) + 2 t R T ( 4 h c λ j t R T Δ Ω 4 π C ( t , λ j ) + α R S L P j ( t ) ) ,
1 / τ ( t , λ j ) := 1 t R T ( 2 α ( λ j ) L ln ( η ( λ j ) R ( t ) ) )   .
P o u t ( t , λ j ) = ( 1 R ( t ) ) R ( t ) ( P j ( t ) exp ( σ ( λ j ) N ( t ) α ( λ j ) L ) + 2 h c λ j t R T Δ Ω 4 π C ( t , λ j ) 1 η ( λ j ) R ( t ) + exp ( σ ( λ j ) N ( t ) α ( λ j ) L ) ) ,
N m + 1 = f ( N m , u j , m )
E m = h ( N m , u j , m )   .
N m + 1 = f ( N m , u j , m ) = f 3 , N ( f 2 ( [ f 1 ( N m ) j = 1 M u j , m / M ] ) )
E m = h ( N m , u j , m ) = f 3 , E ( f 2 ( [ f 1 ( N m ) j = 1 M u j , m / M ] ) ) ,
n := E m λ M h c
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.