1: \documentclass[12pt]{article}
2: \usepackage{graphicx}
3: \usepackage{amssymb}
4:
5: \def\bq{\begin{equation}}
6: \def\eq{\end{equation}}
7: \def\ba{\begin{array}}
8: \def\ea{\end{array}}
9: \def\e{\varepsilon}
10: \def\vir{\mbox{\hspace{3mm},\hspace{6mm}}}
11: \def\sgn{\mbox{\rm sgn\,}}
12: \begin{document}
13: \begin{titlepage}
14: \begin{center}
15: {\huge
16: Experimental and theoretical study of the passively
17: mode-locked Ytterbium-doped double-clad fiber laser}
18: \end{center}
19: \vspace{1.5cm}
20:
21: \noindent
22: Herv\'e Leblond$\rm ^{(a)}$, Mohamed Salhi$\rm ^{(a)}$, Ammar Hideur$\rm ^{(b)}$,
23: Thierry Chartier$\rm ^{(b)}$, Marc Brunel$\rm ^{(b)}$ and Fran\c{c}ois Sanchez$\rm
24: ^{(a)}$\\\vspace{1cm}
25:
26: {\it \noindent
27: (a) Laboratoire POMA, UMR 6136, Universit\'e d'Angers, 2 Bd
28: Lavoisier, 49000 Angers, France\\
29: (b) Groupe d'Optique et
30: d'Optronique, CORIA, UMR 6614, Universit\'e de Rouen, Avenue de l'Universit\'e,
31: site Universitaire du
32: Madrillet, B.P. 12, 76801 Saint-\'Etienne du Rouvray, France}
33:
34: \vspace{2.5cm}
35:
36: {\bf Abstract}\vspace{5mm}
37:
38: We consider an Yb-doped double-clad fiber laser in a unidirectional ring cavity containing a polarizer placed between
39: two half-wave plates. Depending on the orientation of the phase plates, the laser operates in continuous, Q-switch,
40: mode-lock or unstable self-pulsing regime. An experimental study of the stability of the mode locking regime is
41: realized versus the orientation of the half-wave plates. A model for the stability of self-mode-locking and
42: cw operation is
43: developed starting from two coupled nonlinear Schr\"odinger equations in a gain medium. The model is reduced to a
44: master equation in which the coefficients are explicitly dependent on the orientation angles of the phase plates.
45: Analytical solutions are given together with their stability versus the angles.
46:
47:
48: \end{titlepage}
49: \section{Introduction}
50:
51: Self-mode-locking has been successfully obtained in both low power and double-clad fiber lasers using Additive-Pulse
52: Mode locking technique through nonlinear polarization rotation \cite{1}-\cite{6}. The experimental configuration generally consists
53: in a ring cavity containing a polarizing isolator placed between two polarization controllers. The basic principle is the
54: following. The polarization state evolves nonlinearly during the propagation of the pulse in the fiber due to
55: the combined effects of self-phase modulation and cross-phase modulation induced on the two orthogonal polarization
56: components, both resulting from the optical Kerr effect. A polarization controller is adjusted at the output of the
57: fiber such that the polarizing isolator lets pass the central intense part of the pulse but blocks the low-intensity
58: pulse wings. This technique of nonlinear polarization rotation for passive mode locking leads to stable, self-starting
59: pulse trains. Theoretical approaches to this problem can be divided into two categories. The first one has been proposed by
60: Haus {\it et al} \cite{7} and consists in writing a phenomenological scalar equation (the master mode-locking equation) assuming that all
61: effects per pass are small. The model includes the group velocity dispersion (GVD), the optical Kerr nonlinearity and a gain medium.
62: No birefringence is taken into account. Although the mode-locking properties of the laser can be described through this approach
63: for
64: positive and negative GVD, the stability of the mode-locked solutions as a function of
65: the orientation of the polarization controllers cannot be described in this way. On the other hand Kim {\it et al} \cite{8} use an approach based
66: on two coupled nonlinear Schr\"odinger equations. The medium is assumed to be birefringent.
67: The GVD
68: together with the Kerr effect are taken into account. The orientation angles of the eigenaxis of the fiber at both sides
69: of the polarizer are explicitly taken into account through a periodic perturbation. Numerical simulations show that
70: mode-locking can be achieved for negative GVD. To our best knowledge, none of the models take
71: into account all the characteristics of the fiber (birefringence, GVD, gain, optical Kerr effect)
72: and the action of the orientation of the polarization controllers.
73:
74: The aim of this paper is to investigate experimentally and theoretically the mode-locking properties of the Yb-doped double-clad fiber
75: laser. Mode-locking is achieved through nonlinear polarization rotation in a unidirectional ring cavity containing a polarizer
76: placed between two half-wave plates. This paper is organized as follows. Section II is devoted to the experimental results. We first
77: briefly recall the operating regime of the laser as a function of the orientation of the two half-wave plates. A stability diagram
78: of the mode-locked laser is presented \cite{9}. The parameters required in the modelling section are measured. The birefringence
79: of the double-clad fiber is first determined using the magneto-optic method \cite{10}. The total dispersion has been
80: estimated from other experiments where compression of the pulses was realized
81: \cite{11,11bis}. In section \ref{III} the
82: stability of the mode-locking regime is theoretically investigated. The analysis is based on the coupling
83: between the two linearly polarized eigenstates
84: of the fiber \cite{8}\cite{12}. The propagation problem is modelled through two nonlinear Schrodinger equations in a gain medium where the
85: GVD, the nonlinear coefficient and the gain are assumed small over the cavity length. The eigenaxis of
86: the fiber at each side of the polarizer are referenced with respect to the polarizing axis. These angles can be adjusted with
87: two half-wave plates and will thus be considered as variable in the analysis. The polarizer is treated as a periodic
88: perturbation acting as a projecting operator. Using the expressions derived in the propagation problem, we calculate the
89: electric field amplitude just after the polarizer as a function of the corresponding amplitude at the previous round-trip.
90: The round-trip number is then considered as a continuous variable, which allows us to use a multiscale analysis approach.
91: The final equation is a complex Ginzburg-Landau equation whose coefficients explicitly involve the orientation angles
92: of the half-wave plates. Explicit solutions are given and their stabilities are discussed. A stability diagram of
93: the mode-locked regime is calculated and compared to the experimental diagram.
