\documentstyle[aps,preprint]{revtex} \def\baselinestretch{1.2} \textheight 8.5in \topmargin .0in \textwidth 6in \oddsidemargin .25in \evensidemargin 0in \parskip .25in \pagestyle{myheadings} \begin{document} \baselineskip 24pt \title{Travelling Wave Model of a Multimode Fabry-P\'{e}rot Laser\\ in Free Running and External Cavity Configurations} \author{M. Homar$^{1,2}$, J. V. Moloney$^{1}$ and M. San Miguel$^{1,2}$ \\ $^{1}$ Arizona Center for Mathematical Sciences, University of Arizona \\ $^{2}$ Departament de F\'{\i}sica, Universitat de les Illes Balears\\ E--07071 Palma de Mallorca, Spain\\ } \date{~} \maketitle \begin{abstract} We report the results of a numerical study of multimode behavior of a Fabry-P\'{e}rot laser. The model is based on travelling-wave equations for the slowly varying amplitudes of the counterpropagating waves in the cavity, coupled to equations for spatially dependent population inversion and polarization of a two-level active medium. Variations in the material variables on the scale of a wavelength are taken into account by means of an expansion in a Fourier series. Results are given for typical semiconductor laser parameters. Spatially distributed spontaneous emission noise and carrier diffusion are taken into account. The competing roles of Spatial Hole Burning (SHB), spontaneous emission noise and carrier diffusion in determining multimode behavior is elucidated: With no carrier diffusion, spontaneous emission noise excites a large number of modes close to threshold, while SHB leads to a fixed number of significant lasing modes well above threshold. Carrier diffusion washes out the gratings in the material variables, and the resulting strengthening of the inter-mode coupling (cross-saturation) restores dominant single-mode emission well above threshold. We have also studied the effects of optical feedback and opportunities for mode selection with short external cavities: For an external cavity much shorter than the laser cavity length and a small field amplitude reflectivity coefficient a single mode can be selected. For a large reflectivity coefficient two groups of intracavity modes separated by the external cavity mode interspacing are selected. For an external cavity with a round trip time half that of the laser cavity, the laser can be forced with modest feedback to operate on two modes which are both quasi-resonant with the external cavity. Mode selection is not found, even for weak feedback, when the external mode spacing is about $90 \%$ of the laser mode spacing. \end{abstract} \newpage \section{Introduction} The optical spectrum of laser light usually includes several frequencies which are associated with longitudinal cavity modes. Multimode emission degrades the spectral purity required in some laser applications, but allows a higher output power required in other applications. Longitudinal multimode laser behavior is well known in solid state \cite{baer,TSM,khanin,roy,glorieux}, dye\cite{raymer1,raymer2}, semiconductor \cite{agrawal,lee,mandel,yamamoto} and gas lasers\cite{refneal}. In principle, the number of possible lasing modes can be estimated by counting the cavity modes that lie in the spectral region where the unsaturated gain exceeds the loss. However, multimode laser emission involves strong nonlinear mode interaction and competition for the gain. This may lead to mode locking or to a random partition of the total power among modes which oscillate simultaneously (anticorrelated fluctuations), with the extreme behavior being mode hopping. The number of lasing modes with significant power depends on the operating conditions of the laser, and it is determined by several physical mechanisms such as spontaneous emission noise, Spatial Hole Burning (SHB), phase sensitive interactions such as Four Wave Mixing (FWM) and diffusion processes. A conventional theoretical description of multimode laser behavior is based on an expansion of the laser field in a set of normal standing-wave modes appropriate to the laser cavity \cite{sigman}. In this paper we follow a different approach analyzing a Traveling Wave (TW) model for the laser field in a two level active medium filling a Fabry-P\'{e}rot (FP) cavity. The partially reflecting cavity boundary conditions are incorporated and the mode decomposition is avoided. Within this description we discuss the role of different mechanisms in multimode behavior at different pumping levels. A popular description of multimode behavior in semiconductor and solid state lasers features rate equations \cite{agrawal} for the photon numbers $I_i$ associated with the different lasing modes $i$. The variables $I_i$ are nonlinearly coupled among themselves and with a global population inversion variable $N$. In this description one introduces as many gain parameters, for the different modes, as a fixed number of modes are selected {\it a priori} to be relevant. Multimode rate equations, often used for semiconductor lasers \cite{agrawal,lee,yamamoto,gray,ohtsu}, do not incorporate some important physical phenomena such as phase dynamics, phase-amplitude coupling and SHB. The effect of carrier diffusion and SHB in semiconductor lasers was considered in the context of rate equations by phenomenologicaly introducing effective gain parameters for the modes \cite{lee}. A more elaborated approach than standard rate equations has been used to study homogeneously broadened two-level laser models: Studies of multimode solid-state lasers \cite{baer,TSM,glorieux} are based on a set of equations for the $I_i$ which incorporates, as additional dynamical variables, a gain variable associated with each mode $i$. These additional gain variables are used to describe the gratings in the inversion from the longitudinally inhomogeneous gain saturation caused by the optical standing waves associated with each lasing mode in the cavity. A related modeling has been used for dye lasers \cite {raymer1,raymer2}, and justified from Maxwell-Bloch equations by a mode expansion for the lasing field and an adiabatic elimination of the polarization variables. Coupled equations for the complex field mode amplitudes and two spatially independent Fourier amplitudes of the population inversion are obtained in a thin medium approximation which includes effects such as SHB. The relevance of phase sensitive mode-mode interaction for semiconductor lasers has been emphasized by Mandel et al. \cite{mandel} in an analysis similar to the one in \cite {raymer1,raymer2}. They