\documentstyle[pre,preprint,aps]{revtex} \tighten \begin{document} \title{Polarization patterns in Kerr Media} \author{Miguel Hoyuelos, Pere Colet, Maxi San Miguel and Daniel Walgraef\cite{Daniel}} \address{Instituto Mediterr\'aneo de Estudios Avanzados, IMEDEA (CSIC-UIB), Campus Universitat Illes Balears, E-07071 Palma de Mallorca, Spain.\cite{www}} \date{March 18, 1998} \maketitle \begin{abstract} We study spatiotemporal pattern formation associated with the polarization degree of freedom of the electric field amplitude in a mean field model describing a Kerr medium in a cavity with flat mirrors and driven by a coherent plane-wave field. We consider linearly as well as elliptically polarized driving field and situations of self-focusing and self-defocusing. For the case of self-defocusing and linearly polarized driving field there is a stripe pattern orthogonally polarized to the driving field. Such pattern changes into an hexagonal pattern for elliptically polarized driving field. The range of driving intensities for which the pattern is formed shrinks to zero with increasing ellipticity. For the case of self-focusing, changing the driving field ellipticity leads from a linearly polarized (for linearly polarized driving) to a circularly polarized hexagonal pattern (for circularly polarized driving). Intermediate situations include a modified Hopf bifurcation at finite wave number leading to a time dependent pattern of deformed hexagons and a codimension two Turing-Hopf instability resulting in an elliptically polarized stationary hexagonal pattern. Our numerical observations of different spatiotemporal structures are described by appropriate model and amplitude equations. \end{abstract} \pacs{PACS numbers:42.65-k,42.65Sf,47.54+r} \section{Introduction} Spatiotemporal patterns in the transverse direction of an optical field have now been widely studied theoretically and experimentally \cite{Chaossolfract}. In particular, the non linear optical configuration of a thin slice of Kerr material with a single feedback mirror analyzed in \cite{Alessandro} is the basis of many results recently obtained in this field. Studies of optical pattern formation share a number of aspects and techniques with general investigations of pattern formation in other physical systems \cite{CrossHoh}, but they also have specific features such as the role of light diffraction. A special feature of light patterns comes from the vectorial degree of freedom associated with the polarization of the light electric field amplitude. A vectorial degree of freedom also appears in recent studies of two-component Bose-Einstein condensates \cite{busch} modeled by coupled nonlinear Schr\"odinger equations. Consideration of this degree of freedom opens the way to study a rich variety of {\it vectorial} spatiotemporal phenomena. However, in many studies of optical pattern formation this extra degree of freedom has not been taken into account. Those studies correspond to situations in which a linear polarization of light is well stabilized. We will refer to these situations of frozen polarization as the ``scalar case''. An early study of polarization dynamical instabilities in nonlinear optics is due to Kitano et al. \cite{kitano}. For a review on polarization instabilities and multistability see \cite{zheludev}. Space independent polarization instabilities have been also studied in lasers \cite{puccioni,abraham,matlin,serrat}. More recently, vectorial patterns associated with polarization instabilities have been considered in lasers \cite{maxiPRL,gil93,toniPRL,sti,CO2Meucci,VCSELs,prati} as well as in non linear passive optical media. For this latter case, new type of vectorial instabilities have been predicted for cavity \cite{geddes1} or single feedback mirror \cite{Scroggie96} systems. Several experiments in cells with alkaline vapors have been reported either without cavity \cite{Tam77,Gahl} or with a single feedback mirror \cite{Grynberg94,Aumann97}. Dynamically evolving patterns produced in a cell of rubidium vapor with counterpropagating beams have also been studied experimentally \cite{Maitre95}. Within this context of recent experimental studies we address in this paper several aspects of polarization transverse patterns and pattern dynamics in passive optical systems presenting a systematic study of a model system. Pattern formation in non linear cavities for the scalar case was already considered in \cite{MoloneyKerr}. A prototype simple model which has been very useful for the understanding of pattern formation in this case is a mean field model describing a Kerr medium in a cavity with flat mirrors and driven by a coherent plane-wave field \cite{lugi,firt}. This model was extended to take into account the polarization degrees of freedom in \cite{geddes1,corrkerr}. Even if a Kerr material model does not give a faithful description of alkali vapors, it shares with it some basic polarization mechanisms of pattern formation. In addition, the relative simplicity of the model in \cite{geddes1} makes it worthwhile to study it in depth as a general prototype model for the basic understanding of vectorial patterns. We undertake here such study going beyond the situations already considered in \cite{geddes1}. The study in \cite{geddes1} was limited to the case in which the driving field is linearly polarized. Allowing for elliptically polarized driving field, as we do here, gives rise to a rich variety of new phenomena. In addition, the role of elliptically polarized homogeneous solutions in a number of pattern forming instabilities is discussed in detail. Those solutions already exist in the case of linearly polarized driving field. Our study involves a combination of linear stability analysis, numerical simulations and amplitude and model equations. Guided by the results of linear stability analysis we search numerically for different spatiotemporal structures. The general features of these structures are then shown to be described by model and amplitude equations which are justified by general arguments of symmetry, form of the linear instability and identification of relevant non linear couplings. The paper is organized as follows: In Sec. \ref{model} we describe the model we are considering, its spatially homogeneous solutions and general properties of the stability analysis of these states. In Sec. \ref{linearpolarization} we consider the case of linearly polarized driving field, while in Sec. \ref{ellipticalpolarization} we discuss the case of elliptically polarized driving field. Two particular situations of this last case are considered in the two following sections. Section \ref{Blinking Hexagons} is devoted to describe the deformed dynamical hexagons occurring in a modified Hopf bifurcation and Sec. \ref{cod2} discusses the Turing-Hopf codimension two bifurcation. A summary of results and their connection with related studies is given in Sec. \ref{results}. Finally, some general concluding remarks are given in Sec. \ref{conclusions}. \section{Description of the model, reference steady states and stability analysis} \label{model} The system we consider is a Fabry-P\'erot or ring cavity filled with an isotropic Kerr medium. The cavity is driven by an external input field of arbitrary polarization. The situation in which the polarization degree of freedom of the electromagnetic field is frozen was first considered by L. A. Lugiato and R. Lefever \cite{lugi,firt}. Geddes et al \cite{geddes1} generalized the model of \cite{lugi} to allow for the vector nature of the field. Their description of this system is given by a pair of coupled equations for the evolution of the two circularly polarized components of the field envelope $E_+$ and $E_-$, defined by $$ E_\pm = \frac{1}{\sqrt{2}}(E_x \pm iE_y). $$ For an isotropic medium, the equations are \begin{eqnarray} \frac{\partial E_\pm}{\partial t} &=& -(1 + i\eta \theta) E_\pm + i a \nabla^2 E_\pm + E_{0\pm} \nonumber \\ && + i\eta [ A |E_\pm|^2 + (A+B) |E_\mp|^2] E_\pm , \label{1} \end{eqnarray} where $E_{0\pm}$ are the circularly polarized components of the input field, $\eta = +1 \ (-1)$ indicates self-focusing (self-defocusing), $\theta$ is the cavity detuning, $a$ represents the strength of diffraction and $\nabla^2$ is the transverse Laplacian. $A$ and $B$ are parameters related to the components of the susceptibility tensor. As we are considering here an isotropic medium, $A + B/2 = 1$ ($B \le 2$). The case in which $E_{0+}=E_{0-}$ was considered in reference \cite{geddes1} which corresponds to a linearly polarized input field. Here we consider an input field with arbitrary ellipticity $\chi$ defined as \begin{eqnarray} E_{0+} &=& \sqrt{I_0} \cos(\chi/2) , \nonumber \\ E_{0-} &=& \sqrt{I_0} \sin(\chi/2) , \label{2} \end{eqnarray} where $I_0$ is the intensity of the input field. We consider that $E_{0+}$ and $E_{0-}$ are real. This assumption fixes the main axis of the ellipse in the X or Y direction. Note that $\chi=\pi/2$ corresponds to linear polarization along the X axis and that ellipticity increases when the value of $\chi$ is changed away from this value. $\chi=0$ ($\chi=\pi$) corresponds to a right-handed (left-handed) circularly polarized input field. Finally, $\chi=-\pi/2$ corresponds to linear polarization along the Y axis. The intensities of the circularly polarized components of the input field are $I_{0+} = I_0 \cos^2(\chi/2)$ and $I_{0-} = I_0 \sin^2(\chi/2)$. We note that the case of scalar field considered in \cite{lugi,firt} is formally recovered from (\ref{1}) for a circularly polarized input. The circular component of the field which is not excited by the input field decays to zero and the equation for the other component coincides with the one in \cite{lugi} up to a rescaling of the field amplitude. The steady state homogeneous solutions of Eq. (\ref{1}) are reference states from which transverse patterns emerge as they become unstable. These patterns are described in the following