94:
95: \section{Experimental results}
96:
97: In this section we briefly recall the results obtained with an Yb-doped double-clad fiber laser in a unidirectional
98: ring cavity operating at 1.08 $\rm \mu m$ \cite{9}. A polarizing isolator is placed between two half-wave plates. Although this
99: configuration is known to generate ultrashort pulses in low power fiber lasers, it can also lead to very different
100: regimes depending on the orientation of the polarization elements, such as cw operation, Q-switching, mode-locking,
101: or unstable emission regime. We thus perform a systematic study of the operating regime as a function of the relative
102: orientation of the two half-wave plates. The experimental setup is shown in figure \ref{figs1}.
103: \begin{figure}[hbt!]
104: \begin{center}
105: \includegraphics[width=9cm]{figs1.EPS}
106: \caption{\footnotesize Experimental setup.}
107: \label{figs1}
108: \end{center}
109: \end{figure}
110: The 4 meter long Yb-doped double-clad
111: fiber is side-pumped using the v-groove technique \cite{3}. The launching efficiency is about 70 percent. This pumping
112: scheme is very convenient for ring laser applications because the two fiber ends remain free. The fiber core diameter
113: is 7 $\rm \mu m$ allowing a single-mode laser signal propagation. The inner cladding is a $125\times 125\;\;\rm \mu m^2$
114: square leading to a
115: multimode pump propagation. The geometry of the inner cladding ensures an efficient pump absorption. Indeed,
116: the pump power is almost completely absorbed along the doped fiber. In addition, multimode propagation is required
117: for high injection efficiency of large area laser diodes. Two single-mode fibers at 1 $\rm \mu m$ are spliced at both ends of
118: the double-clad fiber leading to a total fiber length of about 8.5 m, the cavity length is about 9 m. Fiber ends
119: are angle-cleaved in order to avoid any Fresnel reflection which could generate a standing-wave operating regime.
120: The splicing losses are less than 0.1 dB. The doped fiber is pumped at 975 nm with a 4 W semiconductor laser.
121: A bulk polarizing optical isolator is used to obtain a travelling wave laser. The latter is placed between
122: two half-wave plates at 1.08 $\rm \mu m$. Each plate can be rotated in a plane perpendicular to the propagation direction.
123: A partially reflecting mirror ensures an output coupling of about 90 percent. The output beam
124: is analyzed through different means. A fast photodiode is used when the laser is cw, Q-switched or unstable.
125: In the regular mode-locking regime, we use an optical autocorrelator to measure the duration of the pulses.
126: Note that there is no compression system in the cavity. An optical spectrum analyzer (Advantest) together with
127: a microwave spectrum analyzer (Tektronix) are also used when needed. Finally, a powermeter allows us to measure
128: the average output power.
129:
130: We investigate now the operating regime of the laser as a function
131: of the orientation of the two waveplates. $\theta_1$ ($\theta_2$)
132: is the orientation of one eigenaxis of the first (second) half-wave
133: plate referenced to the input (output) polarizer of the optical
134: isolator. The angle between the input and output polarizer of the
135: isolator is $\rm 45$ degrees. For a given pumping power we record the
136: evolution of the output
137: intensity versus $\theta_1$ and $\theta_2$. Practically, we fix $\theta_1$ and measure the particular values of $\theta_2$
138: for which a change in the operating regime occurs. It is convenient to represent these results in the plane $(\theta_1,\theta_2)$.
139: It allows to easily identify the output regime of the laser. The experimental results, obtained for a pumping power
140: of about 2.75 W, are summarized in figure \ref{figs2}.
141: \begin{figure}[hbt!]
142: \begin{center}
143: \includegraphics[width=9cm]{figs2t.EPS}
144: \caption{\footnotesize Operating regime of the laser as a function of the orientation of the two half-wave plates $\theta_1$ and $\theta_2$.
145: The pump power was 2.75 W. ML stands for mode-locked regime and cw for continuous wave regime.}
146: \label{figs2}
147: \end{center}
148: \end{figure}
149: We have experimentally observed that the different output regimes are periodic
150: versus $\theta_1$ and $\theta_2$ with a period of 90 degrees.
151: Figure \ref{figs2} shows that the laser essentially operates in cw, mode-lock or unstable regime while the Q-switch behavior appears near
152: the boundary between cw and mode-lock regimes. Note that since the operating regime depends on the pumping rate,
153: the mapping of figure \ref{figs2} is pump power dependent. Typically, when the pump power decreases (increases), the ranges of
154: mode-locking decrease (increase).
155:
156: The different dynamical behaviors observed in our experiment have been discussed in details in reference \cite{9}.
157: Here, we concentrate our attention on
158: the mode-locked operating regime. In figure \ref{figs3} we give the results obtained for $\theta_1 = \rm 20\, deg$
159: and $\theta_2 =\rm 0\, deg$.
160: In this case, the laser is first cw just above threshold (figure \ref{figs3}a). While the pump power is increased, an irregular
161: self-pulsing regime appears which finally becomes a regular self-mode-locking regime. As previously discussed in the introduction,
162: the basic principle responsible for the mode-locking is well known in such configurations. It is based on Kerr effect through
163: nonlinear polarization rotation. This technique is also called Polarization-Additive Pulse Mode-locking. The repetition rate
164: of the pulses corresponds to the free spectral range of the cavity, about 20 MHz in our case. Auto-correlation traces showed
165: a coherence spike of about 150 fs on a longer pedestal of about 60 ps \cite{6}. A typical spectrum in the mode-locking range is given
166: in figure \ref{figs3}b.
167: \begin{figure}[hbt!]
168: \begin{center}
169: \includegraphics[width=9cm]{figs3b.EPS}
170: \caption{\footnotesize $\theta_1 = \rm 20\, deg$ and $\theta_2 = \rm 0\, deg$. (a) laser characteristic and (b) typical optical spectrum in
171: the mode-locked regime.}
172: \label{figs3}
173: \end{center}
174: \end{figure}
175: The mode-locked spectrum bandwidth is more than 25 nm. This spectrum suggests that shorter pulses can be achieved
176: by a suitable dispersion compensation. We have recently shown that femtosecond pulses are obtained with a grating pair inserted
177: in the cavity \cite{6}. In the present mode-locked regime, we measure 600 mW output power for 3.25 W pump power, which corresponds to high
178: energy pulses of about 30 nJ.