also take as starting point Maxwell-Bloch equations for a two-level medium interacting with a number of cavity modes of a laser field. Carrier diffusion is neglected. Through an adiabatic elimination of the polarization variables they obtain coupled equations for the complex amplitudes of the laser modes and population variables which include phase dependent couplings. Phase sensitive interactions have also been discussed in the context of cubic equations for coupled mode amplitudes for gas lasers \cite{lugiato}. All the multimode lasing descriptions mentioned above have in common that they are based on a decomposition of the laser field into cavity modes and in neglecting the fast polarization dynamics. In addition, if the spatial dependence of population inversion is considered, it is assumed to have the same Fourier modes as that of the laser field. A different description of the laser field, which we pursue here, is based upon a direct solution of the wave equation for the field with reflecting boundary conditions appropriate to the laser cavity \cite{fleck,n-m,wfjm}. Such a treatment avoids explicit assumptions regarding the longitudinal structure of the laser. The traveling wave equation for the field is coupled to Bloch equations for population inversion and polarization variables which depend on the spatial coordinate along the cavity. We note that a simple adiabatic elimination of the fast polarization dynamics is not possible within the TW model, since it leads to ill-behaved equations. This fact is similar to the unphysical growth of large wavenumber perturbations found when studying transverse effects by generalized rate equations obtained by adiabatic elimination of the polarization \cite{jakobsen}. In our TW description the frequency dependent gain is then selfconsistently determined from the polarization dynamics. In addition, all the physical mechanisms influencing multimode behavior mentioned above (phase sensitive interactions, SHB, etc.) are, in principle, included in the model. Spontaneous emission noise is also included by means of random polarization sources. We solve this model directly in the space-time domain and for large-signal situations. Another analysis which also considers a propagation equation for a semiconductor laser field is given in \cite{tromborg}: There the equation for the field is coupled to a rate equation for the global carrier number density and modulation response and stability of a given mode solution are discussed through small-signal differential equations which are solved numerically by transforming them into a matrix equation. An important feature of our TW model is the ease with which optical feedback from external mirrors or coupled cavities can be taken into account. A large number of studies exist on optical feedback effects on single-mode semiconductor lasers \cite{petterman,mork,elsasser,macinerney,besnard,feedbackregimes}, but it is only recently that some desciriptions of the interplay of multimode behavior and feedback have been given \cite{wu,ryan}. Strong feedback and multimode behavior have also been considered in dye lasers \cite{munroe}. The analyses in \cite{wu,ryan} are again based on multimode rate equations in which optical feedback from mode $i$ is only directly coupled to the same mode $i$. In our TW model the whole laser field propagates back and forth from the laser cavity to the external mirror. We will focus here on properties of mode selection of a multimode solitary laser by optical feedback. We will consider short external cavities \cite{munroe,tager} for which such phenomena such as coherence collapse for moderate feedback might be more easily avoided. We explore mode selection by changing the length of the external cavity: The field fed back into the laser cavity interacts with different intracavity modes depending on the length of the external cavity. We also explore how the strength of the feedback leads to selection of one mode and to suppression of the other modes. The paper is organized as follows: In section II the TW model is introduced within a slowly varying amplitude approximation for the counterpropagating waves in the FP cavity. The relevance of the different ingredients of the model for solid-state and semiconductor lasers is discussed. In particular, carrier diffusion is taken into account for semiconductor lasers. Results of the numerical analysis are given in Sects. III and IV for typical parameters of a semiconductor laser. In section III, we discuss the factors influencing multimode behavior of a solitary Fabry-P\'{e}rot laser: SHB, spontaneous emission and, for semicondcutor lasers, carrier diffusion. The effect of optical feedback and opportunities for mode selection are discussed in Section IV. \section{ Travelling Wave Model} We consider a Fabry-P\'{e}rot cavity filled with a two-level atomic medium. The Maxwell equation for a linearly polarized electric field in the cavity is \begin{eqnarray} \nabla^{2} E(\vec r, t) -\frac{ \eta ^{2}}{c^{2} }\frac{ \partial ^{2} E(\vec r, t)} {\partial t^{2} }= \frac{ 1}{\epsilon_0 c^{2}}\frac{ \partial ^{2} \Pi (\vec r, t)} {\partial t^{2} } \label{eqnarray:camp} \end{eqnarray} where $E$ is the electrical field, $ \Pi$ is the polarization of the atomic medium, $\eta$ the refraction index of the host medium, $\epsilon_0$ the electrical permittivity and $c$ the speed of light. The optical field inside the cavity consists in two identically polarized counter-propagating waves of the same frequency, \cite{n-m} \begin{eqnarray*} E(\vec r, t) = \left[ E^{+}(z, t)\exp{\left(\imath (k_0z-\omega_c t)\right)} + E^{-}(z, t)\exp{\left(-\imath (k_0z+\omega_c t)\right)} + c.c.