sections for different situations. The steady state homogeneous solutions $E_{s\pm}$ are given by the implicit equation \begin{equation} E_{0\pm} = E_{s\pm} \left\{ 1 - i\eta\left[ \left(1 - {B \over 2} \right) I_{s\pm} + \left(1 + {B \over 2} \right) I_{s\mp} - \theta \right] \right\} , \label{2.5} \end{equation} where $I_{s\pm}= |E_{s\pm}|^2$. For the intensities we have \begin{equation} I_{0\pm} = I_{s\pm} \left\{ 1 + \left[ \left(1 - {B \over 2} \right) I_{s\pm} + \left(1 + {B \over 2}\right) I_{s\mp} - \theta \right]^2 \right\}. \label{3} \end{equation} This gives a pair of coupled cubic polynomials in $I_{s+}$ and $I_{s-}$. Solving for $I_{s+}$ and $I_{s-}$ leads to a polynomial of degree nine from which it is not possible, in principle, to find an analytical expression. For the particular case of linearly polarized input field, Eq. (\ref{3}) admits symmetric ($I_{s+}=I_{s-}=I_s$) and asymmetric ($I_{s+} \neq I_{s-}$) solutions. The symmetric solution corresponds to linearly polarized output light, while the asymmetric is elliptically polarized. For the symmetric solution Eq. (\ref{3}) reduces to the single equation \cite{footnote1} \begin{equation} I_0/2 = I_s(1 + (2I_s - \theta)^2), \label{symmetricsol} \end{equation} which gives an implicit formula for $I_s$. As it is well known, Eq. (\ref{symmetricsol}) implies bistability for $\theta > \sqrt{3}$. We will restrict our analysis to non-bistable regimes, i.e. $\theta <\sqrt{3}$. The asymmetric solution is obtained from the general Eq. (\ref{3}). This solution breaks the ($+$,$-$) symmetry of the problem and is degenerate. There is one solution with $I_{s+}>I_{s-}$ and a second equivalent solution in which $I_{s+}$ and $I_{s-}$ are interchanged. The asymmetric solution only exists for values of $I_0$ greater than a threshold value for which $I_{s+}=I_{s-}=I'$. An example of the symmetric and asymmetric solutions for linearly polarized input is given in Fig. \ref{fig1}. The value of $I'$ is given by \begin{equation} I' = \frac{\theta(B-2) + \sqrt{\theta^2B^2+4(B-1)}}{4(B-1)}. \end{equation} For circularly polarized input light, for example $\chi=0$, Eq. (\ref{3}) reduces to \begin{equation} I_0 = I_{s+}\left\{1 + \left[ \left(1 - {B \over 2}\right) I_{s+} - \theta \right]^2 \right\} , \label{circpolarizedsol} \end{equation} and $I_{s-}=0$. It is clear from Eqs. (\ref{symmetricsol}) and (\ref{circpolarizedsol}) that the solution for circularly polarized input is the same as the symmetric solution for linearly polarized input, up to a rescaling of the intensities. An elliptically polarized input breaks the ($+$,$-$) symmetry of the system. The symmetric solution (\ref{symmetricsol}) found for linearly polarized input no longer exists. Instead there is a single asymmetric solution (\ref{3}) which favors the ellipticity of the input field. When the ellipticity of the input field is decreased, this solution approaches the asymmetric solution obtained for linearly polarized input. An example of this homogeneous elliptically polarized solution, obtained from (\ref{3}), is shown in Fig. \ref{fig3a} for various values of the ellipticity of the input field. This solution favors $I_{s+}$ and an equivalent solution favoring $I_{s-}$ is found when the ellipticity is changed from $\chi$ to $\pi-\chi$. Basic features of the stability of the steady state homogeneous solutions can be analyzed by considering the evolution equations for perturbations $\psi_{\pm}$ defined by \begin{equation} E_{\pm} = E_{s\pm}[1+ \psi_{\pm}] . \label{perturba} \end{equation} From Eqs. (\ref{1}) and (\ref{perturba}) we find \begin{eqnarray} \partial_t \psi_{\pm} &=& -\left[1+i\eta \left(\theta - S \pm {BR\over 2}\right) - ia \nabla^2 \right]\psi_{\pm}\nonumber \\ && +i\eta {S\over 2}\left[\left(1 - {B \over 2}\right) (\psi_{\pm}+\psi_{\pm}^*+\vert\psi_{\pm}\vert^2) + \left(1 +{B \over 2}\right) (\psi_{\mp}+\psi_{\mp}^*+\vert\psi_{\mp}\vert^2) \right] (1+\psi_{\pm})\nonumber \\ &&\pm i\eta {R \over 2} \left[\left(1 - {B \over 2}\right) (\psi_{\pm}+\psi_{\pm}^*+\vert\psi_{\pm}\vert^2) -\left(1 + {B \over 2}\right) (\psi_{\mp}+\psi_{\mp}^*+\vert\psi_{\mp}\vert^2) \right] (1+\psi_{\pm}) , \label{psimasmenos} \end{eqnarray} with \begin{equation} I_{s\pm}={1\over 2}( S \pm R ) . \label{imasmenos} \end{equation} The parameter $R$ measures the deviation from a symmetric solution vanishing for linearly polarized solutions. It is convenient to make a change of variables to the following basis \cite{geddes1}: \begin{equation} \Sigma = \left \{ \begin{matrix} {\sigma_1\cr \sigma_2\cr \sigma_3\cr \sigma_4}\end{matrix} \right \} = \left \{ \begin{matrix} {\Re(\psi_+ + \psi_-)\cr \Im(\psi_+ + \psi_-)\cr \Re(\psi_+ - \psi_-)\cr \Im(\psi_+ - \psi_-)}\end{matrix} \right \} . \end{equation} In this basis, which emphasizes the role of symmetric ($\psi_+=\psi_-$) and anti-symmetric ($\psi_+=-\psi_-$) modes, Eq. (\ref{psimasmenos}) may be written as: \begin{equation} \label{vectevol} \partial_t \Sigma = L \Sigma + N_2(\Sigma\vert\Sigma) + N_3(\Sigma\vert\Sigma\vert\Sigma) , \end{equation} where the linear matrix (in Fourier space) is \begin{equation} L = \left( \begin{array}{cccc} -1 & -\eta (S-\theta_k) & 0 & \eta BR/2 \\ \eta(3S-\theta_k) & -1 & -\eta(B/2-2)R & 0 \\ 0 & \eta BR/2 & -1 & -\eta(S-\theta_k) \\ -3\eta BR/2 & 0 & \eta(S(1-B)-\theta_k) & -1 \end{array} \right) , \label{linmatrix} \end{equation} with \begin{equation} \theta_k=\theta+\eta ak^2. \end{equation} The structure of the linear matrix is particularly simple for a symmetric solution $R=0$ \cite{geddes1}: $L$ becomes a matrix with $2\times 2$ blocks in which the symmetric and antisymmetric modes are decoupled. As a consequence, the linear instabilities lead to the growth of either a symmetric or an antisymmetric mode (see Sec. \ref{linearpolarization}). The eigenvalues $\lambda$ of $L$ are \begin{eqnarray} \lambda_{1,2} &=& -1 \pm \sqrt{(\theta_k-3S)(S-\theta_k)} , \nonumber\\ \lambda_{3,4} &=& -1 \pm \sqrt{(\theta_k+(B-1)S)(S-\theta_k)} . \label{lambdasymmetric} \end{eqnarray} In the general case of elliptically polarized input, the linear unstable modes are not purely symmetric or antisymmetric. The general expression for the 4 independent eigenvalues of $L$ are \begin{eqnarray} &&\lambda_{1,2,3,4} = -1 \pm {1 \over 2}\sqrt{f_1 \pm \sqrt{f_2}} , \nonumber \\ &&f_1 = (2B-8)S^2 - 2\theta_k(B-6)S - 2B(B-1)R^2 - 4\theta_k^2 , \nonumber \\ &&f_2 = 4(B+2)^2S^2(S-\theta_k)^2 - 4B(5B^2-16B+20)R^2S^2 \nonumber \\ & & \; \; \; \; \; \; + 12\theta_kB(B-2)(B-6)R^2S + B^2(B+2)^2R^4 \nonumber \\ & & \; \; \; \; \; \; + 32 B\theta_k^2(B-2)R^2. \label{10} \end{eqnarray} The different eigenvalues correspond to the four different combinations of plus and minus signs in the square roots. Replacing the values of $R$ and $S$, the eigenvalues are given as functions of the steady state intensities $I_{s+}$ and $I_{s-}$. Their dependence on $\eta$ is implicit in $\theta_k$. A given homogeneous steady state solution ($I_{s+}$,$I_{s-}$) becomes unstable when the real part of one eigenvalue becomes positive. These instabilities are described in detail in Sec. \ref{linearpolarization} for linearly polarized input and in Sec. \ref{ellipticalpolarization} for the general case of elliptically polarized input. The non linearities in (\ref{vectevol}) include quadratic $N_2(\Sigma\vert\Sigma)$ and cubic terms $N_3(\Sigma\vert\Sigma\vert\Sigma)$. The structure of these terms also gives some general information on the nature of the instabilities. In particular, if the quadratic non linearity $N_2(\Sigma\vert\Sigma)$ does not vanish, one expects the formation of hexagonal patterns instead of stripes. In addition, a stationary instability (which corresponds to a purely real eigenvalue becoming positive) is expected to be subcritical. For the symmetric solution ($R=0$), the quadratic non linearity is given by \begin{equation} N_2^S(\Sigma\vert\Sigma) = {\eta S\over 2}\left \{ \begin{matrix} {B \sigma_3 \sigma_4 - 2\sigma_1 \sigma_2\cr 3\sigma_1 \sigma_1 + \sigma_2 \sigma_2 +(1-B) \sigma_3 \sigma_3 + \sigma_4 \sigma_4 \cr B\sigma_2 \sigma_3 - 2\sigma_1 \sigma_4\cr 2(1-B)\sigma_1 \sigma_3 - B\sigma_2\sigma_4} \end{matrix} \right \} \,. \label{n1symmetric} \end{equation} Inspection of $N_2^S(\Sigma\vert\Sigma)$ identifies that quadratic non linearities, and therefore hexagonal pattern formation, are only expected for an instability of the symmetric mode. In this case, the critical modes are linear combinations of the modes $\sigma_1$ and $\sigma_2$, and $N_2^S(\Sigma\vert\Sigma)$ plays an important role since the first two components contains products of two unstable modes. Alternatively, if an asymmetric mode becomes unstable, the critical modes are linear combinations of $\sigma_3$ and $\sigma_4$. The third and fourth components of the vector $N_2^S(\Sigma\vert\Sigma)$ contain products of one stable and one unstable mode, but no products of two unstable modes. The adiabatic elimination of the stable modes yields quadratic terms involving two unstable modes, but these are terms of higher order and can be neglected. For an elliptically polarized solution, there is an additional contribution to the quadratic non linearity $N_2$; we find \begin{equation} N_2(\Sigma\vert\Sigma) = N_2^S(\Sigma\vert\Sigma) + {\eta R \over 4} M_2(\Sigma\vert\Sigma), \label{n1} \end{equation} where \begin{equation} M_2(\Sigma\vert\Sigma) = \left \{ \begin{matrix} {2B\sigma_1 \sigma_4 - 4\sigma_2 \sigma_3 \cr 2(4-B)\sigma_1 \sigma_3 +4\sigma_2 \sigma_4 \cr 2B\sigma_1 \sigma_2 - 4\sigma_3 \sigma_4 \cr -3B\sigma_1 \sigma_1 +(4-B)\sigma_3 \sigma_3 -B(\sigma_2 \sigma_2 +\sigma_4 \sigma_4)} \end{matrix} \right \} \,. \label{m1} \end{equation} Therefore, even for a purely asymmetric unstable mode there are important quadratic contributions which involve the unstable modes ($\sigma_3 \sigma_3$, $\sigma_3 \sigma_4$ and $\sigma_4 \sigma_4$), and hexagonal pattern formation is generally expected. \section{Linearly polarized input