179:
180: Consider now the measurements of parameters required in the theoretical section. We first measure the beating length $L_B$
181: of the double-clad fiber using the magneto-optic method \cite{10}. The method consists to apply a sinusoidal magnetic field produced
182: by a coil in a small portion of the fiber and to move the coil along the fiber axis. The fiber is placed between an analyzer and
183: a polarizer. This system is then illuminated and analyzed in transmission. The detected output power is a periodic signal versus
184: the coil displacement. The period in the spatial domain coincides with the beating length of the fiber. We found $L_B = 50\rm\; cm$
185: leading to a birefringence parameter $K = 1/( L_B) = 2\;\rm m^{-1}$. We must also consider the birefringence of the undoped fibers.
186: It has been measured to be of about $1\;\rm m^{-1}$. In the following we will consider an average value
187: of the birefringence of about $1.5\;\rm m^{-1}$. A second important parameter is the GVD.
188: It has been estimated
189: from compression experiments \cite{11bis}. We find $\beta_2 = 0.026\;\rm ps^2/m$.
190:
191: \section{Theory}\label{III}
192: The aim of this section is to develop a theoretical model able to describe the mode-locking properties of the Yb-doped fiber laser
193: as a function of the orientation of the two half-wave plates. For that, we consider a birefringent fiber in a unidirectional ring cavity
194: with an intracavity polarizer. The orientation of the eigenaxis of the fiber at each sides of the polarizer are unspecified and will
195: be taken as variable parameters. This is equivalent to considering that the eigenaxis of the fiber are fixed and two half-wave
196: plates are used to adjust their orientations with respect to the polarizer.
197:
198: \subsection{Derivation of a master equation}
199: \subsubsection{Propagation in the fiber}
200: The starting point are the equations giving the evolution of the two polarization components in a gain medium with Kerr nonlinearity
201: and GVD. In the framework of the eigenaxis of the birefringent fiber moving at the
202: group velocity,
203: the pulse envelope evolution is described by the following system \cite{8}\cite{12}:
204: \bq i\partial_zu-K u-\frac{\beta_2}2\partial_t^2u
205: +\gamma\left(u\left|u\right|^2+Au\left|v\right|^2+Bv^2u^\ast\right)=
206: ig\left(1+\frac1{\omega_g^2}\partial_t^2\right)u,\label{1}\eq
207: \bq i\partial_zv+K v-\frac{\beta_2}2\partial_t^2v
208: +\gamma\left(v\left|v\right|^2+Av\left|u\right|^2+Bu^2v^\ast\right)=
209: ig\left(1+\frac1{\omega_g^2}\partial_t^2\right)v,\label{2}\eq
210: where $\partial_t$ denotes the partial derivative operator $\frac{\partial}{\partial t}$.
211: $\beta_2$ (in $\rm ps^2/m$) is the GVD coefficient. $K$ (in $\rm m^{-1}$)
212: is the birefringent parameter and is related to the $x$ and $y$ refractive indexes through the
213: relation $K = \pi(n_x-n_y)/\lambda$, where lambda is the optical wavelength.
214: $\gamma = 2\pi n_2/(\lambda A_{eff})$ is the nonlinear coefficient, $n_2$ (in $\rm m^2/W)$
215: is the nonlinear index coefficient and $A_{eff}$ (in $\rm m^2$) the effective core area of the fiber.
216: $A$ and $B$ are the dielectric coefficients. In isotropic media, $A = 2/3$ and $B = 1/3$ \cite{12}. This is the case with our
217: silicate fibers. $g$ is the gain coefficient (in $\rm m^{-1}$) and $\omega_g$
218: is the spectral gain bandwidth (in $\rm ps^{-1}$).
219: At first order the gain coefficient is fixed by the fact that it compensates the losses.
220: Note that polarization mode dispersion is not taken into account since the cavity length is short (about 9 m).
221: For numerical simulations we will use the following values for the parameters: $K = 1.5\rm \; m^{-1}$,
222: $ \beta_2 = 0.026\;\rm ps^2/m$, $L = 9\;\rm m$, $\gamma = 3\cdot 10^{-3}\;\rm W^{-1} m^{-1}$ and $\omega_g = 10^{13}\;\rm s^{-1}$.
223:
224: We now assume that the GVD $\beta_2$, the nonlinear coefficient $\gamma$, the gain filtering
225: $\rho={g}/{\omega_g^2}$ are small over one round-trip of the cavity. We introduce a small parameter $\e$ and replace
226: these quantities by $\e\beta_2$, $\e\gamma$ and $\e\rho$. We then look for solutions of the system (\ref{1}-\ref{2}) under the
227: form of a power series in $\e$, as
228: \bq u=u_0+\e u_1+O\left(\e^2\right),\eq
229: \bq v=v_0+\e v_1+O\left(\e^2\right).\eq
230: At order $\e^0$ it yields
231: \bq u_0=\tilde u_0e^{(g-iK)z},\eq
232: \bq v_0=\tilde v_0e^{(g+iK)z},\eq
233: where $\tilde u_0$ and $\tilde v_0$ do not depend on $z$, {\it ie} depend on the time variable $t$ only.
234:
235: Making the transformation
236: \bq u_1=\tilde u_1(z,t)e^{(g-iK)z},\eq
237: \bq v_1=\tilde v_1(z,t)e^{(g+iK)z},\eq
238: the equations obtained at order $\e$ can be integrated with regard to $z$
239: to yield, for the component $u$:
240: \bq \ba{rl}
241: \displaystyle\tilde u_1=&
242: \displaystyle\tilde u_1(0)+z\left(\rho-\frac{i\beta_2}2\right)\partial_t^2\tilde u_0\\
243: &\displaystyle
244: +i\gamma\left(\tilde u_0\left|\tilde u_0\right|^2
245: +A\tilde u_0\left|\tilde v_0\right|^2\right)\frac{e^{2gz}-1}{2g}
246: +i\gamma B\tilde v_0^2\tilde u_0^\ast\frac{e^{(2g+4iK)z}-1}{2g+4iK}.\ea\label{3}\eq
247: To get the complete expression for $u(z)$, we notice that
248: \bq u(0)=\tilde u_0+\e \tilde u_1(0)+O\left(\e^2\right).\eq
249: Then the two components of the wave amplitude can be written as
250: \bq \ba{rl}
251: \displaystyle u(z)=&
252: \displaystyle u(0)e^{(g-iK)z}+\e\biggl[ z\left(\rho-\frac{i\beta_2}2\right)\partial_t^2u(0)
253: \\
254: &\displaystyle+i\gamma\left(u(0)\left|u(0)\right|^2
255: +A u(0)\left| v(0)\right|^2\right)\frac{e^{2gz}-1}{2g}
256: \\
257: &\displaystyle\hspace{1.5cm}
258: +i\gamma B\tilde v(0)^2\tilde u(0)^\ast\frac{e^{(2g+4iK)z}-1}{2g+4iK}\biggr]e^{(g-iK)z}
259: +O\left(\e^2\right),\ea
260: \label{4}\eq
261: \bq \ba{rl}
262: \displaystyle v(z)=&
263: \displaystyle v(0)e^{(g+iK)z}+\e\biggl[ z\left(\rho-\frac{i\beta_2}2\right)\partial_t^2v(0)
264: \\
265: &\displaystyle+i\gamma\left(v(0)\left|v(0)\right|^2
266: +A v(0)\left| u(0)\right|^2\right)\frac{e^{2gz}-1}{2g}
267: \\
268: &\displaystyle\hspace{1.5cm}
269: +i\gamma B\tilde u(0)^2\tilde v(0)^\ast\frac{e^{(2g-4iK)z}-1}{2g-4iK}\biggr]e^{(g+iK)z}
270: +O\left(\e^2\right).\ea
271: \label{5}\eq
272:
273:
274:
275:
276: \subsubsection{Polarizer}
277: Figure \ref{angle} shows the respective orientation of the eigenaxis of the fiber before and after the polarizer (optical isolator).