\right.] \nonumber \end{eqnarray*} $E^{+}(z, t)$ is the slowly varying envelope of the electric field propagating forward in the Fabry-P\'{e}rot cavity and $E^{-}(z, t)$ is the slowly varying envelope of the electric field propagating backwards. The envelopes $E^{+}$ and $E^{-}$ are assumed to depend only on the longitudinal coordinate $z$ and we neglect any dependence on the transverse coordinates $x$, $y$. The carrier frequency $\omega_c$ is the frequency of the cavity mode closest to the maximum of the gain curve. We take $\omega_c$ as the reference frame for the optical frequency of the stationary solution, and $k_0$ is the wave number associated with $\omega_c$, $k_0=\eta\omega_c/c$. In the following we derive equations for $E^{+}(z, t)$ and $E^{-}(z, t)$ coupled to polarization and population inversion following the general scheme in \cite{fleck,n-m,wfjm} The polarization can be written as, \begin{eqnarray*} \Pi(\vec r, t) = P (z, t)\exp{\left(-\imath \omega_c t\right)} \nonumber \end{eqnarray*} where $P (z, t)$ is a slowly varying amplitude in $t$, which still keeps the fast $z$ dependence. Performing the slowly varying envelope approximation, we obtain \begin{eqnarray} \frac{ \eta }{c}\frac{\partial E^{+}}{\partial t} + \frac{\partial E^{+}}{\partial z} &=& \frac{ \imath \omega_c }{2\epsilon_0 c \eta} \langle P \exp{\left(-\imath k_0z\right)} \rangle \label{eqnarray:e+} \\ \frac{\eta}{c}\frac{\partial E^{-}}{\partial t} - \frac{\partial E^{-}}{\partial z} &=& \frac{ \imath \omega_c }{2\epsilon_0 c \eta} \langle P \exp{\left(\imath k_0z\right)} \rangle \label{eqnarray:e-} \end{eqnarray} where $\langle \rangle$ means an average over many wavelenghts. Equations (~\ref{eqnarray:e+}) and (~\ref{eqnarray:e-}) must be supplemented with boundary conditions for the electric field amplitudes at the facets of the FP cavity since we do not make any assumption on the laser longitudinal mode structure. We denote by $r_1$ and $r_2$ the field amplitude reflectivities of the cavity mirrors placed at $z=0$ and $z=L$, respectively: \begin{eqnarray} E^{+}(z=0,t)&=& r_1E^{-}(z=0,t) \nonumber \\ E^{-}(z=L,t)&=& r_2E^{+}(z=L,t) \label{eqnarray:bc} \end{eqnarray} An external cavity can also be taken into account modifying the second boundary condition to include the effect of an external mirror of reflectivity $r_3$ placed on the right of the laser cavity. Thus, our second boundary condition becomes, \begin{eqnarray*} E^{-}(z=L,t)&=& r_2E^{+}(z=L,t)-r_3 \sqrt{1-r_2^{2}} E^{+}(z=L,t-\tau) \end{eqnarray*} where $\tau$ is the round trip time of the field in the external cavity of length $L_{ext}$. Note that these boundary conditions take into account multiple reflections on the external mirror. Boundary conditions determine the complex phase constant, $k=(k_r +\imath k_i)$, of a plane wave solution of the form $E^{\pm}(z,t)=E_0^{\pm}(t) \exp{({\pm}\imath k z)}$. The real part of the wave vector, $k_r =\frac{\pi m}{L}$, yields the cavity modes of the laser resonator. The imaginary part of the wave vector is given by $k_i = |ln(R_1R_2)|/2L$, that accounts for the loss due to the cavity mirrors. The reference wavenumber $k_0$ associated with the carrier frequency $\omega_{c}$ corresponds to a particular cavity mode. The interaction of the electromagnetic wave with the medium is governed by the Bloch equations for two-level atoms. The optical Bloch equations are derived using the expressions for the total field $E(\vec r, t)$ and the polarization $P$ given above. The equations for $P$ and the population inversion $N$ are, \begin{eqnarray} \frac{ \partial P}{\partial t } &=& -\left( \gamma_{\bot} +\imath (\omega_{12} - \omega_{c}) \right) P + \frac{\Theta_{12}\Theta_{21} } {\imath \hbar} N \left( E^{+}e^{\imath k_0z} + E^{-}e^{-\imath k_0z} \right) \label{eqnarray:pol} \\ \frac{ \partial N}{\partial t } &=& - \gamma_{\|}\left( N -N_0 \right) - \frac{2 \imath}{\hbar} \left[\left( E^{+} P ^{*} - E^{-*} P \right)e^{\imath k_0z} + \left( E^{-} P ^{*} - E^{+*} P \right) e^{-\imath k_0z} \right] \label{eqnarray:port} \end{eqnarray} where $\Theta_{12}$ and $\Theta_{21}$ are components of the dipole moment and $\omega_{12}$ is the transition frequency between the two levels. The difference between the transition and carrier frequencies gives the detuning $\delta=\omega_{12} - \omega_{c}$. The decay rate for the polarization and population inversion are $\gamma_{\bot}$ and $\gamma_{\|}$, respectively, and $N_0$ gives the pumping level. \noindent It is convenient to scale the variables in equations (~\ref{eqnarray:e+})--(~\ref{eqnarray:port}) in the following way, \begin{eqnarray*} E^{\pm} \longrightarrow \frac{ \hbar}{\sqrt{2\Theta_{12}\Theta_{21}}} \frac{1}{\sqrt{\gamma_{\bot}\gamma_{\|} }} E^{\pm} \nonumber \\ P \longrightarrow - \frac{2\imath\epsilon_0}{\omega_{c}} \frac{ \hbar}{\sqrt{2\Theta_{12}\Theta_{21}}}\sqrt{ \frac{\gamma_{\bot}} {\gamma_{\|} }}P \nonumber \\ N \longrightarrow\frac{2\hbar\epsilon_0}{\omega_{c}} \frac{1}{2\Theta_{12}\Theta_{21}}N \nonumber \end{eqnarray*} \noindent The longitudinal coordinate is scaled to the length of the laser medium, $L$, and the time to the group velocity over $L$, $c/(\eta L)$, \begin{eqnarray*} z \longrightarrow z/L \nonumber \\ t \longrightarrow \frac{c}{\eta L} t\nonumber \end{eqnarray*} With these scalings, equations (~\ref{eqnarray:e+})--(~\ref{eqnarray:port}) become, \begin{eqnarray} \frac{\partial E^{+}}{\partial t} + \frac{\partial E^{+}}{\partial z} &=& -\langle P \exp{\left(-\imath k_0z\right)} \rangle \label{eqnarray:e2+} \\ \frac{\partial E^{-}}{\partial t} - \frac{\partial E^{-}}{\partial z} &=& -\langle P \exp{\left(\imath k_0z\right)} \rangle \label{eqnarray:e2-} \\ \frac{ \partial P}{\partial t } &=& -\gamma_{\bot}\left[\left( 1 +\imath \frac{\delta}{\gamma_{\bot}} \right) P - N \left( E^{+}e^{\imath k_0z} + E^{-}e^{-\imath k_0z} \right)\right] \label{eqnarray:pol2} \\ \frac{ \partial N}{\partial t } &=& - \gamma_{\|} \left[\left( N -N_0 \right) + \left[\left( E^{+} P ^{*} - E^{-*} P \right)e^{\imath k_0z} + \left( E^{-} P ^{*} - E^{+*} P \right) e^{-\imath k_0z} \right] \right] \label{eqnarray:port2} \end{eqnarray} The spatial variations in polarization and population over wavelength distances are treated by means of a standard Fourier series expansion \cite{lamb}, \begin{eqnarray} P(z,t)= e^{\imath k_0z} \sum_{p=0}^{\infty}P^{+}_{(p)}e^{2\imath p k_0z} + e^{-\imath k_0z} \sum_{p=0}^{\infty}P^{-}_{(p)}e^{-2\imath p k_0z} \label{eqnarray:f1}\\ N(z,t)=N_{(0)} + \sum_{p=1}^{\infty}\left[ N_{(p)}e^{2\imath p k_0z} + (*) \right] \label{eqnarray:f2} \end{eqnarray} Substituting (~\ref{eqnarray:f1}) and (~\ref{eqnarray:f2}) in (~\ref{eqnarray:e2+})--(~\ref{eqnarray:port2}), the following coupled equations are obtained \begin{eqnarray} &&\frac{\partial E^{+}}{\partial t} + \frac{\partial E^{+}}{\partial z} = - P^{+}_{(0)} \label{eqnarray:ep+} \\ &&\frac{\partial E^{-}}{\partial t} - \frac{\partial E^{-}}{\partial z} = - P^{-}_{(0)} \label{eqnarray:ep-} \\ &&\frac{ \partial P^{+}_{(p)}}{\partial t } = -\gamma_{\bot}\left[\left( 1 +\imath \frac{\delta}{\gamma_{\bot}} \right) P^{+}_{(p)} +\left( N_{(p)} E^{+} + N_{(p+1)}E^{-} \right)\right] + \xi_{(p)}^{+}(z,t) \label{eqnarray:polp+} \\ &&\frac{ \partial P^{-}_{(p)}}{\partial t } = -\gamma_{\bot}\left[\left( 1 +\imath \frac{\delta}{\gamma_{\bot}} \right) P^{-}_{(p)} +\left( N_{(p+1)}^{*} E^{+} + N_{(p)}^{*}E^{-} \right)\right]+ \xi_{(p)}^{-}(z,t) \label{eqnarray:polp-} \\ &&\frac{ \partial N_{(p)}}{\partial t } = - \gamma_{\|} \left[ N_{(p)} + \left( E^{-} P _{(p)}^{-*} + E^{+} P_{(p-1)}^{-*} - E^{+*} P _{(p)}^{+} - E^{-*} P_{(p-1)}^{+}\right) \right] \label{eqnarray:portp}\\ &&\frac{ \partial N_{(0)}}{\partial t } = - \gamma_{\|} \left[ \left( N_{(0)} -N_0 \right) + \left( E^{+} P ^{+*}_{(0)} + E^{-} P^{-*}_{(0)} + (*) \right) \right] \label{eqnarray:port0} \end{eqnarray} Equations (~\ref{eqnarray:ep+})-(~\ref{eqnarray:port0}) are numerically solved together with the boundary conditions (~\ref{eqnarray:bc}) for the field amplitudes $E^{\pm}$. In these equations we have also included Langevin noise terms $\xi_{(p)} (z,t)$ as polarization sources. Such spatially distributed noise terms model independent spontaneous emission process in different points of the cavity. They are taken to be Gaussian white noise in space and time with zero mean and correlations $\langle \xi_{0} (z, t) \xi_{0}^{*}(z', t') \rangle = \beta \delta(t - t') \delta(z - z')$. The spontaneous emission noise intensity is measured by $\beta$, which depends, in principle, on the instantaneous value of the population levels \cite{fleck,haken,lax}. In practice we approximate the spontaneous emission intensity by a constant parameter \cite{lax} and, in addition, we will only include noise sources for $p=0$. If we want to take into account spatial variations of the material variables, terms beyond the $p=0$ term must be retained in (~\ref{eqnarray:ep+})-(~\ref{eqnarray:port0}). Such spatial variations come from the optical wave inside the cavity and from the finite reflectivity of the mirrors. In practice, we will truncate our expansions after the first harmonic $p=1$. We refer to the terms of order $p=1$ as "grating terms", because their effect is comparable with the introduction of a spatial grating in the cavity at a spatial frequency that corresponds to half the natural wavelengths of the fields inside the cavity. Adiabatic elimination of the polarization variables within this TW model is not possible since it leads to a set of equations which are intrinsically unstable as can be easily seen by a straightforward linear stability analysis of the reduced equations. The above set of Equations(~\ref{eqnarray:ep+})-(~\ref{eqnarray:port0}) contains a complete description (within the slowly varying approximation) of an homogeneously broadened two-level laser. It should provide a good description of multimode behavior of homogeneously broadened dye or solid state lasers. When this model is used to learn about multimode behavior of semiconductor lasers, $N$ has to be identified with carrier number, $N_0$ with the carrier number at transparency and $\gamma_{\|} N_0$ with the injection current $J$. For semiconductor lasers some other aspects have to be considered, mainly carrier diffusion and inhomogeneous broadening from the band structure. Carrier diffusion can be taken into account by introducing a diffusion term $D \nabla^{2} N$ in (~\ref{eqnarray:port}), where $D$ is the diffusion coefficient. With this term, the equations for the two first orders of the Fourier expansion for the carrier number, are modified as follows: \begin{eqnarray*} &&\frac{ \partial N_{(0)}}{\partial t } = D\nabla^{2} N_{(0)} + J - \gamma_{\|} \left[ N_{(0)} - \left( E^{+} P ^{+*}_{(0)} + E^{-} P^{-*}_{(0)} + (*) \right) \right] \nonumber\\ &&\frac{ \partial N_{(1)}}{\partial t } = - \gamma_{\|} \left[ (1 + \frac{4Dk_0^{2}}{\gamma_{\|}}) N_{(1)} - \left( E^{-*} P _{(0)}^{+*} + E^{+} P_{(0)}^{+*} + E^{+*} P _{(1)}^{+} + E^{-*} P_{(1)}^{-*}\right) \right] \nonumber \end{eqnarray*} For typical parameter values (see table $I$), it is easy to see that the term $D\nabla^{2} N_{(0)}$ can be neglected in comparison with $\gamma_{\|}N_{(0)}$. The equation for $N_{(0)}$ remains then essentially unchanged and the main effect of carrier diffusion appears in the equation for $N_{(1)}$ through a modification, $4Dk_0^{2}$, of the damping coefficient, where $k_0$ the wavenumber of the mode around which we have made the slowly varying amplitude approximation. Our TW homogeneously broadened two-level model includes the coherent coupling between the optical field and the two-level medium by means of the optical Bloch equations, allowing for phase sensitive multimode operation and four-wave mixing. An important characteristic of semiconductor lasers is the strong phase-amplitude dynamical coupling often described by the linewidth enhancement factor $\alpha$-factor \cite{henry}. Such phase sensitivity is also produced by cavity detuning in two-level models. In fact, semiconductor laser behavior has been modeled by identifying a constant $\alpha$-factor with a large detuning (the same for all modes) in a two-level description \cite{mandel}. However, in such a two-level model the maximum gain occurs at zero detuning, where the carrier-induced refractive index change vanishes, while for a semiconductor laser there is asymmetric gain with maximum gain away from resonance. The origin of the $\alpha$-factor in semiconductors has been explained by approximating the electron-hole recombinations at different frequencies by an ensemble of two-level atoms with different transition energies \cite{vahala}. In order to match the typical measured values of $\alpha \sim 1-10$ with a detuning in two-level models, lasing away from resonance must be artificially enforced. This is not possible in our multimode travelling wave model as modes will appear where the different cavity modes find strong gain independent of the detuning of a particular reference mode. Hence the effects of the $\alpha$-factor are not taken into account in (~\ref{eqnarray:ep+})-(~\ref{eqnarray:port0}), but, on the other hand we note that there are highly doped semiconductor lasers for which, at low temperatures, a nearly homogeneously broadened gain is expected \cite{yamamoto}. In summary, (~\ref{eqnarray:ep+})-(~\ref{eqnarray:port0}) supplemented with carrier diffusion incorporate a number of important mechanisms needed to understand multimode behavior in semiconductor lasers, but one should be aware that additional steps might be needed for a full understanding of multimode behavior of inhomogenously broadened semiconductor lasers. In our numerical analysis reported in sections 3 and 4 we use parameter values (see table $I$) which