field} \label{linearpolarization} In this section we discuss transverse polarization patterns in the case of linearly polarized input field, which we take to be X-polarized. We have found (see Fig. \ref{fig1}) two types of homogeneous steady state solutions: a symmetric solution that is also X-polarized and an asymmetric one. The marginal stability curve for each solution is obtained from the eigenvalues of the matrix $L$ given in Eqs. (\ref{lambdasymmetric}) and (\ref{10}). We first consider the stability properties of the solutions with respect to homogeneous perturbations. The symmetric solution becomes unstable for a zero wave number perturbation ($k=0$) for $I_s = I'$. This is the point where the asymmetric solution appears. For $I_s>I'$, the asymmetric solution is stable with respect to homogeneous perturbations. Finite wave number perturbations destabilize the symmetric solution for $I_s < I'$ and the asymmetric solution for $I_s > I'$. In Fig. \ref{fig2} we plot marginal stability curves for $\theta=1$ as a function of $\eta ak^2$, so that positive values of this parameter correspond to self-focusing and negative values to self-defocusing. Figure \ref{fig2}(a) shows the marginal stability curve for the symmetric solution \cite{geddes1}. For the symmetric solution, the shape of the marginal stability curves is, in fact, the same for any value of the detuning $\theta$. This is because the eigenvalues $\lambda_i$ given by Eq. (\ref{lambdasymmetric}) depend only on $\theta_k$ and $S=2I_s$, so a change in the value of $\theta$ is equivalent to a displacement of the origin of $\eta ak^2$ (vertical dashed line) by the same amount. The vertical dashed line separating the self-focusing and self-defocusing cases moves to the right if the detuning $\theta$ is increased, and it intersects the left corner of region I for $\theta=\sqrt{3}$ (the value of $\theta$ beyond which there is bistability). Since in this paper we are only considering the non-bistable regime, the vertical dashed line is always situated to the left of region I. Figure \ref{fig2}(b) shows marginal stability curves for the asymmetric solution which merge continuously with the marginal curves of the symmetric solution for $I_sI_{s-}$. Since the quadratic non linearities are proportional to $R$, the sign of $R$ should determine if the hexagons are of the $0$- or $\pi$-type \cite{Aumann97,shoka,Ackeman95}. As the dynamical evolution of the perturbations $\psi_+$ and $\psi_-$ is associated with $\pm R$ (Eq. (\ref{psimasmenos})), we find the opposite type of hexagons for $|E_+|^2$ and $|E_-|^2$. Changing the ellipticity $\chi$ to $\pi-\chi$ induces a transition, for a given circularly polarized component of the field, from one type of hexagons to the other, similarly to what has been reported in Ref. \cite{Aumann97}. We also note that, near threshold, the hexagonal pattern looks different if we consider the $X$ or $Y$ components of the field. In Fig. \ref{fig6b} we plot $|E_x|^2$ and $|E_y|^2$, for which we find hexagonal pattern of the ``black eyes'' type. Similar patterns have been observed in chemical systems \cite{ouyang}. Here they arise because of the superposition of $E_+$ and $E_-$. In the self-focusing case (positive part of the horizontal axis of Fig. \ref{fig3b}), and for small ellipticity, the islands and tongues of instability of the homogeneous solution are obtained continuously from the ones for the asymmetric solution of linearly polarized input field (Fig. \ref{fig2} (b)). The island I in Fig. \ref{fig3b} (a)) corresponds to a stationary instability of the modes which by continuity go to the symmetric modes when $R\rightarrow 0$. As there are quadratic terms in the amplitude equation, an hexagonal pattern is formed, similarly to the case of linearly polarized input field. The tongues II and III are far away from the threshold for instability of the homogeneous solution and are plotted in this figure only to display how they move as we increase the ellipticity. As in the case of linear polarization of the input field, the tongue II is associated with a Hopf bifurcation and the tongue III with a stationary instability. Figure \ref{fig3b}(b) shows the marginal stability curve for $\chi=78^\circ$. When the input field intensity is increased starting from zero, we have, as usual, a first instability, of the Turing type, where an hexagonal structure emerges. The instability island I is now smaller and there is a window for $I_{s+}$ around 2, where the elliptically polarized homogeneous solution is stable. By further increasing the input field intensity, a second instability appears when the value of $I_{s+}$ crosses the instability threshold of the tongue II. The corresponding eigenvalues of the linear evolution matrix have non zero imaginary parts and cross the imaginary axis at finite wave number, so that this instability is a Hopf bifurcation with broken space translational symmetry. This situation is discussed in detail in Sec. \ref{Blinking Hexagons}. As we can see from the sequence of plots in Fig. \ref{fig3b}, the tongue II moves upwards and the tongue III downwards as the ellipticity of the input field is increased ($\chi$ is decreased). Beyond the island of instability I, the patterns that are expected to form depend crucially on the relative position of the Hopf instability (tongue II) and the stationary instability of tongue III. When the stationary instability is the first to appear on increasing the bifurcation parameter, steady spatial structures may be expected. If the Hopf bifurcation is the first to appear, however, one should obtain wavy spatiotemporal structures. If both instabilities are at the same level, one has a codimension 2 situation, as shown in Fig. \ref{fig3b}(c) for $\chi=73^\circ$. With respect to the situation of Fig. \ref{fig3b}(b), the instability tongue II has moved upwards and to the right, while tongue III has moved downwards until the instabilities associated with each of the two tongues take place at the same value of $I_{s+}$. There is now a large range of values of $I_{s+}$ between the island I and the two tongues for which the elliptically polarized homogeneous solution is stable. Increasing $I_{s+}$ from a value in this range, the homogeneous state has a codimension 2 bifurcation where steady and wavy modes should interact, ending with pure Turing, pure Hopf or mixed modes, according to their non linear interaction. This case is discussed in more detail in Sec. \ref{cod2}. For $\chi \rightarrow 0$, tongue II disappears, and tongue III is the only remaining region of instability. In Fig. \ref{fig3b}(d) we plot the marginal stability curve for a right-handed circularly polarized input field ($\chi=0$). As discussed in Sec. \ref{model}, this case is equivalent to the scalar case, already described in \cite{lugi,firt}. The steady state solution given by Eq. (\ref{circpolarizedsol}) is the same as the symmetric solution for linearly polarized input except for a rescaling of the intensities. After this rescaling, in the self-focusing case, the marginal stability curve is also the same as the one for linearly polarized input, and the same patterns are observed above threshold. In the self-defocusing case, however, we do not have any instability of the homogeneous solution. As stated in Sec. \ref{linearpolarization}, for linearly polarized input field the self-defocusing instability involves the asymmetric modes. Here, there is only one relevant component of the field, and there are not enough degrees of freedom for such an instability to occur. \section{Modified Hopf Bifurcation: Deformed Dynamical Hexagons} \label{Blinking Hexagons} In Fig. \ref{fig7d}, we plot the squared absolute value of $E_+$ and $E_-$ in the near and the far field for an input field intensity such that the value of $I_{s+}$ is slightly above the threshold of the Hopf instability (region II) shown in Fig. \ref{fig3b}(b). We can see that a distorted hexagonal structure appears for $E_+$ and $E_-$. The component $E_-$ is correlated with $E_+$ but has a lower intensity because $\chi < 90^\circ$ gives preference to $E_+$. The structure has a dynamical evolution as shown in Fig. \ref{fig7a}, where we plot four configurations of $|E_+|^2$ at different times. Inspection of the numerical results for the far field indicates that $E_+$ is dominated by a triad of 3 wave vectors with $|\vec k| = k_h$, while $E_-$ is dominated by the triad with opposite wave vectors. In addition, we observe that these 3 wave vectors are not equivalent, since two of them, which form an angle close to 90$^\circ$, carry a higher spectral power than the third one. The modes of $E_+$ with highest intensity are identified in Fig. \ref{fig7e:hexagons_modes}. The two equivalent wave vectors are labeled $\vec k_2$ and $\vec k_3$, and the third dominant wave vector is labeled $\vec k_1$. A basic feature of the dynamical evolution of the pattern can be understood considering the time evolution of each of these modes. The amplitude of any of these Hopf modes $|\vec k| = k_h$ evolves with a fixed frequency. The frequency has the same absolute value for all these modes but different signs, as indicated in Fig. \ref{fig7e:hexagons_modes}. In Fig. \ref{fig7b} we display, as an example, the phase of the mode $\vec k_2$. The frequency obtained from this plot, $\omega = 1.63$, coincides with the imaginary part of the critical eigenvalue associated with the Hopf instability. Since the numerical results are obtained slightly above the instability threshold, we may hopefully interpret them in the framework of reduced dynamics and amplitude equations for the unstable modes. Let us first recall that we are dealing with a Hopf instability with broken translational symmetry. The real part of the most unstable eigenvalues, $\lambda_1$ and $\lambda_2$, is plotted in Fig. \ref{fig7c}. These eigenvalues are complex conjugated for $k \simeq k_h$ and $\lambda_{1,2} (k_h) = \pm i\omega$. The field $E_\pm$ can be