278: $(u_-,v_-)$ are the field components just before the polarizer, $(u_+,v_+)$ just after it.
279: \begin{figure}[hbt!]
280: \begin{center}
281: \includegraphics[width=6cm]{angle.EPS}
282: \caption{\footnotesize Definition of the angles $\theta_-$ and $\theta_+$.}
283: \label{angle}
284: \end{center}
285: \end{figure}
286: The effect of the polarizer can be written as
287: \bq \left(\ba{c}u_+\\v_+\ea\right)=\beta\left(\ba{c}\cos\theta_+\\\sin\theta_+\ea\right)
288: \left(\cos\theta_-\;\;\;\sin\theta_-\right)\left(\ba{c}u_-\\v_-\ea\right),\label{6}\eq
289: where $\beta$ is its transmission coefficient ($\beta < 1$). Immediately after the polarizer,
290: the field has a well-defined linear polarization. Denote by $f_n$ its amplitude at the beginning of the $n^{\rm th}$ round trip.
291: Equation (\ref{5}) can be written as
292: \bq
293: f_{n+1}=\beta\left(\cos\theta_-\;\;\;\sin\theta_-\right)
294: \left(\ba{c}u_ {-,n}\\v_{-,n}\ea\right),\label{7}\eq
295: and
296: \bq \left(\ba{c}u_{+,n+1}\\v_{+,n+1}\ea\right)
297: =\left(\ba{c}\cos\theta_+\\\sin\theta_+\ea\right)f_{n+1},\label{8}\eq
298: where $(u_n, v_n)$ are the field components during the $n^{\rm th}$ round.
299: The field components $u_{+n}=u_n(0)$ and $v_{+n}=v_n(0)$, at the entrance of the fiber immediately after the polarizer,
300: are transformed respectively into $u_n(L)=u_{-n}$ and $v_n(L)=v_{-n}$, at
301: the end of the fiber immediately before the polarizer
302: ($L$ is the fiber length).
303:
304: $u_n(0)$ and $v_n(0)$ are proportional to $f_n$. Using the propagation formulas
305: (\ref{4}) and (\ref{5}), $u_n(L)$ and $v_n(L)$ can be computed.
306: Then $f_{n+1}$ is computed as a function of $f_n$ using (\ref{7}). This yields
307: \bq
308: f_{n+1}=\beta e^{gL}\left\{Q
309: f_n
310: +\e\left[\left(\rho-\frac{i\beta_2}2\right)L
311: Q\partial_t^2f_n
312: +i P f_n\left|f_n\right|^2\right]\right\}
313: +O\left(\e^2\right),\label{9}\eq
314: with
315: \bq Q=\cos(\theta_+-\theta_-)\cos KL-i\cos(\theta_++\theta_-)\sin KL,\label{10}\eq
316: and
317: \bq
318: \ba{r}
319: P=
320: \gamma\displaystyle\biggl(\frac{e^{2gL}-1}{2g}\biggl[
321: \cos(\theta_+-\theta_-)\cos KL-i\cos(\theta_++\theta_-)\sin KL\hspace{1.1cm}\vspace{1.5mm}\\
322: \displaystyle+\frac{A-1}2\sin2\theta_+
323: \left[\sin(\theta_++\theta_-)\cos KL-i\sin(\theta_+-\theta_-)\sin KL\right]\biggr]
324: \hspace{.5cm}\vspace{1.5mm}\\
325: \displaystyle + \frac B2\sin2\theta_+\biggl[\sin\theta_+\cos\theta_-e^{-iKL}
326: \frac{e^{(2g+4iK)L}-1}{2g+4iK}\hspace{.5cm}\vspace{1.5mm}\\
327: \displaystyle
328: +\cos\theta_+\sin\theta_-e^{iKL}
329: \frac{e^{(2g-4iK)L}-1}{2g-4iK}\biggr]\biggr).\ea\label{11}\eq
330:
331:
332:
333: \subsubsection{Gain threshold and continuous limit for $f_n$}
334:
335:
336: The dominant behavior of $\left(f_n\right)$ is given by the zero
337: order term in (\ref{9}):
338: \bq f_{n+1}=\beta e^{gL}Q
339: f_n
340: +O\left(\e\right).\label{12}\eq
341: A steady state over a large number of round trips could be reached if this first order evolution
342: is the multiplication by a phase factor, but only in this case.
343: Therefore the modulus of $\beta e^{gL}Q$ must be 1. This gives the threshold gain value:
344: \bq\ba{rl}g_0=&\displaystyle\frac{-1}{2L}\ln\left(\beta^2\left|Q\right|^2\right)\vspace{1.5mm}\\
345: =&\displaystyle
346: \frac{-1}{2L}\ln\left(\beta^2\left[\cos^2(\theta_+-\theta_-)
347: -\sin2\theta_+\sin2\theta_-\sin^2KL\right]\right)
348: .\ea\label{13}\eq
349: Denote by $e^{i\psi}$ the quantity $\beta e^{g_0L}Q$.
350: The exact condition is $\left|\beta e^{gL}Q\right|=1+O\left(\e\right)$,
351: the gain $g$ therefore writes $g=g_0+\e g_1$, in which $g_1$ is still free.