correspond to the characteristic time scales of a semiconductor laser. Equations (~\ref{eqnarray:ep+})--(~\ref{eqnarray:port0}), with the given boundary conditions (~\ref{eqnarray:bc}), are integrated numerically with a semi-implicit finite difference integration scheme \cite{fleck}. We use an integration step $\Delta z = \Delta t = 1/300$, which corresponds to a spatial step of $0.75 \mu m$ and a temporal step of $9$ $fsec$ for $\eta = 3.5$ and $L=250$ $\mu m$. As initial conditions we use a random distribution of the electric field along the cavity. We take a real electric field distributed between $0$ and $1$ with an amplitude of $10^{-4}$. For the polarization and the population variables we take initial values $P^{\pm}=0$ and $N= 0.95 N_0$. We have checked that steady state properties do not change when using a non-random initial field profile of the same amplitude and which satisfies boundary conditions. The stochastic part of the equation for the polarization is integrated using an Euler algorithm of order $(\Delta t)^{1/2}$ \cite{maxi82,greiner}. Independent Gaussian random numbers are generated at each point in the spatio-temporal grid using an optimized version of the Burlish-Stoer algorithm \cite{raul}. \section{Factors influencing multimode behavior of a Fabry P\'{e}rot laser} The TW model described above includes three main factors which influence mutimode behavior: a) Spatial hole burning (SHB), b) Spontaneous emission noise and c) Carrier diffusion. These three effects compete with different relative importance depending on the operation point of the laser above threshold. In this section we consider these factors for a solitary laser with no external mirror ($r_3=0$). Spatial Hole Burning appears via the "grating terms" associated with the Fourier modes in (~\ref{eqnarray:f1})--(~\ref{eqnarray:f2}) for $p>0$. The short range spatial variation of the atomic variables described by these Fourier modes tends to be washed out by carrier diffusion. On the other hand, close to threshold spontaneous emission noise is expected to be more important in exciting many modes, while it should become less relevant for high power emission well above threshold. For solid state lasers there is no diffusion of the atoms of the active medium and spatial hole burning is the dominant effect, while for semiconductor lasers carrier diffusion is very important and the noise strength is high. In the following we discuss these factors separately to elucidate their effects and relative importance. As a reference situation we show in Fig. 1. the Field Power Spectrum (FPS) obtained when spontaneous emission noise and carrier diffusion are neglected and the Fourier expansion in (~\ref{eqnarray:f1})--(~\ref{eqnarray:f2}) is truncated at $p=0$ (no SHB). The FPS is obtained by a standard FFT routine with a time window of $2$ $ns$, ($2^{16}$ points) when the laser has reached a steady state $20 ns$ after switch-on. We obtain clear single mode emission for the injection current $J=1.5 J_{th}$ shown. We find such single mode behavior for different current values, closer to threshold and also well above threshold. \subsection{Multimode emission with no carrier diffusion} {\bf a) Spatial Hole Burning:} Fig.~2 shows the FPS obtained by neglecting spontaneous emission noise and carrier diffusion and taking into account SHB by truncating the Fourier expansion in (~\ref{eqnarray:f1})--(~\ref{eqnarray:f2}) at $p=1$. It is clear that the "grating terms" induce the growth of additional modes above threshold. Only very close to threshold (Fig.~2.a), (up to $J=1.03 J_{th}$), the behavior remains single-mode. Further above threshold SHB causes the excitation of side modes located at the frequencies given by the cavity length and the group velocity inside the cavity, giving rise to multimode laser emission. Fig.~2 shows the results obtained for zero detuning. For a detuning of $\delta=70 GHz$ we obtain the same qualitative behavior, but the power spectrum is no longer symmetric around the cavity frequency in the case of multimode behavior since the gain peak is then $\delta=70 GHz$ away from the reference cavity mode (see also Fig.~5a). We point out ( Fig.~3), that for very high levels of pumping ($J=15 J_{th}$ and above), we observe a splitting of the FPS into two groups of modes as reported for solid state lasers in \cite{khanin}. {\bf b) Spontaneous Emission Noise:} Fig.~4 shows the FPS (also for zero detuning) and the same pumping levels as Fig.~2, but with spontaneous emission noise added. Noise excites side modes at pump levels for which SHB alone does not destroy single mode emission. Noise has no important effects for large pump values. For $J>1.9 J_{th}$ the Side Mode Suppresion Ratio (SMSR) of the modes excited by spontaneous emission is above 20 dB. In general both SHB and spontaneos emission noise contribute to multimode emission. We note that a finite time interval of averaging in the presence of noise causes small asymmetries in the FPS while did not occur for $\beta=0$ (Figs.2 and 3). Figs.~5.a-d show the FPS for a detuning $\delta=70 GHz$ taking into account SHB and spontaneous emission noise. The nonzero detuning leads to a nonsymmetric FPS. As the pumping level increases, the number of lasing modes increases as well as the power of the side-modes with respect to the main one. However, there is a level of pumping ($J> 2.5 J_{th}$) beyond which additional side-modes do not grow and the number of relevant longitudinal modes remains constant. In addition, for $J> 4 J_{th}$, the power ratio between the different modes remains constant. Except for the different time scales involved, the general qualitative behavior as the pumping is increased (as shown in Fig. 5a-d) it is the one typically observed in solid state lasers \cite{khanin}. \subsection{The Effect of Carrier Diffusion} As discussed in section II, the main effect of diffusion in our mathematical modelling is to give rise to an effective damping rate $ \gamma_{\|} + 4Dk_0^{2}$ for the first Fourier mode of the carrier number $N_1$. This higher damping rate reduces the effect of SHB. Fig.