projected on the eigenvectors of the linear evolution matrix $L$ and using Eq. (\ref{perturba}) may be written as \begin{eqnarray} \left( \begin{array}{c} E_+ \\ E_+^* \\ E_- \\ E_-^* \end{array} \right) = \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) + \left( \begin{array}{c} E_{s+} \psi_+ \\ E_{s+}^* \psi_+^* \\ E_{s-} \psi_- \\ E_{s-}^* \psi_-^* \end{array} \right) = \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) \nonumber\\ + \sum_{\vec k} e^{i\vec k \vec r} \left( V_1 \sigma_{1,\vec k} + V_2 \sigma_{2,\vec k} + V_3 \sigma_{3,\vec k} + V_4 \sigma_{4,\vec k} \right), \end{eqnarray} where $V_i$ are the four eigenvectors, in Fourier space, and $\sigma_{i,\vec k}$ are the corresponding amplitudes. The usual procedure of reducing the dynamics to the dynamics of the unstable modes only, leads to the adiabatic elimination of the modes $V_3 \sigma_{3,k}$ and $V_4 \sigma_{4,k}$ because $\lambda_3$ and $\lambda_4$ are stable eigenvalues. Taking into account that the eigenvalues $\lambda_1$ and $\lambda_2$ are most unstable for $\vec k = \vec k_h$ with $|\vec k_h| = k_h$, we may write, close to the instability \cite{CrossHoh} \begin{eqnarray}\label{redfield} &&\left( \begin{array}{c} E_+ \\ E_+^* \\ E_- \\ E_-^* \end{array} \right) = \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) \nonumber \\ &&+ \sum_{\vec k_h} \left( V_1 \sigma_{1,\vec k_h} e^{i \vec k_h \vec r + i \omega t} + V_2 \sigma_{2,\vec k_h} e^{i \vec k_h \vec r - i \omega t} \right) + \dots . \end{eqnarray} The amplitudes, $\sigma_{1,\vec k_h}=\sigma_{1,\vec k_h}(\vec X,T)$ and $\sigma_{2,\vec k_h}=\sigma_{2,\vec k_h}(\vec X,T)$, only depend on the slow variables $\vec X=\varepsilon^{1/2}\vec r$ and $T=\varepsilon^{-1}t$, where $\varepsilon= {I_{s+}-I_{s+}^c \over I_{s+}^c}$ is the reduced distance to the instability threshold (we are using $I_{s+}$ as the bifurcation parameter and $I_{s+}^c$ is the critical value at the Hopf bifurcation). For each particular pattern, their evolution equations, or amplitude equations, may be derived with standard procedures \cite{daniel}. However, it is often convenient to study instead order parameter equations of the Swift-Hohenberg type. These equations reduce to the correct amplitude equations near the onset of instability but take care of the orientational degeneracy of the unstable wavectors, preserve the correct symmetries of the problem, allow the description of transitions between patterns of different symmetries and contain rapid spatiotemporal variations which may be important for pattern selection or transient dynamics \cite{CrossHoh,shoka,manneville}. In the present case, we consider a model order parameter dynamics of the type \begin{eqnarray}\label{SHOsc} \partial_t\sigma_{\pm} &=& [\varepsilon - \xi_h^2(k_h^2 +\nabla^2 )^2 \pm i\omega ]\sigma_{\pm} + v\sigma_{\mp}^2 \nonumber \\ && - (1 \pm i\beta)\sigma_{\pm}^2\sigma_{\mp} , \end{eqnarray} where the subscript $+$ and $-$ refer to the sign of the frequency, so that the complex $\sigma_{\pm}$ are proportional to the wave packet $\sigma_{1(2),\vec q}\exp (i\vec k \vec r \pm i\omega t)$, with $\vert \vec k\vert \simeq k_h$. These equations contain quadratic non linearities, as discussed in Sec. \ref{model}, and are thus equivalent to the equations describing oscillatory convection in hydrodynamic systems with no ``up-down" symmetry \cite{brand}. In this case, the authors obtain monoperiodic regular states of hexagonal symmetry that correspond to a two-dimensional standing wave. We will call these states pulsating hexagons. Our system has, however, the following originality. The eigenvalue $\lambda_2$ is real and only slightly negative for the modes $\sigma_{0}$ with a wavevector such that $\vert \vec k\vert = k_s \simeq \sqrt 2 k_h$ (see Fig. \ref{fig7c}). As $k_s \simeq \sqrt 2 k_h$, the modes $\sigma_{0}$ may be coupled with pairs of Hopf modes $\sigma_{\pm}$ with orthogonal wave vectors via quadratic resonances. Since the modes $\sigma_{0}$ are only slightly damped, one should incorporate them in the order parameter dynamics, which then becomes \begin{eqnarray}\label{SHgen} \partial_t\sigma_{\pm} &=& [\varepsilon - \xi_h^2(k_h^2 +\nabla^2 )^2 \pm i\omega ]\sigma_{\pm} + v_0\sigma_{\mp}^2 + v_1\sigma_{\pm}\sigma_{0} \nonumber\\ && + v_2 \sigma_{0}^2 - (1 \pm i\beta)\sigma^2_{\pm}\sigma_{\mp} -( \gamma \pm i\delta)\sigma^2_{0}\sigma_{\pm}\nonumber\\ \partial_t\sigma_{0} &=& [\mu - \xi_s^2(k_s^2 +\nabla^2 )^2 ]\sigma_{0} + \bar v_1 \sigma_{+}\sigma_{-} +\bar v_2 \sigma_{0}^2- \sigma_{0}^3 \nonumber\\ & & - u\sigma_{0}\sigma_{+}\sigma_{-} , \end{eqnarray} where $\mu<0$ is the linear damping of the homogeneous mode ($\mu \propto \lambda_2(k_s)$). The kinetic coefficients could be obtained numerically from Eq. (\ref{perturba}). However, this formidable task may be avoided, since we are mainly interested in generic dynamical behaviors, which essentially depend on their signs and orders of magnitude. Let us look for the possible asymptotic solutions of this system. Systems described by the dynamics (\ref{SHOsc}) have been mainly studied in one-dimensional geometries where the resulting patterns correspond to traveling or standing waves \cite{CrossHoh,haken,lomdahl}. These solutions are recovered here. Effectively, the uniform amplitude equations for critical unidirectional counterpropagating traveling waves corresponding to the modes $\sigma_0=0$, $\sigma_+ = L\exp (i\vec k_h\vec r +i\omega t) + R^*\exp (-i\vec k_h\vec r + i\omega t)$ and $\sigma_-= L^*\exp (-i\vec k_h\vec r - i\omega t) + R\exp (i\vec k_h\vec r - i\omega t)$ ($\vert \vec k_h\vert = k_h$) may be deduced from (\ref{SHgen}), and are \begin{eqnarray}\label{TW} \dot L &=& \varepsilon L - (1+i\beta )L(\vert L\vert^2 +2 \vert R\vert^2 )\nonumber\\ \dot R &=& \varepsilon R - (1-i\beta )R(\vert R\vert^2 +2 \vert L\vert^2) . \end{eqnarray} The non linear cross-couplings coefficients of the field equations are twice the self-coupling coefficients. In this situation, like in reaction-diffusion systems with scalar couplings \cite{shoka}, traveling waves are stable structures whereas standing waves are unstable. As we are considering two-dimensional systems, we have to study the stability of such waves versus modulations with wave vectors pointing in other directions. In the absence of coupling between $\sigma_{\pm}$ and $\sigma_0$, one should obtain pulsating hexagons, as in \cite{brand}. The coupling between oscillatory $\sigma_{\pm}$ and steady $\sigma_0$ modes may modify this picture, however. Effectively, let us consider the linear stability of a traveling wave defined by $\sigma_-=A \exp \left( i k_h{x+y\over\sqrt{2}} - i\omega_0 t \right) $, $\sigma_+= A^* \exp \left( -i k_h{x+y\over\sqrt{2}} + i \omega_0 t \right) $, with $\vert A\vert^2=\varepsilon$ and $\omega_0 =\omega -\beta\varepsilon$ which corresponds to a solution of (\ref{TW}) with $L=0$ and $R=A e^{i\beta \varepsilon t}$. (The wave vector direction is arbitrary, and we may choose this particular one, anticipating results of the following discussion. Based on the wave vector definition of Fig. \ref{fig7e:hexagons_modes}, $x$ and $y$ being the spatial coordinates in the plane, the right traveling wave just described corresponds to mode $A$.) Mode $A$ is quadratically coupled to the modes $B \exp \left( i k_h{x-y\over\sqrt{2}}- i \omega_0 t \right) $ and $D e^{ik_sy}$ and their complex conjugates. Taking $\sigma_-=A \exp \left( i k_h{x+y\over\sqrt{2}} - i \omega_0 t \right) + B \exp \left( i k_h{x-y\over\sqrt{2}}- i \omega_0 t \right)$, $\sigma_+= A^* \exp \left( -i k_h{x+y\over\sqrt{2}} + i \omega_0 t \right) + B^* \exp \left( -i k_h{x-y\over\sqrt{2}} + i \omega_0 t \right) $, and $\sigma_0 = D e^{ik_sy} + D^* e^{-ik_sy}$, the corresponding linearized amplitude equations are of the form \begin{eqnarray} \dot B &=& -\varepsilon B + v_1 A D^* - \dots \nonumber\\ \dot D &=& (\mu - u\varepsilon) D + \bar v_1 A B^* - \dots . \end{eqnarray} The characteristic equation of the corresponding evolution matrix is \begin{equation} s^2 - s((1+u)\varepsilon -\mu )+\varepsilon (u\varepsilon -\mu) - v_1\bar v_1\varepsilon = 0 , \end{equation} and the traveling waves are thus unstable for \begin{equation} \varepsilon < \varepsilon_c = {1\over u }(v_1\bar v_1 +\mu ) . \end{equation} If $v_1\bar v_1$ is negative, $\varepsilon_c < 0$ and unidirectional traveling waves are stable. This is, for example, the case in systems where the order parameter dynamics contains non linear couplings between the gradients of the field \cite{proctorjones}. Alternatively, if $v_1\bar v_1$ is positive, and we may suppose that this is the case here, $\varepsilon_c $ is positive when $\mu $ has sufficiently small absolute value (we recall that $\mu<0$). Hence, unidirectional traveling waves are unstable versus two-dimensional spatiotemporal patterns in the range $0<\varepsilon < \varepsilon_c$. We will now try to identify these patterns, taking into account the fact that the dynamics favors propagating waves and contains quadratic non linearities. Hence, we consider that the dynamics may be reduced to the dynamics of the modes A, B, C, D and their respective complex conjugates, as defined in Fig. \ref{fig7e:hexagons_modes}. Starting with mode A, modes B and D should be included in the description because A is unstable versus B and D. Mode C should be also included because of the coupling between A and B. The coupling between A and B also generates higher harmonics with frequency $-2\omega$ (vector $\vec k_5$ in Fig. \ref{fig7e:hexagons_modes}) which can be observed in the far field shown in Fig. \ref{fig7d}. However, since the dynamics of these harmonics is slaved to the dynamics of A and B, they are not considered here. The amplitude equations for modes A, B, C and D obtained from (\ref{SHgen}) taking into account quadratic non linearities between wavy modes are \begin{eqnarray} \partial_t A &=& \varepsilon A+4\xi_h^2(\vec k_2.\vec\nabla )^2A + v_1 DB + v_0C^*B^* e^{i(3\omega t +\kappa x)}\nonumber\\ &-&(1+i\beta)A(\vert A\vert^2 + 2 (\vert B\vert^2 +\vert C\vert^2))-(\gamma+i\delta)A\vert D\vert^2\nonumber\\ \partial_t B &=&\varepsilon B+ 4\xi_h^2(\vec k_3.\vec\nabla )^2B + v_1 AD^* + v_0 C^*A^* e^{i(3\omega t +\kappa x)}\nonumber\\ &-& (1+i\beta)B(\vert B\vert^2 +2 (\vert A\vert^2 + \vert C\vert^2))-(\gamma+i\delta)B\vert D\vert^2\nonumber\\ \partial_t C &=& \varepsilon C+ 4\xi_h^2(\vec k_1.\vec\nabla )^2C + v_0 A^*B^* e^{i(3\omega t +\kappa x)} \nonumber\\ &-& (1+i\beta)C(\vert C\vert^2 +2 (\vert B\vert^2 + \vert A\vert^2)) -(\gamma+i\delta)C\vert D\vert^2\nonumber\\ \partial_t D&=&\mu D+ 4\xi_s^2(\vec k_4.\vec\nabla )^2D + \bar v_1 AB^* - \dots , \end{eqnarray} where $\kappa =(1-\sqrt 2 )k_h $. $D$ may be adiabatically eliminated and $D\simeq -{\bar v_1 AB^*\over \mu}$. The mismatch between the modes A, B and C may be absorbed in their phases ($\phi_i \to \bar\phi_i - \kappa x/3$). Writing $I= R_I\exp \phi_I$ and defining the total phase as $\Phi = \bar\phi_A +\bar\phi_B+\bar\phi_{C} -3\omega t $, we finally get the following equations for uniform amplitudes and phases, up to cubic non linearities: \begin{eqnarray}\label{dyn} \partial_t R_A &=& \varepsilon_0 R_A+ v_1 R_BR_{C} \cos \Phi - R_A(R_A^2 +\bar\gamma R_B^2 +2 R_{C}^2)\nonumber\\ \partial_t R_B &=& \varepsilon_0 R_B+ v_1 R_AR_{C} \cos \Phi - R_B(R_B^2 +\bar\gamma R_A^2 +2 R_{C}^2)\nonumber\\ \partial_t R_{C} &=& \varepsilon_1 R_{C} + v_1R_AR_B\cos \Phi - R_{C}(R_C^2 + 2 (R_A^2+R_B^2))\nonumber\\ \partial_t\bar\phi_A &=& - v_0{R_BR_{C}\over R_A}\sin \Phi + O(\beta,\delta,R^2)\nonumber\\ \partial_t\bar\phi_B &=& - v_0{R_AR_{C}\over R_C}\sin \Phi + O(\beta,\delta,R^2)\nonumber\\ \partial_t\bar\phi_{C} &=& - v_0{R_AR_B \over R_{C}}\sin \Phi + O (\beta,\delta,R^2)\nonumber\\ \partial_t\Phi &=& -3\omega - v_1{2R_{C}^2+R_A^2 \over R_{C}}\sin \Phi + O (\beta,\delta,R^2) , \end{eqnarray} where $\varepsilon_0 = \varepsilon - {2\over 9}\xi_h^2(1-\sqrt 2 )^2k_h^4 \simeq \varepsilon - 0.038\xi_h^2k_h^2$, $\varepsilon_1 = \varepsilon - {4\over 9}\xi_h^2(1-\sqrt 2 )^2k_h^4 \simeq \varepsilon - 0.075\xi_h^2k_h^2$ and $\bar\gamma = 2 +{v_2\bar v_1 \over \mu}$. From these equations, it turns out that $A$ and $B$ are equivalent, $R_A = R_B \ne R_C$ and $\bar\phi_A - \bar\phi_B = \varphi$, where $\varphi$ is an arbitrary constant. We may choose $\varphi=0$ for simplicity. Except for sufficiently small $\omega$, where the system (\ref{dyn}) may admit fixed point solutions, this system of equations is expected to generate time-periodic solutions corresponding to pulsating deformed hexagons. Over a period, the mean value of the amplitude of the modes A and B should be equal while the mean value of the amplitude of C should be smaller ($\varepsilon_1 <\varepsilon_0$, and $\bar\gamma <\gamma$). In the absence of coupling with the $D$ mode ($\mu << 0$), one should recover the pulsating hexagons found by Deissler and Brand \cite{brand}. It is the particularity of this system to present a resonant interaction with a slowly evolving stable mode which induces the deformation of the hexagonal pattern. In general $A$ and $B^*$ modes need not make a particular angle for their quadratic resonance with a stationary mode. If they make an angle $\psi$ with the y-axis, they should be coupled with the stable $D$ mode (with $k_s \hat y = 2k_h\sin \psi \hat y$, where $\hat y$ is the unit vector along the Y-axis) in such a way that \begin{equation} \partial_t D(\psi) = [\mu - 4\xi_s^2(k_s^2- 4\sin^2\psi k_h^2)^2]D(\psi) + \bar v_1 AB^* +\dots . \end{equation} Hence, the adiabatic elimination of this mode leads to a renormalized coefficient $\bar\gamma$ which is minimum for $\sin \psi = k_s / 2k_h$, The angle $\psi$ is then half of the selected angle between $A$ and $B^*$ wave vectors. Changing the detuning $\theta$, the ratio between $k_s$ and $k_h$ can be tuned, which then changes the distortion of the hexagonal pattern. If one now wishes to reconstruct the field from the results of this weakly non linear analysis, one obtains from equations (\ref{redfield}) and (\ref{dyn}), at leading order, \begin{eqnarray}\label{reconstruct} \left( \begin{array}{c} E_+ \\ E_+^* \\ E_- \\ E_-^* \end{array} \right) &-& \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) = \sum_{\vec k_h} \left( V_1 \sigma_{1,\vec k_h} e^{i \vec k_h \vec r + i \omega t} + V_2 \sigma_{2,\vec k_h} e^{i \vec k_h \vec r - i \omega t} \right) + \dots \nonumber\\ &=& V_1 e^{i \omega t} \left( A^* e^{-i \vec k_2 \vec r} + B^* e^{-i \vec k_3 \vec r } + C^* e^{-i \vec k_1 \vec r} \right) + V_2 e^{-i \omega t} \left( A e^{i \vec k_2 \vec r} + B e^{i \vec k_3 \vec r} + C e^{i \vec k_1 \vec r } \right) + \dots \nonumber\\ &=& V_1 e^{ i \omega t } \left( 2R_A \cos {k_h y\over \sqrt 2} e^{-i {k_h x\over \sqrt 2} - i\bar\phi_A } + R_{C} e^{i k_h x - i\bar\phi_{C}}\right) \nonumber\\ && + V_2 e^{ -i \omega t} \left( 2R_A \cos {k_h y\over \sqrt 2}e^{i {k_h x\over \sqrt 2} + i\bar\phi_A} + R_{C} e^{-i k_h x + i\bar\phi_{C}}\right) + \dots . \end{eqnarray} From (\ref{dyn}), it may be seen that $\bar\phi_A$ and $\bar\phi_C $ are functions of $\Phi$. Since $\omega$ is finite, close to the instability threshold one may expand $\Phi$ around $3\omega t$ \cite{kuramoto}, and (\ref{reconstruct}) becomes \begin{eqnarray}\label{reconstructbis} \left( \begin{array}{c} E_+ \\ E_+^* \\ E_- \\ E_-^* \end{array} \right) - \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) & = & V_1 e^{ i \omega t } \left( 2R_A \cos {k_h y\over \sqrt 2} e^{-i {k_h x\over \sqrt 2} } + R_{C} e^{i k_h x } \right) \nonumber\\ && + V_2 e^{ -i \omega t} \left( 2R_A \cos {k_h y\over \sqrt 2}e^{i {k_h x\over \sqrt 2} } + R_{C} e^{-i k_h x }\right) + \dots \end{eqnarray} where $R_A$ and $R_B$ still contain time-dependent contributions of frequencies $3\omega, 6\omega, \dots $. The corresponding spatiotemporal patterns are thus different from pulsating hexagons, since, besides their deformation, they are built on a triad of traveling waves propagating in the $\vec k_1$, $\vec k_2$ and $\vec k_3$ (or $-\vec k_1$, $-\vec k_2$ and $-\vec k_3$) directions, leading to what we call deformed dynamical hexagons. These conclusions are in qualitative agreement with the numerical results presented in Figs. \ref{fig7d}, \ref{fig7a} and \ref{fig7b}, which tell us, furthermore, that the eigenvectors $V_1$ and $V_2$ should have dominant contributions to $E_+$ and $E_-$, respectively. Effectively, it appears that the dominant contributions to $E_-$ come from the modes $k_1$, $k_2$ and $k_3$ with frequency $-\omega$, while $E_+$ is built on the modes $k_1$, $k_2$ and $k_3$ with frequency $\omega$. We thus think that the main feature which determines the properties of the patterns presented in Figs. \ref{fig7d} and \ref{fig7a} is the fact that a constructive coupling occurs between a nearly marginal stationary mode and unstable oscillatory modes. Couplings between steady and oscillatory modes have already been shown to be able to induce subharmonic Hopf bifurcations in one-dimensional reaction-diffusion systems in codimension 2 situations \cite{Lima96,DeWit96}. Here, this particular coupling is allowed by the two-dimensional geometry of the system, which induces the bifurcation to spatiotemporal patterns with deformed hexagonal shape. \section{Codimension two Hopf and Turinglike instabilities} \label{cod2} In this section, we consider the situation shown in Figs. \ref{fig3a}(c) (steady state) and \ref{fig3b}(c) (marginal stability), corresponding to an ellipticity $\chi = 73^\circ$ and detuning $\theta=1$. We see from the marginal stability curve that there are two different wave numbers that become unstable at nearly the same value of the control parameter $I_{s+}$. Similar situations can be obtained changing simultaneously the ellipticity and the detuning. For example, for $\chi = 67^\circ$ and $\theta=0.6$ the same type of situation occurs (the relation between the two critical wave numbers changes but the qualitative results are not affected). In Fig. \ref{fig8b} we plot the unstable eigenvalues $\lambda_1$ and $\lambda_2$ as functions of the wave number $k$ near the instability threshold for $\chi = 67^\circ$ and $\theta=0.6$. In the plot of $\lambda_2$ we can see that the modes $k_h$ and $k_s$ become simultaneously unstable, and, furthermore, $\lambda_2(k_h)$ is complex, while $\lambda_2(k_s)$ is real. Hence, we have a codimension two situation where the oscillatory instability corresponding to a Hopf bifurcation with broken spatial symmetry and the stationary Turinglike instability are close together. We show in Fig. \ref{fig8a} numerical results for the near and the far field of $E_+$ at three different times during the transient following this codimension two bifurcation. These results display the competition between the Hopf and the static modes. At short times the Hopf modes dominate, but at long times a static hexagonal pattern is formed. In the vicinity of the codimension two point, the field variables may be written as: \begin{eqnarray} \left( \begin{array}{c} E_+ \\ E_+^* \\ E_- \\ E_-^* \end{array} \right) &=& \left( \begin{array}{c} E_{s+} \\ E_{s+}^* \\ E_{s-} \\ E_{s-}^* \end{array} \right) \nonumber \\ &&+ \sum_{\vec k_h}\left( V_1 \sigma_{1,\vec k_h} e^{i \vec k_h \vec r + i \omega t} + V_2 \sigma_{2,\vec k_h} e^{i \vec k_h \vec r - i \omega t} \right) \nonumber \\ &&+ \sum_{\vec k_s}V_3 \sigma_{0,k_s} e^{i \vec k_s \vec r } , \end{eqnarray} where $\vert \vec k_h\vert = k_h$ and $\vert \vec k_s \vert = k_s$ are the critical wave numbers associated with each of the instabilities mentioned before. This expansion is the generalization of Eq. (\ref{redfield}) to the codimension 2 situation, where one has to expand the fields on all the unstable modes of the problem, including here the Turinglike unstable mode. The amplitudes $\sigma_{1,\vec k_h}(\vec X,T)$, $\sigma_{2,\vec k_h}(\vec X,T)$ and $\sigma_{0,\vec k_s}(\vec X,T)$ only depend on the slow variables of the problem $\vec X=\varepsilon^{1/2}\vec x$ and $T=\varepsilon^{-1}t$. Furthermore, we define $\mu$ as the reduced distance to the stationary instability threshold, ($\mu = {I_{s+}-I_{s+}^c \over I_{s+}^c}$, where now $I_{s+}^c$ is the critical value of the bifurcation parameter at this Turinglike instability; $\mu$ is positive, contrary to the case discussed in the preceding section). The structure of the evolution equations for these amplitudes may easily be obtained using the symmetries of the problem \cite{daniel}. Contrary to the case discussed in Sec. \ref{Blinking Hexagons}, in the present situation $k_s < k_h$ so that there should be no quadratic couplings between oscillatory hexagonal planforms and steady modes. In the absence of such resonances, oscillatory and stationary modes are first coupled through cubic non linearities. For example, the dynamics of a pair of unidirectional counterpropagating traveling waves of amplitudes $A$ and $B$, coupled to an arbitrary number of steady modes of amplitudes $R_i$ correspond to the following coupled Ginzburg-Landau and Swift-Hohenberg equations: \begin{eqnarray}\label{hopfturing} \partial_t A &=& GL_A(\varepsilon ,A,B) -g (1+id)A\Sigma_i\vert R_i\vert ^2\nonumber\\ \partial_t B &=& GL_B^*(\varepsilon ,A,B) -g (1-id)B\Sigma_i\vert R_i\vert ^2\nonumber\\ \partial_t R_i &=& SH_i({R_j}) -w R_i(\vert A\vert ^2 + \vert B\vert ^2) \end{eqnarray} where, in the absence of mean flow or group velocity, $GL_A(\varepsilon ,A,B)=\varepsilon A+(1+i\alpha )\partial_x^2A - (1+i\beta )A(\vert A\vert^2+ \gamma \vert B\vert ^2)$, with $\vec k_h \Vert \hat x$. $SH_i({R_j})$ represents the generic evolution terms of the Swift-Hohenberg type close to a Turinglike instability. In the absence of ``up-down" symmetry, as it is the case here, quadratic non linear couplings between stationary modes are important. The corresponding dynamical operator may then be written, for an arbitrary triad of modes, as ($i,j = 1,2,3$): \begin{eqnarray}\label{hex} SH_1({R_j}) &=&\mu R_1 + (\vec k_{s1}\vec \nabla )^2 R_1+ vR_2^*R_3^* -\vert R_1\vert ^2 R_1 \nonumber\\ && - u (\vert R_2\vert ^2 +\vert R_3\vert^2)R_1\nonumber\\ SH_2({R_j}) &=&\mu R_2 + (\vec k_{s2}\vec \nabla )^2 R_2+ vR_1^*R_3^* -\vert R_2\vert ^2 R_2 \nonumber\\ && - u (\vert R_1\vert ^2 +\vert R_3\vert^2)R_2\nonumber\\ SH_3({R_j}) &=&\mu R_3 + (\vec k_{s3}\vec \nabla )^2 R_3+ vR_1^*R_3^* -\vert R_3\vert ^2 R_3 \nonumber\\ && - u (\vert R_2\vert ^2 +\vert R_1\vert^2)R_3 , \end{eqnarray} with $\vec k_{s1}+\vec k_{s2}+\vec k_{s3}=0$. Since we are dealing with a system with scalar non linear couplings, $\gamma$, $u$, $g$ and $w$ should be larger than one. Codimension 2 situations have been extensively studied in one-dimensional reaction-diffusion systems where Turing and zero-wave number Hopf instability thresholds are close together \cite{shoka,Lima96,DeWit96}. According to the non linear couplings between unstable modes, the resulting patterns may be pure Turing, pure Hopf, or mixed mode patterns. We are considering here two-dimensional geometries and a Hopf bifurcation with finite wave vector. The Eqs. (\ref{hopfturing}) and (\ref{hex}) admit as steady states pure stripes of amplitude $\sqrt{\mu}$ or hexagonal planforms, where $R_i = R_0$ and $A=B=0$, with $$ R_0 = {v + \sqrt{v^2 +4(1+2u)\mu}\over 2(1+2u)} . $$ Stripes are unstable versus hexagons for $\mu < v^2/(u-1)^2$, which is expected to be the case here since $v$ is finite and $\mu \ll 1$. They also admit as asymptotic solutions traveling waves (of amplitude $\vert A\vert = \vert A_0\vert = \sqrt \varepsilon$, $B=0$, R$=0$ or $\vert B\vert = \vert B_0\vert = \sqrt \varepsilon$, $A=0$, R$=0$) and mixed modes. The stability of hexagonal patterns versus wavy modes may be studied with the following linearized equations: \begin{eqnarray} \partial_t A &=& \varepsilon A+(1+i\alpha )\partial_x^2A -3g (1+id)\vert R_0\vert ^2 A\nonumber\\ \partial_t B &=& \varepsilon B+(1+i\alpha )\partial_x^2B -3g (1+id)\vert R_0\vert ^2 B . \end{eqnarray} The result is that hexagonal planforms are stable for \begin{equation} \varepsilon < {3g \over 2(1+2u)}(v + \sqrt{v^2 +4(1+2u)\mu}) , \end{equation} which is the case to be expected here since $\varepsilon$ is small and $v$ finite. Since we suppose that $\gamma$ is larger than 1, the pure wavy solutions of Eq. (\ref{hopfturing}) are traveling waves of amplitude $\sqrt{\varepsilon}$. The evolution of the steady modes $R_i$ in the presence of pure traveling waves ($|A_0| = \sqrt{\varepsilon}$, $|B_0|=0$ or $|B_0| = \sqrt{\varepsilon}$, $|A_0|=0$) is given by \begin{equation} \partial_t R_i= SH_i({R_j}) -w \varepsilon R_i . \end{equation} Traveling waves are then linearly stable if $\mu - w \varepsilon < 0$, which should be the case here. Hence, for the situation discussed in this section, hexagonal and wave patterns are expected to be simultaneously stable. The condition for the existence of mixed hexagon-traveling wave modes is found to be \begin{equation} 1+2u > 3g w \end{equation} and is not expected to be satisfied in reaction-diffusion dynamics with scalar non linear couplings. It is important to note that the Hopf bifurcation is supercritical while the Turinglike transition to hexagons is subcritical. As a result of their supercriticality, wavy patterns grow first. Although these patterns are linearly stable versus stationary hexagonal planforms, the dynamics of the latter present destabilizing quadratic non linearities. The result is that the hexagonal patterns grow faster and finally take over. Since steady hexagons are stable versus waves, they should thus be the final pattern, although wave patterns may appear as transients during the first part of the evolution. This is indeed what is observed in the numerical simulations of Fig. \ref{fig8a}: at time $t=50$ there is a dominance of Hopf modes with arbitrary orientations with weakly excited Turing modes, as clearly seen in the far field. At late times ($t=1700$) the Hopf modes have lost the competition and only Turing modes giving an hexagonal pattern survive. Complicated dynamical competition occurs at intermediate times. In Fig. \ref{fig_compet}, we show the integrated power of Hopf and Turing modes as functions of time. The Hopf modes are seen to grow supercritically first, but eventually the subcritical Turing modes grow faster until they overcome the Hopf modes. \section{Summary and discussion of results} \label{results} In previous sections we have reported a rich phenomenology for the broad range of values of the cavity detuning ($2/B - 1 < \theta < \sqrt{3}$) that we have explored. We summarize here these results and discuss their connection to other related studies. We have revisited in Sec. \ref{linearpolarization} the case of linearly polarized driving field for self-focusing as well as self-defocusing situations. For the self-defocusing case there is an instability leading to a stripe stationary pattern which is orthogonally polarized to the driving field \cite{geddes1}. However, increasing the intensity of the driving field the pattern disappears leading to a final homogeneous elliptically polarized pattern. The transient dynamics involves the spatial coexistence of two equivalent elliptically polarized homogeneous states separated by moving interfaces, like in a process of phase separation dynamics \cite{gunton}. Spatial coexistence of domains of different structures has been reported in \cite{Residori96} for a liquid crystal light valve with rotated feedback loop, while stationary spatial coexistence of circularly polarized states has been reported in alkali vapors driven by a linearly polarized field in a single mirror system \cite{Grynberg94} and in cells without mirrors \cite{Tam77,Gahl}. For a linearly polarized driving field and a self-focusing situation, there is an instability leading to an hexagonal pattern with the same polarization of the driving field \cite{geddes1}. This is the same process as in a scalar model \cite{lugi,firt}. When the intensity of the driving field is increased the hexagonal pattern is destabilized \cite{firthsolfract}. We have observed that a further increase of the driving field intensity leads to a complicated spatiotemporal dynamics with bursting spots that create propagating circular waves. A related phenomenon has been reported in a model which includes the dynamics of atomic variables \cite{Berre97}. For an elliptically polarized driving field (Sec. \ref{ellipticalpolarization}) and a self-defocusing situation, the stripe pattern is converted into an hexagonal pattern in each of the two independent vectorial components of the electric field. A transition from bright to dark hexagons (or viceversa) in each of the field components is obtained by changing the ellipticity of the driving field. In addition, the range of parameters (cavity detuning and input intensity) for which a pattern exists shrinks to zero as the ellipticity departs from its value for linearly polarized driving field. Beyond a certain ellipticity, the homogeneous solution, which now is elliptically polarized, never looses stability. The change from stripes or squares to hexagons for any finite ellipticity follows from general symmetry considerations also made in \cite{geddes2} for a Kerr medium with counterpropagating beams. It has been predicted \cite{Scroggie96} and observed \cite{Aumann97} in a Na cell with single feedback mirror. It has also been invoked as an explanation of the observations in \cite{Maitre95} for a cell of rubidium vapor in two counterpropagating beams. The transition from bright to dark hexagons by changing the ellipticity of a driving field has been discussed in \cite{Ackeman95} and observed in \cite{Aumann97} for Na vapor with a single feedback mirror. For an elliptically polarized driving field and a self-focusing situation, the field ellipticity is a tuning parameter that permits to explore several situations. For circularly polarized driving field the scalar case is recovered and a circularly polarized hexagonal pattern emerges. Two particularly interesting cases for intermediate ellipticity involve dynamical patterns occurring through a Hopf instability at a finite wave number $k_h$. The first of these instabilities considered in Sec. \ref{Blinking Hexagons} leads to a time dependent pattern consisting of deformed dynamical hexagons. The Hopf bifurcation is modified by a linearly damped stationary mode $k_s>k_h$. The quadratic resonance of these two modes gives rise to deformed hexagons. The spatiotemporal pattern can be described by a superposition of a triad of $k_h$ Hopf modes, each one associated with a traveling wave with the same frequency, but one of them has an amplitude different from the other two. A second interesting case considered in Sec. \ref{cod2} is a codimension two bifurcation in which a stationary Turinglike instability occurs simultaneously with the Hopf instability. Now the stationary unstable mode is such that $k_h>k_s$, and we do not find quadratic resonance. The transient dynamics involves complicated states with strong competition of the Hopf and stationary modes. The final state is a stationary hexagonal pattern in which the Hopf modes have died. Dynamical patterns and codimension Turing-Hopf bifurcations have been considered in related studies. A first case of dynamical hexagons, in a Kerr medium with counterpropagating beams, is discussed in \cite{geddes2}, but they arise as a secondary pure Hopf instability. Dynamical hexagons arising from a codimension two situation, for alkali vapors with a single feedback mirror, are considered in \cite{LogvinEPL,Logvin96}. The difference with the situations we have found is twofold. First, the codimension two situation is different from ours because $k_s>k_h$, so that a mixed Turing-Hopf mode occurs through a quadratic resonances. Second, the dynamical pattern of \cite{LogvinEPL,Logvin96} is different from our deformed dynamical hexagons because the Turinglike mode is linearly unstable and therefore appears with a large amplitude. In spite of these differences, we note that the deformation of hexagons in \cite{Logvin96} has an origin similar to ours. The competition of unstable Hopf modes and unstable Turinglike modes with $k_s>k_h$ has also been considered experimentally in \cite{Residori96} for a large system. By appropriate control of parameters, either of the two types of modes can dominate. Close to the codimension two situation, spatial coexistence of domains in which one of the two modes dominates is observed \cite{last_news}. A codimension two bifurcation involving a Turinglike mode and a Hopf mode, but of zero wave number ($k_h=0$), has also been recently considered in an optical parametric oscillator with saturable losses \cite{tlidi}. \section{Conclusions} \label{conclusions} In this paper we have presented a systematic analysis of a prototype vectorial model of pattern formation in nonlinear optics describing a Kerr medium in a cavity with flat mirrors and driven by a coherent plane-wave field. We have considered linearly as well as elliptically polarized driving fields and situations of self-focusing and self-defocusing. We have described, by numerical simulations, amplitude and model equations, a rich variety of phenomena that illustrate the relevance of the polarization degree of freedom in optical spatiotemporal dynamics. In particular, we have shown that this degree of freedom allows for new asymmetric homogeneous solutions, induces new instabilities, and changes the basic symmetry of the pattern formed beyond an instability. A particular relevant aspect of our results is to show that the ellipticity of the pump can be used as a tuning parameter, easily accessible to the experimentalist, that permits to explore different types of pattern forming instabilities. For example, we have shown that by changing the ellipticity we find situations which include: i) a modified Hopf bifurcation at finite wave number leading to a time dependent pattern of deformed hexagons, ii) a codimension two Turing-Hopf instability resulting in an elliptically polarized stationary hexagonal pattern. Small changes in the ellipticity also change the symmetry of the pattern, for example from stripes to hexagons. Another general relevant aspect of our results is that the information on two different patterns is encoded in each of the two independent polarization components which are easily separated by a polarizer. From a fundamental point of view this opens the possibility to study spatiotemporal correlations between two different patterns emerging from the same physical system \cite{sti}. Correlations between different polarization components close to the onset of pattern formation in the model of Ref. \cite{geddes1} have already been studied in \cite{corrkerr}. We finally mention that vectorial degrees of freedom are also important in pattern formation in lasers where vectorial topological defects of the field amplitude can occur if the rotational symetry of the system is not broken by a pump field \cite{gil93,prati,pismenbook,2dVCGLE}. Corresponding phenomena should also exist in other fields such as two-component Bose-Einstein condensates \cite{busch}. This calls for further systematic investigation of vectorial spatiotemporal phenomena for which a number of experimental results are becoming available. \section{Acknowledgements} We want to acknowledge interesting discussions with Dr. M. Santagiustina. We also acknowledge Dr. A. Vicens for a careful reading of the manuscript. This work is supported by QSTRUCT (Project ERB FMRX-CT96-0077). Financial support from DGICYT (Spain) Project PB94-1167 is also acknowledged. M. H. wants to acknowledge financial support from the FOMEC project 290, Dep. de F\'{\i}sica FCEyN, Universidad Nacional de Mar del Plata, Argentina. \begin{thebibliography}{123} \bibitem[+]{Daniel} Permanent address: Center for Nonlinear Phenomena and Complex Systems, Universit\'e Libre de Bruxelles, Campus Plaine, Blv. du Triomphe B.P 231, 1050 Bruxelles. \bibitem[*]{www} Electronic address: http://www.imedea.uib.es/PhysDept/ \bibitem{Chaossolfract} {\it Nonlinear Optical Structures, Patterns, Chaos}, edited by L. A. Lugiato, special issue of Chaos, Solitons and Fractals, {\bf 4}, 1251 (1994) and references therein. \bibitem{Alessandro} D. Alessandro and W. J. Firth, Phys. Rev. Lett. {\bf 66}, 2597 (1991); Phys. Rev. A {\bf 46}, 537 (1992). \bibitem{CrossHoh} M. C. Cross and P. C. Hohenberg, Rev. Mod. Phys. {\bf 65}, 851 (1993). \bibitem{busch} Th. Busch, J. I. Cirac, V. M. Perez-Garcia and P. Zoller, preprint (cond-mat/9712123). \bibitem{kitano} M. Kitano, T. Yabuzaki and T. Ogawa, Phys. Rev. Lett. {\bf 46}, 926 (1981). \bibitem{zheludev} M. I. Zheludev, Sov. Phys. Usp. {\bf 32}(4), 357 (1989). \bibitem{puccioni} G. P. Puccioni, M. V. Tratnik, J. E. Sipe and G. L. Oppo, Opt. Lett. {\bf 12}, 242 (1987). \bibitem{abraham} N. B. Abraham, M. San Miguel and E. Arimondo, Opt. Comm. {\bf 117}, 344 (1995); {\bf 121}, 168 (1995). \bibitem{matlin} N. B. Abraham, M. D. Matlin, R. S. Gioggia, Phys. Rev. A {\bf 53}, 3514 (1996). \bibitem{serrat} C. Serrat, A. Kulminskii, R. Vilaseca and R. Corbalan, Opt. Lett. {\bf 20}, 1353 (1995). \bibitem{maxiPRL} M. San Miguel, Phys. Rev. Lett. {\bf 75}, 425 (1995). \bibitem{gil93} L. Gil, Phys. Rev. Lett. {\bf 70}, 162 (1993). \bibitem{toniPRL} A. Amengual, D. Walgraef, M. San Miguel and E. Hern\'andez-Garc\'{\i}a, Phys. Rev. Lett. {\bf 76}, 1956 (1996). \bibitem{sti} A. Amengual, E. Hern\'andez-Garc\'{\i}a, R. Montagne and M. San Miguel, Phys. Rev. Lett. {\bf 78}, 4379 (1997). \bibitem{CO2Meucci} C. Taggiasco, R. Meucci and N. B. Abraham, Opt. Comm. {\bf 133}, 507 (1997). \bibitem{VCSELs} J. Martin Regalado, S. Balle and M. San Miguel, Opt. Lett. {\bf 22}, 460 (1997). \bibitem{prati} F. Prati, G. Tissoni, M. San Miguel and N. B. Abraham, Opt. Comm. {\bf 143}, 133 (1997). \bibitem{geddes1} J. B. Geddes, J. V. Moloney, E.M. Wright and W.J. Firth, Opt. Comm. {\bf 111}, 623-631 (1994). \bibitem{Scroggie96} A. J. Scroggie and W. J. Firth, Phys. Rev. A {\bf 53}, 2752 (1996). \bibitem{Tam77} A. C. Tam and W. Happer, Phys. Rev. Lett. {\bf 38}, 278 (1977). \bibitem{Gahl} A. Gahl, J. Seipenbusch, A. Aumann, M. M\"oller and W. Lange, Phys. Rev. A {\bf 50}, R917 (1994). \bibitem{Grynberg94} G. Grynberg, A. Maitre, and A. Petrossian, Phys. Rev. Lett. {\bf 72}, 2379 (1994). \bibitem{Aumann97} A. Aumann, E. B\"uthe, Yu. A. Logvin, T. Ackemann and W. Lange, Phys. Rev. A {\bf 56}, R1709 (1997). \bibitem{Maitre95} A. Maitre, A. Petrossian, A. Blouin, M. Pinard, G. Grynberg, Opt. Comm. {\bf 116}, 153 (1995). \bibitem{MoloneyKerr} J.V. Moloney, H. M. Gibbs, Phys. Rev. Lett. {\bf 48}, 1607 (1982); A. Quarzeddini, H. Adachihara and J. V. Moloney, Phys. Rev. A {\bf 38}, 2005 (1988) \bibitem{lugi} L. A. Lugiato and R. Lefever, Phys. Rev. Lett. {\bf 58}, 2209 (1987). \bibitem{firt} W. J. Firth, A. J. Scroggie, G. S. McDonald and L. Lugiato, Phys. Rev. A {\bf 46}, R3609 (1992). \bibitem{corrkerr} M. Hoyuelos, P. Colet and M. San Miguel, Phys. Rev. E, to appear (1997). \bibitem{footnote1} In \cite{geddes1} $I_0$ is defined as the intensity of any of the components of the input field, so in the equivalent equation in \cite{geddes1} $I_0$ is not divided by 2. \bibitem{num_method} The numerical scheme used for integration of Eq. (\ref{1}) is described in detail in R. Montagne, E. Hern\'andez-Garc\'{\i}a, A. Amengual and M. San Miguel, Phys. Rev. E {\bf 56}, 151 (1997). The method is pseudospectral and second-order accurate in time, and is similar to the so-called two-step method. Lattices of size $128\times 128$ were used. \bibitem{firthsolfract} A. J. Scroggie, W. J. Firth, G. S. McDonald, M. Tlidi, R. Lefever and L. A. Lugiato, Chaos Solitons \& Fractals {\bf 4}, 1323 (1994). \bibitem{NLS} See, for example, K. Rypdal and J.J. Rasmussen, Phys. Scr. {\bf 33}, 498 (1986). \bibitem{goldman80} M. V. Goldman, K. Rydal, B. Hafizi, Phys. Fluids {\bf 23}, 945 (1980). \bibitem{geddes2} J. B. Geddes, R. A. Indik, J. V. Moloney and W. J. Firth, Phys. Rev. A {\bf 50}, 3471 (1994). \bibitem{shoka} P. Borckmans, G. Dewel, A.De Wit and D. Walgraef, in {\it Chemical Waves and Patterns}, edited by R. Kapral and K. Showalter (Kluwer, Dordrecht, 1994), pp. 323-363. \bibitem{Ackeman95} T. Ackemann, Y. A. Logvin, A. Heuer and W. Lange, Phys. Rev. Lett. {\bf 75}, 3450 (1995). \bibitem{ouyang} Q. Ouyang, G. H. Gunaratne, H.L. Swinney, Chaos {\bf 3}, 707 (1993). \bibitem{daniel} D.Walgraef, {\it Spatio-Temporal Pattern Formation} (Springer-Verlag, New York, 1996). \bibitem{manneville} P. Manneville, {\it Dissipative Structures and Weak Turbulence} (Academic Press, San Diego, 1990). \bibitem{brand} H. R. Brand and R. J. Deissler, {\it Pattern Formation near an Oscillatory Instability for Systems without up-down Symmetry} (unpublished). \bibitem{haken} M. Bestehorn and H. Haken, Phys. Rev A {\bf 42}, 7195 (1990). \bibitem{lomdahl} H.R. Brand, P.S. Lomdahl and A.C. Newell, Physica {\bf D 23}, 345 (1986). \bibitem{proctorjones} C.A. Jones and M.R.E. Proctor, Phys. Lett. {\bf A 121}, 224 (1987); M.R.E. Proctor and C.A. Jones, J. Fluid Mech. {\bf 188}, 301 (1988). \bibitem{kuramoto} Y. Kuramoto, {\it Chemical Oscillations, Waves and Turbulence} (Springer-Verlag, Berlin, 1984). \bibitem{Lima96} D. Lima, A. De Wit, G. Dewel and P. Borckmans, Phys. Rev. E {\bf 53}, R1305 (1996). \bibitem{DeWit96} A. De Wit, D. Lima, G. Dewel and P. Borckmans, Phys. Rev. E {\bf 54}, 261 (1996). \bibitem{gunton} J. d. Gunton, M. San Miguel and P. Sahni, in {\it Phase transitions and critical phenomena} vol. 8. Eds. C. Domb and J. Lebowitz. Academic Press (New York, 1983). \bibitem{Berre97}M. Le Berre, A. S. Patrascu, E. Ressayre, and A. Tallet, Phys. Rev. A {\bf 56} 3150 (1997). \bibitem{LogvinEPL} Y. A. Logvin, T. Ackemann and W. Lange, Europhys. Lett. (1997) \bibitem{Logvin96} Y. A. Logvin, B. A. Samson, A. A. Afanas'ev, A. M. Samson and N. A. Loiko, Phys. Rev. E {\bf 54}, R4548 (1996). \bibitem{Residori96} S. Residori, P. L. Ramazza, E. Pampaloni, S. Boccaletti and F. T. Arecchi, Phys. Rev. Lett. {\bf 76}, 1063 (1996). \bibitem{last_news} After the completion of this work we have learnt of a theoretical study \cite{pismen} of the experiments in \cite{Residori96} and other related situations. The authors of \cite{pismen} consider a quadratic resonance of a Turinglike mode and two Hopf modes of finite wave number in a similar way as our analysis of the deformed dynamical hexagons. \bibitem{pismen} L. M. Pismen and B. Y. Rubinstein, {\em Quasicrystalline and Dynamic Planforms in Nonlinear Optics} (unpublished). \bibitem{tlidi} M. Tlidi, P. Mandel and M. Haelterman, Phys. Rev. E {\bf 56}, 6524 (1997). \bibitem{pismenbook} L. Pismen, {\it Vortices in Nonlinear Fields} (Oxford University Press, 1998) \bibitem{2dVCGLE} M. Hoyuelos, E. Hern\'andez-Garc\'{\i}a, P. Colet and M. San Miguel (unpublished) \end{thebibliography} \begin{figure} %\psfig{figure=figure1.ps,width=14cm} \caption{Steady state homogeneous solutions, as a function of the input field intensity, for linearly polarized light, $\chi=90^\circ$. The solid line is the symmetric solution. The dashed line corresponds to the asymmetric solutions. The two branches of the asymmetric solution give the values of $I_{s+}$ and $I_{s-}$ in one of the two degenerate solutions, and they meet the symmetric solution for $I_{s+}=I_{s-}=I'$. Parameter values: $a=1$, $B=1.5$ and $\theta=1$. These parameter values are the same for all the figures except where otherwise noticed. The quantities plotted in all the figures are dimensionless.} \label{fig1} \end{figure} \begin{figure} %\psfig{figure=figure2.ps,width=14cm} \caption{Steady state solutions for different values of the ellipticity $\chi$ of the input field. The solid lines correspond to the value of $I_{s+}$ and the dashed lines to $I_{s-}$. Input field ellipticity values: (a) $\chi=87^\circ$, (b) $\chi=78^\circ$, (c) $\chi=73^\circ$, (d) $\chi=0^\circ$. Note that $I_{s-}=0$ in Fig. (d).} \label{fig3a} \end{figure} \begin{figure} %\psfig{figure=figure3.ps,width=14cm} %\vspace*{-3cm} \caption{Marginal stability curves for linearly polarized input field as described in the text. (a) Stability of the symmetric solution and (b) stability of symmetric solution (up to $I'$) and asymmetric solution. The dotted line displays the value of $I'$. The vertical dashed line separates the self-focusing case (to the right) from the self-defocusing one (left).} \label{fig2} \end{figure} \begin{figure} %\vspace*{5cm} %\psfig{figure=figure4.ps} %\vspace*{-5cm} \caption{Plots of $|E_+|^2$ for increasing values of the intensity of the linearly polarized input field in the self-defocusing case ($\eta=-1$). From left to right and from top to bottom: $I_{s+}=0.8$ ($I_0=1.8$), $I_{s+}=1.5$ ($I_0=3.1$), $I_{s+}=2.6$ ($I_0=5.1$) and $I_{s+}=3.0$ ($I_0=6.0$). Gray scale values: black$=0.1$, white$=3.8$.} \label{fig4} \end{figure} \begin{figure} %\vspace*{5cm} %\psfig{figure=figure5.ps} %\vspace*{-2.5cm} \caption{Plots of $|E_+|^2$ for increasing values of the input field in the self-focusing regime ($\eta=1$). In this figure the gray scale is logarithmic. From top to bottom $I_{s+} = 0.48$ ($I_0=0.96$, black$=0.16$, white $=3.5$), $I_{s+} = 0.55$ ($I_0=1.1$, black$=0.0018$, white$=9.8$) and $I_{s+} = 1.7$ ($I_0=3.2$, black$=0.0028$, white$=17$).} \label{fig5} \end{figure} \begin{figure} %\psfig{figure=figure6.ps,width=14cm} \caption{Marginal stability curves for the same values of the ellipticity $\chi$ of the input field as in Fig. \protect\ref{fig3a}: (a) $\chi=87^\circ$, (b) $\chi=78^\circ$, (c) $\chi=73^\circ$, (d) $\chi=0^\circ$.} \label{fig3b} \end{figure} \begin{figure} %\vspace*{5cm} %\psfig{figure=figure7.ps} %\vspace*{-5cm} \caption{ $|E_+|^2$ for increasing values of the quasi linearly polarized input field ($\chi=87^\circ$) in the self-defocusing case ($\eta=-1$). From left to right and from top to bottom: $I_{s+}=0.8$ ($I_0=1.8$), $I_{s+}=1.7$ ($I_0=3.1$), $I_{s+}=2.1$ ($I_0=3.8$) and $I_{s+}=2.3$ ($I_0=4.2$). Gray scale values: black$=0.13$, white$=2.6$.} \label{fig6} \end{figure} \begin{figure} %\vspace*{2cm} %\psfig{figure=figure8.ps} %\vspace*{-10cm} \caption{ $|E_x|^2$ (left figure, black$=1.2$, white$=1.5$) and $|E_y|^2$ (right figure, black$=0$, white$=0.58$) for $I_{s+}=0.8$ ($I_0=1.8$). The values of the other parameters are the same as in Fig. \protect \ref{fig6}.} \label{fig6b} \end{figure} \begin{figure} %\vspace*{10cm} %\psfig{figure=figure9.ps} %\vspace*{-5cm} \caption{From left to right and from top to bottom: $|E_+(x,y)|^2$ (near field, black$=1.1$, white$=6.5$), $|E_+(\vec k)|^2$ (far field, white$=0$, black$=0.02$), $|E_-(x,y)|^2$ (near field, black$=0$, white$=1.2$) and $|E_-(\vec k)|^2$ (far field, white$=0$, black$=0.019$). The far fields are drawn in logarithmic scales. The homogeneous mode ($k=0$) is in the center of the far field plots and has been eliminated in all figures. Parameters: $I_0=3.92$, $\chi=78^\circ$ and $\eta=1$.} \label{fig7d} \end{figure} \begin{figure} %\vspace*{10cm} %\psfig{figure=figure10.ps} %\vspace*{-5cm} \caption{Four configurations of $|E_+|^2$ during one period $T$ of oscilation ($T=3.86$). From left to right and from top to bottom: $t=0$, $t=0.97$, $t=1.93$ and $t=2.90$. The values of the parameters are the same as in Fig. \protect\ref{fig7d}. Gray scale values: black$=1.1$, white$=6.5$.} \label{fig7a} \end{figure} \begin{figure} %\hspace*{1cm}\psfig{figure=figure11.eps,width=14cm} %\vspace*{1cm} \caption{Definition of the unstable Hopf and slaved stationary modes coupled through quadratic non linearities and described by the dynamical system (\ref{dyn}). For asymptotic times, the modes $k_1$, $k_2$ and $k_3$ dominate and build the dynamical hexagons described in the text. The intensity range of these modes obtained from simulations (see $|E_+(\vec k)|^2$ in Fig. \protect \ref{fig7d}) is from 0.0008 (damped static modes) to 0.019 (brightest Hopf mode).} \label{fig7e:hexagons_modes} \end{figure} \begin{figure} %\hspace*{-1.5cm}\psfig{figure=figure12.ps} \caption{Phase of the Hopf mode $k_2$ (see Fig. \protect\ref{fig7e:hexagons_modes}) as a function of time. The values of the parameters are the same as in Fig. \protect \ref{fig7d}. A period $T=3.86$ is obtained from the plot.} \label{fig7b} \end{figure} \begin{figure} %\psfig{figure=figure13.ps} \caption{Real part of most unstable eigenvalues $\lambda_1$ and $\lambda_2$. The values of the parameters are the same as in Fig. \protect \ref{fig7d}.} \label{fig7c} \end{figure} \begin{figure} %\psfig{figure=figure14.ps} \caption{Real part of unstable eigenvalues $\lambda_1$ and $\lambda_2$. Parameters: $I_0=10.9$, $\chi=67^\circ$ and $\eta=1$.} \label{fig8b} \end{figure} \begin{figure} %\vspace*{-2cm} %\psfig{figure=figure15.ps} %\vspace*{1cm} \caption{Near field ($|E_+(x,y)|^2$) and far field ($|E_+(\vec k)|^2$, in logarithmic scale) at three different times for the Turing-Hopf competition. From top to bottom: $t=50$ (near field: black$=3.6$, white$=7.1$; far field: white$=0$, black$=5.3\times 10^{-4}$), $t=700$ (near field: black$=4.6$, white$=5.5$; far field: white$=0$, black$=2.1\times 10^{-4}$) and $t=1700$ (near field: black$=0.6$, white$=16$; far field: white$=0$, black$=0.039$). The values of the parameters are the same as in Fig. \protect \ref{fig8b}.} \label{fig8a} \end{figure} \begin{figure} %\hspace*{-1.5cm}\psfig{figure=figure16.ps} \caption{Time dependence of the amplitudes $|A_s|^2$ (solid line) and $|A_h|^2$ (dashed line), corresponding to wave numbers $k_s$ and $k_h$, in logarithmic scale. The amplitudes were calculated integrating over a circle of radius $k_s$ or $k_h$ in Fourier space. The plot shows the exponential growth of the Hopf modes at short times, the competition between the modes and the final domination of the static modes.} \label{fig_compet} \end{figure} \end{document}