352: Expanding $e^{\e g_1L}$ in a power series of $\e$, equation (\ref{9}) becomes
353: \bq
354: f_{n+1}=e^{i \psi}\left(1+\e g_1L\right)f_n
355: +\e\left(\rho-\frac{i\beta_2}2\right)Le^{i \psi}\partial_t^2f_n
356: +i \e \frac{Pe^{i\psi}}Q
357: f_n\left|f_n\right|^2+O\left(\e^2\right).\label{14}\eq
358:
359:
360:
361: Consider now some function $f$ of a continuous variable $z$, obeying an equation of the form
362: \bq
363: i\partial_zf=\left({\cal A}+i\e{\cal B}\right)f+\e{\cal C}\partial_t^2f
364: +\e{\cal D}f\left|f\right|^2,\label{15}\eq
365: where $\cal A$ and $\cal B$ are real, and $\cal C$ and $\cal D$ complex coefficients.
366: We look for solutions under the form $f=f_0+\e f_1+\ldots$. The solution at first order is
367: \bq f_0=\tilde f_0 e^{-i{\cal A}z}.\label{16}\eq
368:
369: We write
370: \bq f_1=\tilde f_1(z,t) e^{-i{\cal A}z},\label{16b}\eq
371: and obtain:
372: \bq f(L)=\left[f(0)+\e\left({\cal B}f(0)-i{\cal C}\partial_t^2f(0)-
373: i{\cal D}f(0)\left|f(0)\right|^2\right)L\right]e^{-i{\cal A}L}+O\left(\e^2\right).
374: \label{17}\eq
375: Equations (\ref{14}) and (\ref{17}) are identified according to
376: $f(0)\equiv f_{n}$ and $f(L)\equiv f_{n+1}$,to yield
377: \bq e^{-i{\cal A}L}=e^{i\psi},\label{18}\eq
378: \bq {\cal B}=g_1,\label{19}\eq
379: \bq {\cal C}=\frac{\beta_2}2+i\rho,\label{20}\eq
380: \bq {\cal D}=\frac{-P}{QL}.\label{21}\eq
381:
382:
383: Then the continuous equation (\ref{15}) yields some interpolation of
384: the discrete sequence $\left(f_n\right)$.
385: £££
386: The continuous approximation is relevant when the number of round trips is very large;
387: further, mode-locked pulses actually correspond to this situation.
388: The small correction of order $\e$ in equations (\ref{15}) or (\ref{17})
389: gives account for the variations of $f$ on `propagation distances', which are here number of round trips,
390: very large, of order $1/\e$.
391: This can be shown in a rigorous way using the multiscale formalism, commonly used for the derivation of
392: the model equations in the soliton theory \cite{tan67a}.
393: We introduce a slow variable $\zeta=\e z$, in such a way that
394: \bq\partial_z=\partial_{\hat z}+\e \partial_{\zeta}.\label{22}\eq
395: The values of $\zeta$ about 1 correspond to number of round trips about $1/\e$.
396: $f$ is expanded as above as $f_0+\e f_1+\ldots$, and the
397: first order obviously yields (\ref{16}), we make the transform (\ref{16b})
398: and $\tilde f_1$ must satisfy the following equation:
399: \bq i\partial_{\zeta}\tilde f_0+i\partial_{\hat z}\tilde f_1=
400: i{\cal B}\tilde f_0+{\cal C}\partial_t^2\tilde f_0+
401: {\cal D}\tilde f_0\left|\tilde f_0\right|^2.\label{23}\eq
402: Equation (\ref{23}) can be written as
403: \bq \partial_{\hat z}\tilde f_1={\cal F}(\tilde f_0).\eq
404: The long distance evolution of $\tilde f_0$ is obtained from the requirement
405: that $\tilde f_1$ does not grow linearly with $z$.
406: This yields $\partial_{\hat z}\tilde f_1=0$ and
407: \bq i\partial_{\zeta}\tilde f_0=
408: i{\cal B}\tilde f_0+{\cal C}\partial_t^2\tilde f_0+
409: {\cal D}\tilde f_0\left|\tilde f_0\right|^2.\label{24}\eq
410: Or, using real coefficients only:
411: \bq i\partial_{\zeta}\tilde f_0=
412: i g_1\tilde f_0+\left(\frac{\beta_2}2+i\rho\right)\partial_t^2\tilde f_0+
413: \left({\cal D}+i{\cal D}_i\right)\tilde f_0\left|\tilde f_0\right|^2.\label{24bis}\eq
414: Equation (\ref{24}) or (\ref{24bis}) is the cubic complex Ginzburg-Landau (CGL) equation. Note that equation
415: (\ref{24}) is formally identical to the master equation proposed by Haus {\it et al} \cite{7}. However,
416: the coefficients of (\ref{24}) explicitly depend on the orientation of the eigenaxis of the fiber at both sides of the polarizer.
417: An essential feature is the arising of a nonlinear gain or absorption ${\cal D}_i$, which results from the combined effects of
418: the nonlinear rotation of the
419: polarization, the losses due to the polarizer, and the linear gain. The value and the sign of ${\cal D}_i$ depend on the
420: angles $\theta_+$ and $\theta_-$ between the eigenaxis of the fiber and the polarizer.
421:
422: \subsection{Different regimes for the CGL equation}
423: \subsubsection{The stationary solution and its modulational instability}
424: The CGL equation (\ref{24bis}) admits the following nonzero stationary solutions, {\it ie} with a constant modulus:
425: \bq f_1=A e^{i\left(\kappa\zeta-\Omega t\right)}\label{n1}\eq
426: with
427: \bq \Omega^2=\frac1\rho\left({\cal D}_i\vert A\vert^2+g_1\right)\label{n2}\eq
428: and
429: \bq \kappa=\frac{\beta_2}{2\rho}\left({\cal D}_i\vert A\vert^2+g_1\right)-{\cal D}_r\vert A\vert^2.\label{n3}\eq
430: In the particular case $\Omega=0$ it is a constant solution (independent of $t$),
431: with a fixed amplitude
432: \bq A=\sqrt{\frac{-g_1}{{\cal D}_i}},\label{n4}\eq
433: and \bq \kappa=\frac{{\cal D}_r}{{\cal D}_i}g_1.\label{n5}\eq
434: It exists only if ${\cal D}_ig_1<0$, {\it ie} if the excess of linear gain $g_1$
435: and the effective nonlinear gain ${\cal D}_i$ have opposite signs.