~5 shows a comparison of the FPS at different current levels for the case of zero diffusion, in the left column, and for a given diffusion coefficient, in the right column. In this case, the central mode grows faster than the adjacent modes as the current increases and eventually the amplitude of the adjacent modes effectively saturates. Thus we observe restoration of single mode emission for large enough pumping level as a consequence of diffusion. However, smaller values of the diffusion coefficient do not fully kill multimode emission far above threshold. For such multimode behavior the total output power of the laser exhibits fast oscillations associated with mode beating and slower oscillations of the envelope. The latter oscillations are the ordinary relaxation oscillations of the laser (see inset Fig.~6a). For intermediate pump levels and for the typical diffusion coefficient of semiconductors used in Fig.~5 we have observed simultaneous operation of few modes, but we have not observed mode-hopping of the form of switching of most of the power from one mode to another with nearly constant total power when the average spectrum showed only a few modes. An early discussion of multimode behavior in semiconductor lasers was given in \cite{lee} in terms of rate equations with a parabolic approximation for the dependence of the gain parameters of different modes as a function of their wavelength. Emphasis in that work was on the effect of spontaneous emission noise. They showed that, as the current increases, the width of the emitted mode spectrum envelope narrowed. In their model the main mode grew faster than the adjacent modes as the current was increased near threshold, and eventually, the amplitude of the side modes saturated with increasing the current. Hence, over a certain power level, additional power increases with increasing current were due to the growth of a single mode. This general behavior is the one we find when we neglect SHB and carrier diffusion. The latter two effects are discussed phenomenologically in \cite{lee} by decreasing the value of the gain coefficients of the modes as a consequence of diffusion. The onset of SHB is identified with the output power level for which such a decrease is equal to the difference in gain between the lasing mode and its nearest neighbor. The analysis in \cite{lee} describes different regions of multimode behavior depending on pump level and cavity length. The general picture was a transition from noise dominated multimode emission close to threshold to single mode emission for higher injection. A second transition to multimode emission dominated by SHB was predicted further above threshold. In our TW model the different gain at different cavity frequencies and the gain modifications due to SHB and diffusion appear as a consequence of the consistent dynamical equations. In our numerical analysis we obtain a transiton from multimode to single mode behavior asthe injection current is increased keeping all other parameters fixed (Figs.5e-h). For a fixed injection current, single or multimode behavior depends on the value of the diffusion coefficient, so that, the injection current for which single mode operation is achieved could be changed by the possible dependence of the diffusion coefficient on the injection current which we have neglected. We finally note that, for the range of parameters we have explored, we do not observe the second transition from single to multimode behavior obtained in \cite{lee} as the injection current is further increased. \section{Optical feedback and mode selection} Optical feedback from an external mirror combines with the intracavity fields and can enhance the power of a particular intracavity mode. As a consequence it can lead to mode selection depending on the feedback level and the external cavity length. We first consider weak feedback levels to avoid instabilities of the output intensity. We have analyzed short external cavities of different lengths up to $L_{ext}=1 cm$. For the shortest cavity considered with $L_{ext}=75$ $\mu m$ and weak feedback, single mode selection is achieved, while for the longest external cavities considered, many intracavity modes are excited by the weak feedback. For strong feedback and the shortest cavity considered, two groups of intracavity modes can be selected. For this analysis of feedback effects we have chosen an intermediate value of carrier diffusion for which multimode emission occurs in a large range of pump values ($D=2\times 10^{-4}$ $m^{2}s^{-1}$) and the same detuning used for Fig.~5 ($\delta=70 GHz$). For the shortest external cavity of $L_{ext}=75$ $\mu m$, the external cavity modes are separated by $2000 GHz$, while the laser cavity modes, for our paremeters values, have an intermode spacing of $171 GHz$. The laser cavity mode closest to the maximum of the gain curve which resonates with one of the external modes is excited and can be selected. The intracavity mode spectrum is centered in such a way that the mode at zero frequency coincides with an external cavity peak. In this case, we excite the zero order mode, $m=0$. We can select the mode at a frequency of $\approx171 GHz$, $m=1$ by changing the external cavity by $\frac{171}{2000}\frac{ \lambda}{2}$. The selection of the mode $m=1$ is seen in Fig.~6, where single mode emission with a SMSR of 20 dB is obtained for an external mirror field reflectivity $r_3$ of just $1\%$. We show in Fig.~6 the FPS at the rigth mirror for different values of the external mirror reflectivity (right column), and the total output intensity as a function of time (left column), in the steady state regime of laser emission. In the output intensity variations one can identify the fundamental modulation fequency as being the beat frequency between adjacent intracavity modes. From a) to d), we observe how the amplitude of these oscillations is reduced being almost zero when nearly single mode emission is achieved (d). In the inset of Fig.~6.a) and d), we plot the intensity as a function of time for a longer time interval of 2 nsec. The envelope of the oscillations of the output power exhibits slower oscillations associated with the relaxation oscillations of the laser. These oscillations are present for all weak feedback levels with a frequency which is only slightly modified by weak feedback: For the solitary laser the frequency is $\approx 1.21 GHz$ (Fig.6a) and it becomes $\approx 1.18 GHz$ for $1\%$ field reflectivity (Fig.6e). Results on the effect of weak feedback from an external cavity of $L_{ext}=450$ $\mu m$ (giving a mode spacing of 333GHz) are shown in Fig.~7. For this intermediate length of the external cavitites that we have considered, the external mode separation coincides with twice the internal one, so that the laser might be forced to reduce its multimode emission to lasing in two modes. We consider here a lower value of pumping level than for the shortest cavity in Fig.~6 to have a larger number of lasing modes. We observe that two modes can be selected with a SMSR of $\approx 20 dB$ for $r_3 = 2\%$. Note that the peaks in the spectrum alternate in height, as alternate modes of the laser are nearly resonant with the external cavity as well. The behavior shown in Fig.~8 for an external cavity of $L_{ext}= 1$ $cm$ (with mode spacing of $\approx 15.4 GHz$) is typical of what is observed for $L_{ext}>0.1 cm$. For these external cavity lengths mode selection is not found, and instead, weak feedback undamps many intracavity modes for a given value of the field reflectivity $r_3$. The excitation of many modes occurs, for the parameter values in Fig.~8, for $r_3=2\%$, but time dependence of the output power becomes chaotic for field reflectivities $r_3>5\%$. Finally, we have performed a preliminary study for different feedback levels in the region of moderate and strong feedback. While a standard classification of feedback regimes is well known for single mode lasers \cite{feedbackregimes}, no such classification seems available for multimode lasers. The consideration of strong feedback levels is not an extra difficulty in our Travelling Wave model since it properly takes into account multiple external reflections between the two mirrors that define the external cavity. However, we have checked that these multiple external reflections do not introduce significant effects for $r_3 \leq 5\%$. Our results for the shortest external cavity are shown in Fig.~9 which is a continuation of the sequence detailed in Fig.~6 for increasing field reflectivity $r_3$. In the region of weak feedback ($r_3 < 5\%$) shown in Fig.~6 mode selection is achieved. For moderate feedback levels ($r_3 \approx 5-15\%$) the laser output becomes unstable exhibiting a broad Field Power Spectrum (Fig.~9.f). However, for high feedback levels as shown in Figs.~9.g to ~9.j, two groups of intracavity modes are selected and a few-mode emission is recovered. The spacing between these two groups of modes ($\approx 2000 GHz$) is given by the external roundtrip time or separation of external cavity resonances. The internal cavity modes are still well defined within the selected groups due to the broad external cavity spectrum. \acknowledgements We have greatly benefited from discussions with C. Ning and R. Indik and useful comments from N. B. Abraham. This work has been supported by AFOSR contract F49620-94-1-0144DEF. MH and MSM also acknowledge financial support from Comision Interministerial de Ciencia y Tecnologia (CICYT, Spain) project TIC93-0744. \begin{thebibliography}{99} \bibitem{baer} T. Baer, "Large-amplitude fluctuations due to longitudinal mode coupling in diode-pumped intracavity-doubled Nd:YAG lasers", J. Opt. Soc. Am. B, {\bf 3}, 1175-1180 (1986). \bibitem{TSM} C. L. Tang, H. Statz and G. de Mars, "Spectral output and spiking behavior of solid-state lasers", J. Appl. Phys., {\bf 34}, 2289-2295 (1963). \bibitem{khanin} Ya. Khanin, {\em Principles of Laser Dynamics}, to be published by Elsevier. \bibitem{roy} K. Wiesenfeld, C. Bracikowski, G. James and R. Roy, Phys. Rev. Lett., {\bf 65}, 1749 (1990). \bibitem{glorieux} K. Otsuka, P. Mandel, S. Bielawski, D. Derozier and P. Glorieux, "Alternative time scale in multimode lasers", Phys. Rev.{\bf A}, {\bf 46}, 1693-1695 (1992). \bibitem{raymer1} M. G. Raymer, Z. Deng and M. Beck, "Strong-field dynamics of a multimode, standing-wave dye laser", J. Opt. Soc. Am. B, {\bf 5}, 1588-1595, (1988). \bibitem{raymer2} I. McMackin, C. Radzewicz, M. Beck, and M. G. Raymer, "Instabilities and chaos in a multimode, standing-wave, cw dye laser", Phys. Rev.{\bf A}, {\bf 38}, 821-832 (1988). \bibitem{agrawal} G. P. Agrawal and N. K.Dutta, {\em Long-Wavelength Semiconductor Laser}, Van Nostrand Reinhold Company (1986). \bibitem{lee} Tien-Pei Lee, Charles A. Burrus, John A. Copeland, Andrew G. Dentai and Dietrich Marcuse, "Short-Cavity InGaAsP Injection Lasers: Dependence of Mode Spectra and Single-Longitudinal-Mode Power on Cavity Length", IEEE J. Quantum Electron., {\bf QE-18}, 1101-1113 (1982). \bibitem{mandel} P. Mandel, C. Etrich and K. Otsuka, "Laser Rate Equations with Phase-Sensitive Interactions", IEEE J. Quantum Electron., {\bf QE-29}, 836-843 (1993). C. Etrich, P. Mandel, N. B. Abraham and H. Zeghlache, "Dynamics of a Two-Mode Semiconductor Laser", IEEE J. Quantum Electron., {\bf QE-28}, 811-821 (1992). \bibitem{yamamoto} S. Inoue, H. Ohzu, S. Machida and Y. Yamamoto, "Quantum correlation between longitudinal-mode intensities in a multimode squeezed semiconductor laser", Phys. Rev. {\bf A}, {\bf 46}, 2757-2765 (1992). \bibitem{refneal} T. J. Bridges and W. W. Rigrod, "Output Spectra of the Argon Ion Laser", IEEE J. Quantum Electron., {\bf QE-1}, 303-308 (1965). \bibitem{sigman} A. E. Siegman, {\em Lasers}, University Sciences Books (1986). \bibitem{gray} G. Gray and R. Roy, "Noise in nearly-single-mode semiconductor lasers", Phys. Rev.{\bf A}, {\bf 40}, 2452-2462 (1989). \bibitem{ohtsu} M. Ohtsu and Y. Teramachi, "Analyses of Mode Partition and Mode Hopping in Semiconductor Lasers", IEEE J. Quantum Electron., {\bf QE-25}, 31-38 (1989). \bibitem{lugiato} L. A. Lugiato, K. Wang and Y. Sun, "Phase-dependent laser equations", Optics Comm., {\bf 105}, 397-410 (1994). \bibitem{fleck} J. A. Fleck, Jr., "Ultrashort-Pulse Generation by Q-Switched Lasers", Phys. Rev.{\bf B}, {\bf 1}, 84-100 (1971). \bibitem{n-m} Alan C. Newell and J. V. Moloney, {\em Nonlinear Optics}, Addison-Wesley Publishing Company (1991). \bibitem{wfjm} W. Forysiak and J. V. Moloney, ``Dynamics of synchronously pumped mode-locked color-center lasers'', Phys. Rev.{\bf A}, {\bf 45}, 3275-3288 (1992); ``Mode-locking dynamics of synchronously pumped color-center lasers: Fabry-Perot and ring geometries'', Phys. Rev. {\bf A}, {\bf 45}, 8110-8120 (1992). \bibitem{jakobsen} P. K. Jakobsen, J. V. Moloney, A. C. Newell and R. Indik, "Space-time dynamics of wide-gain-section lasers", Phys. Rev.