436: We perform a linear stability analysis for the constant solution $f_c=Ae^{i\kappa\zeta}$.
437: We seek for solutions of the form $f_1=f_c\left(1+u\right)$, $u$ being very small. It is seen that it must satisfy the
438: following equation:
439: \bq -\kappa u+i\partial_\zeta u=ig_1 u+\left(\frac{\beta_2}2+i\rho\right)\partial_t^2u+{\cal D}A^2\left(2u+u^\ast\right).\label{n6}\eq
440: Due to the term $u^*$, equation (\ref{n6}) is not linear.
441: The eventual instability is of the type discovered first by Benjamin and Feir in the frame of water waves \cite{ben67}.
442: It has been shown that this kind of instability involves
443: the nonlinear interaction of two modulation terms
444: with conjugated phases with the square $A^2$ of the constant solution \cite{stu78}.
445: Thus we write $u$ as:
446: \bq u=u_1e^{\left(\lambda \zeta-i\omega t\right)}+u_2e^{\left(\lambda^\ast \zeta+i\omega t\right)}\label{n7}\eq
447: It is found that $\lambda$ satisfies an equation of the form
448: \bq \lambda^2+2b\lambda+c=0.\label{n8}\eq
449: Using the expression (\ref{n4}) of the fixed amplitude $A$, we get the following values of the real constants
450: $b$ and $c$:
451: \bq b=g_1+\rho\omega^2,\label{n9}\eq
452: \bq c=\left(g_1\frac{{\cal D}_r}{{\cal D}_i}+\frac{\beta_2}2\omega^2\right)^2+b^2-
453: g_1^2\left(1+\frac{{\cal D}_r^2}{{\cal D}_i^2}\right).\label{n10}\eq
454: A short analysis of equation (\ref{n8}) show that a necessary condition for $u$ to remain bounded is that
455: $b$ is positive. This proves that modulational instability occurs as soon as the excess of linear gain $g_1$ is negative
456: and the nonlinear gain ${\cal D}_i$ is positive. Recall indeed that the existence of the constant nonzero solution
457: requires that these two quantities have opposite signs.
458: When $g_1>0$ and ${\cal D}_i<0$, we show that the modulational instability never occurs.
459: Numerical computation shows indeed that the quantity $\left({\cal D}_r\beta_2/2+{\cal D}_i\rho\right)$ is always negative,
460: which ensures, using elementary analysis, that $c$ is always positive, and the result follows.
461:
462:
463: If we admit that the excess of linear gain $g_1$ will self-adjust to a value for which a stable solution exists,
464: the above discussion shows that the stability of the constant solution depends on the sign of the
465: effective nonlinear gain ${\cal D}_i$.
466: It is stable when ${\cal D}_i<0$ only.
467: We expect to observe continuous laser emission when the constant solution is stable,
468: and only in this case. Thus the domains of continuous emission should coincide with the regions where ${\cal D}_i$ is
469: negative. The sign of ${\cal D}_i$ as function of the angles $(\frac{-1}2\theta_-,\frac12\theta_+)$ is drawn on figure \ref{ddi}.
470: It is easily checked that the relation between the angles $\theta_1$, $\theta_2$ used in the experiments and the
471: angles $\theta_+$, $\theta_-$ defined by figure \ref{angle} is such that $\theta_1=\frac{-1}2\theta_-+\theta_{10}$ and
472: $\theta_2=\frac{1}2\theta_++\theta_{20}$, with some fixed value of $\theta_{10}$ and $\theta_{20}$.
473: \begin{figure}[hbt!]
474: \begin{center}
475: \includegraphics[width=9cm]{ddi.EPS}
476: \caption{\footnotesize Sign of the effective nonlinear gain ${\cal D}_i$ as a function of the orientation
477: of the polarizer. White: ${\cal D}_i<0$, and the constant solution of the CGL equation is stable. Gray:
478: ${\cal D}_i>0$, and the constant solution of the CGL equation is unstable.}
479: \label{ddi}
480: \end{center}
481: \end{figure}
482: The white domain on figure \ref{ddi} corresponds to the negative values of the nonlinear gain ${\cal D}_i$. It is thus the domain where the constant solution
483: of the CGL equation is stable. Continuous laser emission should occur in this domain.
484: \subsubsection{Localized solutions}
485: A localized analytical solution of the CGL equation (\ref{24bis}) can also be written \cite{sot97a}. It has the following expression:
486: \bq f_1=a(t) e^{i\left(d \ln a(t)-\omega \zeta\right)}\label{n11},\eq
487: with \bq a(t)=BC\mbox{sech }\left(Bt\right)\label{n12}\eq
488: and \bq B=\sqrt{\frac{g_1}{\rho d^2-\rho-\beta_2 d}},\label{n13}\eq
489: \bq C=\sqrt{\frac{3d\left(4\rho^2+\beta_2^2\right)}{2\left(\beta_2 {\cal D}_i-2\rho {\cal D}_r\right)}},\label{n14}\eq
490: \bq d=\frac{3\left(\beta_2{\cal D}_r+2\rho {\cal D}_i\right)+\sqrt{9\left(\beta_2{\cal D}_r+2\rho {\cal D}_i\right)^2+8
491: \left(\beta_2{\cal D}_i-2\rho {\cal D}_r\right)^2}}{2\left(\beta_2{\cal D}_i-2\rho {\cal D}_r\right)},\label{n15}\eq
492: \bq\omega=\frac{-g_1\left(4\rho d+\beta_2 d^2-\beta_2\right)}
493: {2\left(\rho d^2-\rho-\beta_2 d\right)}.\label{n16}\eq
494: The inverse $B$ of the pulse length is real only if the quantity ${\cal T}=\left(\rho d^2-\rho-\beta_2 d\right)$ and the
495: excess of linear gain $g_1$ have the same sign.
496: The regions where ${\cal T}$ is either positive or negative are specified on figure \ref{pltruc2}.
497: \begin{figure}[hbt!]
498: \begin{center}
499: \includegraphics[width=9cm]{pltruc2.EPS}
500: \caption{\footnotesize Sign of the excess of nonlinear gain value $g_1$ for which the exact localized solution exists,
501: as a function of the orientation
502: of the polarizer. White: ${\cal T}$ and $g_1$ positive. Gray:
503: ${\cal T}$ and $g_1$ negative.}
504: \label{pltruc2}
505: \end{center}
506: \end{figure}
507: The background with zero amplitude is stable when the excess of linear gain $g_1$ is negative, and unstable in the opposite case.