{\bf A}, {\bf 45}, 8129-8137 (1992). \bibitem{tromborg}B. Tromborg, Hans Erik Lassen and Henning Olesen, "Travelling wave analysis of semiconductor lasers: Modulation responses, mode stability and quantum mechanical treatment of noise spectra", IEEE J. Quantum Electron., {\bf QE-30}, 1875-1889 (1994), and references there in. \bibitem{petterman} K. Petermann, {\em Laser Diode Modulation and Noise}, Kluwer Academic Publishers (1991). \bibitem{mork} L. N. Langley, K. A. Shore and J. Mork, "Dynamical and noise properties of laser diodes subject to strong feedback", Optics Lett., {\bf 19}, 2137-2139 (1994). \bibitem{elsasser} J. Sacher, W. Elsasser and E. O. G\"obel, "Nonlinear Dynamics of Semiconductor Laser Emission Under Variable Feedback Conditions", IEEE J. Quantum Electron., {\bf QE-27}, 373-379 (1991). \bibitem{macinerney} Jun Ye, Hua Li and J. McInerney, "Period-doubling route to chaos in a semiconductor laser with weak optical feedback", Phys. Rev. {\bf A},{\bf 47}, 2249-2252 (1993). "Detailed analysis of coherence collapse in semiconductor lasers", IEEE J. Quantum Electron., {\bf QE-29} 2421 (1994). \bibitem{besnard} P. Besnard, B. Meziane and G. M. Stephan, "Feedback Phenomena in a Semiconductor Laser Induced by Distant Reflectors", IEEE J. Quantum Electron., {\bf QE-29}, 1271-1284 (1993). \bibitem{feedbackregimes} R. W. Tkach and A. R. Chraplyvy, "Regimes of feedback in 1.5 Micrometer distributed feedback lasers", J. Lightwave Technol.,{\bf LT-4}, 1665-1661 (1986). \bibitem{wu} Huoy-jong Wu and Hung-chun Chang, "Mode Partition in Semiconductor Lasers with Optical Feedback", IEEE J. Quantum Electron., {\bf QE-29}, 2154-2162 (1993). \bibitem{ryan} A. T. Ryan, G. P. Agrawal, G. R. Gray and E. C. Gage, "Optical-Feedback-Induced Chaos and its Control in Multimode Semiconductor Lasers", IEEE J. Quantum Electron., {\bf QE-30}, 668-679 (1994). \bibitem{munroe} M. Munroe, S. E. Hodges, J. Cooper and M. G. Raymer, "Total intensity modulation and mode hopping in a coupled-cavity as a result of external cavity length variations", Optics Lett., {\bf 19}, 105-107 (1994). \bibitem{tager} A. A. Tager and B. B. Elenkrig, "Stability regimes and High-Frequency Modulation of Laser Diodes with Short External Cavity", IEEE J. Quantum Electron., {\bf QE-29}, 2886-2890 (1993). \bibitem{lamb} S. Stenholm and W. E. Lamb, Jr., "Semiclassical Theory of a High-Intensity Laser", Phys. Rev., {\bf 181}, 618-635 (1969). \bibitem{haken} H. Haken and H. Sauermann, "Nonlinear Interaction of Laser Modes", Z. Phys. {\bf 173}, 261 (1963). \bibitem{lax} M. Lax, "Classical Noise. V. Noise in Self-Sustained Oscillators", Phys. Rev. {\bf A}, {\bf 160}, 290-306 (1967). \bibitem{henry} C. H. Henry, "Theory of spontaneous emission noise in open resonators and its application to lasers and optical amplifiers", J. Lightwave Technol.,{\bf LT-4}, 288-297 (1986). \bibitem{vahala} K. Vahala and A. Yariv, "Semiclassical Theory of Noise in Semiconductor Lasers", IEEE J. Quantum Electron., {\bf QE-19}, 1096-1101 (1983). \bibitem{maxi82} J. M. Sancho, M. San Miguel, S. L. Katz and J. D. Gunton, "Analytical and numerical studies of multiplicative noise", Phys. Rev. {\bf A}, {\bf 26}, 1589-1607 (1982). \bibitem{greiner} A. Greiner, W. Strittmatter and J. Honerkamp, "Numerical Integration of Stochastic Differential Equations", J. Stat. Phys., {\bf 51}, 95 (1988). \bibitem{raul} R. Toral and A. Chakrabarti, "Generation of Gaussian distributed random numbers by using a numerical inversion method", Computer Physics Comm., {\bf 74}, 327-334 (1993). \end{thebibliography} \begin{table} \caption{Meaning and values of the different parameters appearing in the model} \begin{tabular}{|c|c|c|c|} Parameter & Meaning & Value & Units \\ \hline \hline $r_1$ & Field reflectivity of left side mirror & 99.5$\%$ &-\\ $r_2$ & Field reflectivity of right side mirror & 56.6$\%$ &-\\ L & Length of the diode cavity & 250 & $\mu m$ \\ $\gamma_{\|}$ & Inverse carrier lifetime & 2 $\times 10^{8}$ & $s^{-1}$ \\ $\gamma_{\bot}$ & Inverse polarization characteristic time & 1 $\times 10^{13}$ & $s^{-1}$ \\ $\beta $ & Spontaneous emission rate & 1$\times 10^{4}$ & $s^{-1}$ \\ $\eta$ & Index of refraction & 3.5 &-\\ $\delta$ & Detuning & 0 or 70 & GHz \\ \end{tabular} \end{table} \newpage \begin{figure} \caption{:Field power spectrum (log scale) for $J=1.5 J_{th}$ neglecting SHB, spontaneous emission noise ($\beta=0$) and diffusion ($D=0$). $\delta=0$.} \end{figure} \begin{figure} \caption{:Field power spectrum (log scale) for: a) $J=1.02 J_{th}$, b) $J=1.7 J_{th}$ and c) $J=3.5 J_{th}$, neglecting spontaneous emission noise ($\beta=0$) and diffusion ($D=0$). $\delta=0$.} \end{figure} \begin{figure} \caption{:Field power spectrum (log scale) for $J=15 J_{th}$ neglecting spontaneous emission noise ($\beta=0$) and diffusion ($D=0$). $\delta=0$.} \end{figure} \begin{figure} \caption{:Field power spectrum (log scale) for same pumping levels than in Fig.2. Spontaneous emission noise $\beta=1\times 10^{4}$ $s^{-1}$. Diffusion is neglected ($D=0$) and $\delta=0$.} \end{figure} \begin{figure} \caption{:Field power spectrum (linear scale) for no diffusion (left column) and for $D=4\times 10^{-4} m^2 s^{-1}$ (right column) and different pumps values: a,e) $J=1.05 J_{th}$, b,f) $J=1.5 J_{th}$, c,g) $J=2.5 J_{th}$ and d,h) $J=3.5 J_{th}$. Spontaneous emission noise $\beta=1\times 10^{4}$ $s^{-1}$ and $\delta=70$ GHz.} \end{figure} \begin{figure} \caption{: Total intensity as a function of time (left column) and Field Power Spectrum (right column) for $J=1.5 J_{th}$, $L_{ext} = 75.055\mu m$, $D=10^{-4} m^{2}s^{-1}$, $\delta=70$ GHz, $\beta=1\times 10^{4}$ $s^{-1}$ and different external reflectivities: a,e) $r_3=0\%$, b,f) $r_3=0.1\%$, c,g) $r_3=0.5\%$ and d,h) $r_3=1\%$. The inserts in a) and d) correspond to a plot of the total intensity as a function of time in a large time window. } \end{figure} \begin{figure} \caption{:Total intensity as a function of time (left column) and Field Power Spectrum (right column) for $J=1.3 J_{th}$, $L_{ext} = 450\mu m$, $D$, $\delta$ and $\beta$ the same than in Fig.5, and different external reflectivities: a,e) $r_3=0\%$, b,g) $r_3=0.2\%$, c,h) $r_3=0.5\%$, d,i) $r_3=1\%$ and e,j) $r_3=2\%$.} \end{figure} \begin{figure} \caption{:Total intensity as a function of time (left column) and Field Power Spectrum (right column) for $J=1.4 J_{th}$, $L_{ext} = 1 cm$, $D$, $\delta$ and $\beta$ the same than in Fig.5, and different external reflectivities: a,d) $r_3=0\%$, b,e) $r_3=0.75\%$ and c,f) $r_3=5\%$.} \end{figure} \begin{figure} \caption{: Total intensity as a function of time (left column) and Field Power Spectrum (right column) for $J=1.5 J_{th}$, $L_{ext} = 75\mu m$, $D$, $\delta$ and $\beta$ the same than in Fig.5, and different external reflectivities: a,f) $r_3=10\%$, b,g) $r_3=15\%$, c,h) $r_3=25\%$, d,i) $r_3=35\%$ and e,j) $r_3=40\%$.} \end{figure} \end{document}