508: When $g_1>0$, the exact localized solution (\ref{n11}) is unstable due to the instability of the background \cite{sot97a}.
509: Qualitatively, localized pulse formation can be expected
510: when the excess of linear gain $g_1$ is negative and the nonlinear gain ${\cal D}_i$ is positive, as is suggested on figure~\ref{pulse}.
511: \begin{figure}[hbt!]
512: \begin{center}
513: \includegraphics[width=9cm]{pulse.EPS}
514: \caption{\footnotesize Schematic representation of the effect of the nonlinear gain ${\cal D}_i$, of the excess of
515: linear gain $g_1$, and of both the dispersion $\beta_2$ and the effective self-phase modulation ${\cal D}_r$ on a localized pulse.}
516: \label{pulse}
517: \end{center}
518: \end{figure}
519: The effective self-phase modulation ${\cal D}_r$ is always negative, and the dispersion $\beta_2$ is positive.
520: Therefore their conjugated effect
521: leads to the increase of the pulse width.
522: If $g_1$ is positive and ${\cal D}_i$ negative (figure \ref{pulse}a), the nonconservative effects decrease the amplitude at the top of the pulse, and
523: increase it at the bottom: no stable localized pulse can be formed. If on the contrary $g_1$ is negative and ${\cal D}_i>0$
524: (figure \ref{pulse}b),
525: energy appears at the
526: top of the pulse and disappears at its bottom, yielding some pulse narrowing, which could be expected to balance the
527: broadening caused by the nonlinear index variations.
528: In fact the analytical localized solution (\ref{n11}) is never stable when $\beta_2 {\cal D}_r<0$, but can be
529: stabilized when higher order nonlinear terms are taken into account, as shows the study of the so-called quintic CGL equation
530: \cite{sot97a}. The stabilizing term should be a quintic nonlinear absorption.
531: Observe that here the effective nonlinear gain ${\cal D}_i$
532: follows from the linear gain and the nonlinear evolution of the polarization in the fiber, without
533: considering cubic nonlinear gain in the initial model
534: (\ref{1}-\ref{2}). In an analogous way, an effective
535: quintic nonlinear gain term in the master equation (\ref{24}) does not require that a quintic nonlinear gain term
536: is present in the initial model, either a cubic nonlinear gain or a quintic nonlinear index should be enough.
537: (\ref{1}-\ref{2}).
538: The precise determination of this higher order term
539: is left for further study. On these grounds, the exact solution (\ref{n11}) corresponds to a potentially stable pulse
540: when $g_1 <0$ and ${\cal D}_i>0$, and to a completely unstable pulse when $g_1 >0$ and ${\cal D}_i<0$.
541:
542: \subsubsection{The different regimes}
543: According to the conclusions of the previous section, and assuming that
544: the excess of linear gain $g_1$ will self-adjust
545: to the value for which a stable solution exists, mode-locked laser emission can be expected
546: in the region where ${\cal D}_i>0$ and ${\cal T}<0$. This is the domain in dark gray marked ML on figure \ref{concl}.
547: In the region where ${\cal D}_i$ is negative, the constant solution of the CGL equation is stable, and
548: continuous laser emission is expected to occur. It is the white domain marked cw on the figure \ref{concl}.
549: In the region where ${\cal T}$ is positive and ${\cal D}_i$ negative,
550: no stable localized neither constant
551: solution exist. The laser behavior is expected to be unstable in this region, in light gray on the figure \ref{concl}.
552: The theoretical results summarized on the figure \ref{concl} can be compared to the experimental results of figure \ref{figs2}.
553: \begin{figure}[hbt!]
554: \begin{center}
555: \includegraphics[width=9cm]{CONCL3B.EPS}
556: \caption{\footnotesize Operating regime of the laser as a function of the angles $\theta_+$ and $\theta_-$ between the polarizer and the
557: eigenaxis of the fiber, according to the theory. ML stands for mode-locked regime and CW for continuous wave regime.}
558: \label{concl}
559: \end{center}
560: \end{figure}
561: A discrepancy between the two figures appears at first glance: the periodicity of the
562: behavior regarding the variable $\theta_2$ or $\theta_+/2$ seems to be 45 degrees according to the theory,
563: while it is 90 degrees according to the experiment. Indeed,
564: the two elongated domains of mode-locked behavior are not equivalent according to experimental observation,
565: while they seem to be identical on the theoretical figure \ref{concl}.
566: \begin{figure}[hbt!]
567: \begin{center}
568: \includegraphics[width=9cm]{codi.EPS}
569: \caption{\footnotesize Contour plot of the effective nonlinear gain ${\cal D}_i$
570: as a function of the angles $\theta_+$ and $\theta_-$ between the polarizer and the
571: eigenaxis of the fiber. }
572: \label{codi}
573: \end{center}
574: \end{figure}
575: A contour plot of the effective nonlinear gain ${\cal D}_i$ as a function of the angles is drawn on figure \ref{codi}.
576: It is seen that the values of ${\cal D}_i$ are not the same in the two theoretical domains of mode-locked emission.
577: Thus they are in fact not equivalent, and the periodicity of the theoretical result is 90 degrees, as that of the
578: experimental one. Further it has been shown \cite{sot97a} that an excessive value of the nonlinear gain might
579: prevent pulse stabilization.
580: A more accurate analysis would thus very likely show that
581: the regions where ${\cal D}_i$ takes its
582: highest values are out of the domain of stability of the localized pulse.
583: This confirms that the two horizontal domains of mode-locked regime drawn on figure \ref{concl} are not equivalent.
584: Further, it could explain a part of the discrepancy between theoretical and experimental results.
585:
586: Before we conclude, let us consider the influence of the group velocity dispersion
587: on the stability of the mode-locking solutions. This is important because GVD compensation
588: is required to generate subpicosecond pulses. Experimentally, we have obtain
589: 666 fs pulses with a grating pair inserted in the cavity \cite{6,11bis}.
590: Under these conditions, the total GVD is about $\beta_2 = 0.0005\,\rm ps^2/m$.
591: Hence, stable ultrashort pulse generation is possible for such low values of the GVD.
592: We have tested our master equation for decreasing values of $\beta_2$.
593: We have found that a lowering of $\beta_2$ results in a reduction of the
594: stable mode-locking regions. In addition, mode-locking completely disappears for
595: $\beta_2\lesssim 0.0015 \,\rm ps^2/m$.
596: These results are in agreement with the predictions of the Haus's model \cite{7}
597: but not with our experimental results.
598: The validity of the model presented here is therefore limited to GVD values not
599: too close to zero. Higher order terms of the GVD could be required to improve the
600: model near the zero-dispersion point. This problem deserves further experimental
601: and theoretical work.
602:
603:
604:
605:
606: \section{Conclusion}
607: We have given a rigorous derivation of the master equation, a CGL equation, that allows to discuss
608: theoretically the different operating regimes of the passively mode-locked Yb-doped double-clad fiber laser.
609: This equation involves a single amplitude, but gives an account
610: of the evolution of the polarization inside the fiber through its coefficients. An effective nonlinear
611: gain arises from the conjugated effects of the linear gain and the self-phase modulation. Its
612: sign depends on the orientation of the polarizer.
613: The existence and stability of the constant nonzero solution of the CGL equation has been discussed. It gives
614: an account of the continuous emission regime of the laser. An analytical localized solution has also been given.
615: Although it is properly speaking unstable, it allows to discuss the formation of localized pulses, {\it ie} the
616: mode-locked regime of the laser. An unstable regime has also been identified. These theoretical results are in good accordance
617: with the experiments. The study of higher order corrections to this model should improve these results.
618: \newpage
619: \begin{thebibliography}{References}
620:
621: \bibitem{1} K. Tamura, H.A. Haus and E.P. Ippen, Self-starting additive pulse mode-locking erbium fiber ring laser,
622: {\it Electron. Lett.} {\bf 28}, 2226-2228, 1992.
623: \bibitem{2} H.A. Haus, K. Tamura, L.E. Nelson and E.P. Ippen,
624: Stretched-pulse additive pulse mode-locking in fiber ring lasers:
625: theory and experiment, {\it IEEE Jour. Quant. Electron.} {\bf 31}, 591-598, 1995.
626: \bibitem{3} V.J. Matsas, T.P. Newson, D.J. Richardson and D.N. Payne,
627: Selfstarting passively mode-locked finer ring soliton laser exploiting nonlinear polarization rotation,
628: {\it Electron. Lett. } {\bf 28}, 1391-1393, 1992.
629: \bibitem{4} V. Cautaerts, D.J. Richardson, R. Pashotta and D.C. Hanna, Stretched pulse Yb3+:silica fiber laser,
630: {\it Opt. Lett.} {\bf 22}, 316-318, 1997.
631: \bibitem{5} R. Hofer, M. Hofer and G.A. Reider,
632: High energy, sub-picosecond pulses from a Nd-doped double-clad fiber laser,
633: {\it Optics Com.} {\bf 169}, 135-139, 1999.
634: \bibitem{6} A. Hideur, T. Chartier, M. Brunel, S. Louis, C. \"Ozkul and F. Sanchez,
635: Generation of high energy femtosecond pulses from a side-pumped cYb-doped double-clad fiber laser,
636: {\it Appl. Phys. Lett.} {\bf 79}, 3389 (2001).
637: \bibitem{7} H.A. Haus, J.G. Fujimoto and E.P. Ippen, Structures for additive pulse mode locking,
638: {\it J. Opt. Soc. Am. B} {\bf 8}, 2068-2076, 1991.
639: \bibitem{8} A.D. Kim, J.N. Kutz and D.J. Muraki,
640: Pulse-train uniformity in optical fiber lasers passively mode-locked by nonlinear polarization rotation,
641: {\it IEEE Jour. Quant. Electron. } {\bf 36}, 465-471, 2000.
642: \bibitem{9} A.Hideur, T.Chartier, M.Brunel, M.Salhi, C.Ozkul and F.Sanchez,
643: Mode-lock, Q-switch and cw operation of an Yb-doped double-clad fiber ring laser, {\it Optics Com.} {\bf 198}, 141-146, 2001.
644: \bibitem{10} T. Chartier, A. Hideur, C. \"Ozkul, F. Sanchez, G. Stephan,
645: Polarization properties of single-mode optical fibers with random fluctuations of birefringence,
646: {\it Appl. Optics.} {\bf 40}, 5343-5353 (2001).
647: \bibitem{11} M.L. Dennis and I.N. Duling III,
648: Experimental study of sideband generation in femtosecond fiber lasers,
649: {\it IEEE Jour. Quant. Electron.} {\bf 30}, 1469-1477, 1994.
650: \bibitem{11bis} A. Hideur, T. Chartier, C. \"Ozkul, M. Brunel and F. Sanchez,
651: Experimental study of pulse compression in a side-pumped Yb-doped
652: double-clad mode-locked fiber laser, {\it Appl.
653: Phys. B} {\bf 74}, 121-124 (2002).
654: \bibitem{12} G.P. Agrawal, {\it Nonlinear Fiber Optics}, Academic Press, Second Edition, 1995.
655: \bibitem{13} D.J. Ripin and L. Goldberg,
656: High efficiency side-coupling of light into optical fibers using embedded V-grooves,
657: {\it Electron. Lett.} {\bf 31}, 2204-2205, 1995.
658: \bibitem{tan67a}
659: T. Taniuti and C.-C. Wei, Reductive perturbation method in nonlinear
660: wave propagation I. {\it J. Phys. Soc. Japan}, {\bf 24}, 941-946 (1968).
661: T. Taniuti and N. Yajima,
662: Perturbation method for a nonlinear wave modulation. III
663: {\it J. Math. Phys.}, {\bf 14} (10), 1389-1397 (1973).
664: \bibitem{ben67}
665: T.B. Benjamin and J.E. Feir, The desintegration of wave
666: trains in
667: deep water. Part.1. Theory,
668: {\it J. Fluid. Mech.} {\bf 27} (3), 417-430 (1967). \\
669: \bibitem{stu78}
670: J.T. Stuart, F.R.S. and R.C. DiPrima,
671: The Eckhaus and Benjamin-Feir resonance mechanisms,
672: {\it Proc. R. Soc. Lond. A} {\bf 362}, 27-41 (1978).
673: \bibitem{sot97a}
674: J.M. Soto-Crespo, N.N. Akhmediev, V.V. Afanasjev, and S.Wabnitz,
675: Pulse solutions of the cubic-quintic complex Ginzburg-Landau equation in the case
676: of normal dispersion,
677: {\it Phys. Rev. E} {\bf 55} (4), 4783-4796 (1997).
678: \end{thebibliography}
679:
680:
681: \end{document}
682: