math0310095/lms.tex
1: \documentclass{lms}
2: 
3: %% CARACTERES
4: \usepackage{amssymb,amsmath} 
5: \usepackage[mathscr]{euscript} 
6: \usepackage[all]{xy}
7: 
8: \newcommand{\R}{\ensuremath{\mathbb{R}}}
9: \newcommand{\C}{\ensuremath{\mathbb{C}}}
10: \newcommand{\N}{\ensuremath{\mathbb{N}}}
11: \newcommand{\Z}{\ensuremath{\mathbb{Z}}}
12: \renewcommand{\O}{\mathcal{O}}
13: \renewcommand{\Re}{\mathop{\rm Re}} 
14: \renewcommand{\Im}{\mathop{\rm Im}}
15: \newcommand{\tr}{\mathop{\rm tr}} \newcommand{\Ker}{\mathop{\rm Ker}}
16: \newcommand{\iN}{\hbox{ {\leaders\hrule\hskip.2cm}{\vrule height .22cm} }}
17: \newcommand{\pa}[3][]{ \frac{ \partial^{#1} {#2} }{ \partial {#3}^{#1} } }
18: \renewcommand{\t}[1]{ {}^t \!  #1 } 
19: \newcommand{\at}[1]{\big|_{#1}}
20: \newcommand{\ds}{{\displaystyle } } 
21: \newcommand{\g}{\ensuremath{\mathfrak{g}}}
22: \newcommand{\gC}{\ensuremath{\mathfrak{g}^{\C}}}
23: \renewcommand{\u}{\ensuremath{\mathfrak{u}}}
24: \newcommand{\su}{\ensuremath{\mathfrak{su}}}
25: \renewcommand{\sl}{\ensuremath{\mathfrak{sl}}} \renewcommand{\phi}{\varphi}
26: \newcommand{\Ad}{\mathop{\rm Ad}} \newcommand{\ad}{\mathop{\rm ad}}
27: \renewcommand{\.}{\cdot} \newcommand{\lf}{\hfill\break}
28: 
29: \newtheorem{theorem}{Theorem}[section]
30: \newtheorem{proposition}[theorem]{Proposition}
31: \newtheorem{lemma}[theorem]{Lemma} 
32: \newtheorem{corollary}[theorem]{Corollary}
33: 
34: \newnumbered{definition}[theorem]{Definition}
35: \newnumbered{example}[theorem]{Example}
36: \newunnumbered{remark}[theorem]{Remark}
37: 
38: \title{Hamiltonian stationary tori in the complex projective plane}
39: \author{Fr\'ed\'eric H\'elein \and Pascal Romon}
40: 
41: \classno{53C55 (primary), 53C42, 53C25, 58E12 (secondary)}
42: 
43: \begin{document}
44: \maketitle
45: 
46: 
47: 
48: \section{Introduction}
49: Hamiltonian stationary Lagrangian surfaces are Lagrangian surfaces of a given
50: four-dimensional manifold endowed with a symplectic and a Riemannian structure,
51: which are critical points of the area functional with respect to a particular
52: class of infinitesimal variations preserving the Lagrangian constraint: 
53: the compactly supported Hamiltonian vector fields.  The
54: Euler--Lagrange equations of this variational problem are highly simplified when
55: we assume that the ambient manifold ${\cal N}$ is K\"ahler.  In that case we can
56: make sense of a {\em Lagrangian angle function} $\beta$ along any
57: simply-connected Lagrangian submanifold $\Sigma \subset {\cal N}$ (uniquely
58: defined up to the addition of a constant).  And as shown in \cite{ScWo} the mean
59: curvature vector of the submanifold is then $\vec{H}= J\,\nabla \beta$, where
60: $J$ is the complex structure on ${\cal N}$ and $\nabla \beta$ is the gradient of
61: $\beta$ along $\Sigma$. It turns out that $\Sigma$ is Hamiltonian stationary if and
62: only if $\beta$ is a harmonic function on $\Sigma$.\\
63: 
64: \noindent A particular subclass of solutions occurs when $\beta$ is constant:
65: the Lagrangian submanifold is then simply a minimal one.  In the case where
66: ${\cal N}$ is a Calabi--Aubin--Yau manifold, such submanifolds admit an
67: alternative characterization as {\em special Lagrangian}, a notion which has
68: been extensively studied recently because of its connection with string theories
69: and the mirror conjecture, see \cite{SYZ}.\\
70: 
71: \noindent An analytical theory of two-dimensional Hamiltonian stationary
72: Lagrangian submanifolds was constructed by R. Schoen and J. Wolfson \cite{ScWo},
73: proving the existence and the partial regularity of minimizers.  In contrast our
74: results in the present paper rest on the fact that, for particular ambient
75: manifolds ${\cal N}$, Hamiltonian stationary Lagrangian surfaces are solutions
76: of an integrable system.  This was discovered first in the case when ${\cal
77: N}=\C^2$ in \cite{HR1} and \cite{HR2}.  In a subsequent paper \cite{HR3} we
78: proved that the same problem is also completely integrable if we replace $\C^2$
79: by any two-dimensional Hermitian symmetric space.  Among these symmetric spaces
80: one very interesting example is $\C P^2$, because any simply-connected
81: Lagrangian surface in $\C P^2$ can be lifted into a Legendrian surface in $S^5$.
82: Furthermore the cone in $\C^3$ over this Legendrian surface is actually a
83: singular Lagrangian three-dimensional submanifold in $\C^3$; and the cone in
84: $\C^3$ is Hamiltonian stationary if and only if the surface in $\C P^2$ is so.\\
85: 
86: \noindent A similar correspondence has been remarked and used in \cite{J},
87: \cite{McI3} and \cite{Has} in the case of minimal Lagrangian surfaces in $\C
88: P^2$ and allows these Authors to connect results on minimal Lagrangian surfaces
89: in $\C P^2$ \cite{Sh} to minimal Legendrian surfaces in $S^5$ \cite{MM} and
90: special Lagrangian cones in $\C^3$.\\
91: 
92: \noindent Our aim in this paper is the following:
93: \begin{itemize}
94: \item to expound in details the correspondence between Hamiltonian stationary
95: Lagrangian surfaces in $\C P^2$ and Hamiltonian stationary Legendrian surfaces
96: in $S^5$ and a formulation using a family of curvature free connections of this
97: integrable system (theorem \ref{1.6.theobis}).  
98: We revisit here the formulation given in \cite{HR3}, using {\em twisted loop groups}.  
99: Roughly speaking it rests on the identifications 
100: $\C P^2\simeq SU(3)/S(U(2)\times U(1))$ and $(S^5,\hbox{ contact structure}) \simeq
101: \left( U(3)/U(2)\times U(1), A^3_3=0\right)$, where $A^3_3$ is a component of
102: the Maurer--Cartan form.  We also show that this problem has an alternative
103: formulation, analogous to the theory of K. Uhlenbeck \cite{U} for harmonic maps
104: into $U(n)$, using {\em based loop groups}.
105: 
106: \item to define the notion of {\em finite type} Hamiltonian stationary
107: Legendrian surfaces in $S^5$: we give here again two definitions, in terms of
108: twisted loop groups (which is an analogue to the description of finite type
109: harmonic maps into homogeneous manifolds according to \cite{BP}) and in terms of
110: based loop groups (an analogue to the description of finite type harmonic maps
111: into Lie groups according to \cite{BFPP}).  We prove the equivalence between the
112: two definitions because we actually need this result for the following.  We
113: believe that this fact should be well known to some specialists in the harmonic
114: maps theory, but we did not find it in the literature.
115: 
116: \item we prove in theorem \ref{4.MainTheo} that all Hamiltonian stationary 
117: Lagrangian tori in $\C P^2$ (and
118: hence Hamiltonian stationary Legendrian tori in $S^5$) are of finite type.  This
119: is the main result of this paper.  Our proof focuses on the case of Hamiltonian
120: stationary tori which are not minimal, since the minimal case has been studied
121: by many authors (\cite{BFPP}, \cite{Sh}, \cite{MM}, \cite{Has},
122: \cite{McI1},\cite{McI2}, \cite{McI3}, \cite{J}).  The method here is adapted
123: from the similar result for harmonic maps into Lie groups in \cite{BFPP}.
124: However the strategy differs slightly: we use actually the two existing
125: formulations of finite type solutions, using twisted or based loop groups.  One
126: crucial step indeed is the construction of a {\em formal Killing field},
127: starting from a given torus solution.  This step can be slightly simplified here
128: in the twisted loop groups formulation, because the semi-simple element we start
129: with is then just constant.  However proving that the formal Killing field is
130: {\em adapted} requires more work in the twisted loop groups formulation
131: (actually we were not able to do it directly) than in the based loop groups
132: formulation; here we take advantage from the two formulations to avoid the
133: difficulties and to conclude.
134: 
135: \item lastly we give some examples of Hamiltonian stationary Legendrian tori in
136: $S^5$: we construct in theorem \ref{homogeneous} a family of solutions which are equivariant in some sense
137: under the action of the torus, that we call {\em homogeneous} Hamiltonian
138: stationary tori.  These are the simplest examples that one can build.
139: \end{itemize}
140: 
141: \noindent Let us add that the structure of the integrable system studied here
142: fits in a classification of elliptic integrable systems proposed by C.L. Terng
143: \cite{T}, as a 2nd $(U(3),\sigma,\tau)$-system\footnote{We have here exchanged
144: the notations $\sigma$ and $\tau$ with respect to \cite{T} in order to be
145: consistent with our notations in \cite{HR3}.}, where $\sigma$ is an involution
146: of $U(3)$ such that its fixed set is $U(3)^\sigma\simeq U(2)\times U(1)$ and
147: $U(3)/U(2)^\sigma\simeq \C P^2$ and $\tau$ is a 4th order automorphism (actually
148: $\tau^2=\sigma$) which encodes the symplectic structure on $\C P^2$ or the
149: Legendrian structure on $S^5$.\\
150: 
151: \noindent {\em Notations} --- For any matrix $M\in GL(n,\C)$, we denote by
152: $M^\dagger :=\,^t\overline{M}$.
153: 
154: 
155: 
156: \section{Geometrical description of Hamiltonian stationary Lagrangian surfaces
157: in $\mathbb{C}P^n$}
158: 
159: \subsection{The Lagrangian angle}
160: 
161: The complex projective space $\mathbb{C}P^n$ can be identified with the quotient
162: manifold $S^{2n+1}/S^1$. It is a complex manifold with complex structure $J$. We
163: denote by $\pi:S^{2n+1}\longrightarrow \mathbb{C}P^n$
164:  the canonical projection a.k.a. Hopf fibration, and equip $\mathbb{C}P^n$ with the
165: Fubini-Study Hermitian metric, denoted by $\langle \cdot , \cdot \rangle _{\mathbb{C}P^n} =
166: \langle \cdot , \cdot \rangle  - i\omega(\cdot , \cdot )$, where
167: $\langle \cdot , \cdot \rangle$ is a Riemannian metric and $\omega$ is the
168: K\"ahler form\footnote{note that the sign
169: convention may vary in the literature, e.g.\,some Authors use $\langle \cdot ,
170: \cdot \rangle _{\mathbb{C}P^n} = \langle \cdot , \cdot \rangle +
171: i\omega(\cdot , \cdot )$.}. 
172: For each $z\in S^{2n+1}$ we let ${\cal H}_z$ be the complex $n$-subspace 
173: in $T_{z} S^{2 n+1}\subset \mathbb{C}^{n+1}$ which is Hermitian orthogonal to $z$ 
174: (and hence to the fiber of $d\pi_{z}$). By construction of the Fubini-Study metric,
175: $d\pi_z:{\cal H}_z\longrightarrow T_{\pi(z)}\mathbb{C}P^n$ is 
176: an isometry between complex Hermitian spaces. We call the
177: subbundle ${\cal H}:= \cup_{z\in S^{2n+1}}{\cal H}_z$ of $T S^{2n+1}$
178: the {\em horizontal distribution}. It defines in a natural way a connection
179: $\nabla^\mathit{Hopf}\simeq \nabla ^H$ on the Hopf bundle 
180: $\pi:S^{2n+1}\longrightarrow \mathbb{C}P^n$, whose curvature is $2i\omega$. 
181: As a consequence \cite{Re2}:
182: \begin{proposition}\label{2.1.prop}
183: Let $\Omega$ be a {\em simply connected} open subset of $\mathbb{R}^n$ and 
184: $u:\Omega \longrightarrow \mathbb{C}P^n$
185: be a smooth Lagrangian immersion, i.e.\,such that $u^*\omega = 0$. Then there exists a lift
186: \begin{center}
187: \hskip2cm
188: \xymatrix{
189: & S^{2n+1} \ar@{->}[d]^\pi \\
190: \Omega \ar@{->}[ur]^{\widehat{u}} \ar@{->}[r]^u & \mathbb{C}P^n
191: }
192: \end{center}
193: such that $\left(u^*\nabla^H\right) \widehat{u} = 0$ (where $u^*\nabla^H$ is 
194: the pull-back by $u$ of the connection $\nabla^H$). This lift is unique up to multiplication
195: by a unit complex number. Moreover the pull-back by $\widehat{u}$ of the symplectic
196: form  $\omega$ on $\mathbb{C}^{n+1}$ vanishes; we say that $\hat{u}$ is 
197: \emph{Legendrian}.
198: \end{proposition}
199: 
200: Taking $u,\hat{u}$ as above, we define, for any orthonormal framing
201: $(e_{1},\ldots,e_{n})$ of $T\Omega$, the Lagrangian angle $\beta$ by
202: \[
203: e^{i \beta} = dz^1 \wedge \ldots \wedge dz^{n+1} \big( \hat{u}, d\hat{u}(e_{1}),
204: \ldots, d\hat{u}(e_{n}) \big) .
205: \]
206: which makes sense because $( \hat{u}, d\hat{u}(e_{1}),\ldots, d\hat{u}(e_{n}))$ is
207: a Hermitian-orthonormal frame, for any $x\in\Omega$. Furthermore, the result is
208: independent from the choice of the framing, and depends on the choice of the lift
209: $\hat{u}$ only through multiplication by a unit complex constant. Hence $\beta$ is
210: defined up to an additive constant and $d\beta$ is always well-defined along any
211: Lagrangian immersion $u$. Another characteristic property of the Lagrangian angle
212: relates it to the mean curvature vector field $\vec{H}$ along $u$:
213: \begin{equation}\label{2.2.H=JDb}
214: \vec{H} = {1\over n}J\nabla \beta
215: \end{equation}
216: or equivalently $d\beta = - \vec{H} \iN \omega$ (see~\cite{Br,Da} for details). 
217: 
218: \subsection{Hamiltonian stationary Lagrangian submanifolds}
219: %
220: A \emph{Hamiltonian stationary} Lagrangian submanifold $\Sigma$ in
221: $\mathbb{C}P^n$ is a Lagrangian submanifold which is a critical point of the
222: $n$-volume functional ${\cal A}$ under first variations which are {\em
223: Hamiltonian vector fields} with compact support.  This means that for any smooth
224: function with compact support $h\in {\cal
225: C}^\infty_c(\mathbb{C}P^n,\mathbb{R})$, we have
226: \[
227: \delta {\cal A}_{\xi_h}(\Sigma) := \int_\Sigma \left\langle \vec{H},
228: \xi_h\right\rangle_E d\hbox{vol} = 0,
229: \]
230: where $\xi_h$ is the Hamiltonian vector field of $h$, i.e.\,satisfies
231: $\xi_h \iN \omega +dh = 0$ or $\xi_h = J\nabla h$.  We also remark that if
232: $f\in {\cal C}^\infty_c(\Sigma,\mathbb{R})$, then there exist smooth extensions
233: with compact support $h$ of $f$, i.e.\,functions $h\in {\cal
234: C}^\infty_c(\mathbb{C}P^n,\mathbb{R})$ such that $h_{|\Sigma} = f$, and moreover
235: the normal component of $\left(\xi_h\right)_{|\Sigma}$ does not depend on the
236: choice of the extension $h$ (it coincides actually with $J\nabla f$, where
237: $\nabla$ is here the gradient with respect to the induced metric on $\Sigma$).
238: So we deduce from above that $\delta {\cal
239: A}_{\xi_h}(\Sigma) = {1\over n}\int_\Sigma \left\langle \nabla \beta, \nabla
240: f\right\rangle_E d\hbox{vol}$.  This implies the following.
241: \begin{corollary}
242: Any Lagrangian submanifold $\Sigma$ in $\mathbb{C}P^n$ is Hamiltonian stationary
243: if and only if $\beta$ is a harmonic function on $\Sigma$, i.e.
244: \[
245: \Delta_\Sigma \beta = 0.
246: \]
247: \end{corollary}
248: 
249: \noindent This theory extends to non simply connected surfaces $\Sigma$
250: with the following restrictions.  Let $\gamma$ be a homotopically non trivial
251: loop.  The Legendrian lift of $\gamma$ needs not close, so that in general its
252: endpoints $p_1, p_2 \in S^{2n+1}$ are multiples of each other by a factor $e^{i
253: \theta}$.  The same holds for the Lagrangian angle: $\beta ( p_2 ) \equiv \beta
254: ( p_1 ) + (n+1) \theta \bmod 2 \pi$ (since the tangent plane is also shifted by
255: the Decktransformation $z \longmapsto e^{i \theta} z$).  In particular $\beta$
256: is not always globally defined on surfaces in $\mathbb{C} P^n$ with non trivial
257: topology, unless the Legendrian lift is globally defined in $S^{2n+1} /
258: \mathbb{Z}_{n+1}$ (here $\mathbb{Z}_{n+1}$ stands for the $n+1$-st roots of
259: unity in $SU(n+1)$).
260: 
261: 
262: \subsection{Conformal Lagrangian immersions into $\mathbb{C}P^2$}
263: We now set $n= 2$.  We suppose that $\Omega$ is a simply connected open subset
264: of $\mathbb{R}^2\simeq \mathbb{C}$ and consider a conformal Lagrangian immersion
265: $u:\Omega\longrightarrow \mathbb{C}P^2$.  This implies that we can find a
266: function $\rho:\Omega\longrightarrow \mathbb{R}$ and two sections $E_1$ and
267: $E_2$ of $u^*T\mathbb{C}P^2$ such that $\forall (x,y)\in \Omega$,
268: $(E_1(x,y),E_2(x,y))$ is an Euclidean orthogonal basis over $\mathbb{R}$ of
269: $T_{u(x,y)}u(\Omega)$ and
270: \[
271: d u = e^\rho\left( E_1d x + E_2d y\right).
272: \]
273: We observe that, due to the fact that $u$ is Lagrangian, $(E_1,E_2)$ is a also a
274: {\em Hermitian} basis over $\mathbb{C}$ of $T_{u(x,y)}\mathbb{C}P^2$.\\
275: 
276: \noindent Let \xymatrix{ \Omega\ar@{->}[r]^{\widehat{u}}\ar@{->}[dr]^u & S^5
277: \ar@{->}[d]^\pi \\
278:  & \mathbb{C}P^2 } be a parallel lift of $u$ as in Proposition \ref{2.1.prop}
279:  and $(e_1,e_2)$ be the unique section of $\widehat{u}^*{\cal H}\times
280:  \widehat{u}^*{\cal H}$ which lifts\footnote{recall that the condition that
281:  $v\in \left(\widehat{u}^*{\cal H}\right)_{(x,y)}$ means that $v$ is in the
282:  horizontal subspace ${\cal H}_{\widehat{u}(x,y)}$} $(E_1,E_2)$.  Then we have
283: \begin{equation}\label{duchapeau}
284: d\widehat{u} = e^\rho\left( e_1d x + e_2dy\right).
285: \end{equation}
286: Note that $\forall (x,y)\in \Omega$, $(e_1(x,y),e_2(x,y))$ is a Hermitian basis
287: of ${\cal H}_{\widehat{u}(x,y)}$, which is Hermitian orthogonal to
288: $\widehat{u}(x,y)$.  Hence $\forall (x,y)\in \Omega$,
289: $(e_1(x,y),e_2(x,y),\widehat{u}(x,y))$ is a Hermitian basis of $\mathbb{C}^3$.
290: Thus this triplet can be identified with some $\widehat{F}(x,y)\in U(3)$.  We
291: hence get the diagram \xymatrix{ & U(3)\ar@{->}[d]^{(\cdot \cdot *)}\\
292: \Omega \ar@{->}[ur]^{\widehat{F}} \ar@{->}[r]^{\widehat{u}} \ar@{->}[dr]^u &
293: S^5\ar@{->}[d]^{\pi}\\
294: & \mathbb{C}P^2 }, where $(\cdot \cdot *)$ is the mapping
295: $(e_1,e_2,e_3)\longmapsto e_3$.\\
296: 
297: \noindent We define the Maurer--Cartan form $\widehat{A}$ to be the 1-form on
298: $\Omega$ with coefficients in $\u(3)$ such that $d\widehat{F} = \widehat{F}\cdot
299: \widehat{A}$.  Then we remark that the horizontality assumption $\langle
300: d\widehat{u},\widehat{u}\rangle _{\mathbb{C}^3} = 0$ exactly means that
301: \begin{equation}\label{A33=0}
302: \widehat{A}^3_3 = 0.
303: \end{equation}
304: Moreover the Lagrangian angle function $\beta_{\widehat{u}}$ along $\widehat{u}$
305: can be computed by
306: \[
307: e^{i\beta_{\widehat{u}}} = dz^1 \wedge dz^2 \wedge dz^3 (e_1,e_2,\widehat{u}) 
308: = \det \widehat{F}.
309: \]
310: As in \cite{HR1} we consider a larger class of framings of $u$ as follows.
311: \begin{definition}
312: A \emph{Legendrian framing of $u$ along $\widehat{u}$} is a map
313: $F:\Omega\longrightarrow U(3)$ such that
314: \begin{itemize}
315: \item $(\cdot \cdot *)\circ F = \widehat{u}$ \item $\det F =
316: e^{i\beta_{\widehat{u}}}$.
317: \end{itemize}
318: \end{definition}
319: It is easily seen that the first condition is equivalent to the fact that there
320: exists a smooth map $G:\Omega \longrightarrow U(3)$ (a gauge transformation) of
321: the type
322: \[
323: G(x,y) = \left(\begin{array}{cc}g(x,y) & 0 \\ 0 & 1\end{array}\right),\quad
324: \hbox{where }g:\Omega\longrightarrow U(2)
325: \]
326: such that \[F(x,y) = \widehat{F}(x,y)\cdot G^{-1}(x,y).
327: \]
328: And then the second one is equivalent to say that $g$ takes values in $SU(2)$.
329: 
330: \subsection{A splitting of the Maurer--Cartan form of a Legendrian framing}
331: Using (\ref{duchapeau}) and (\ref{A33=0}) one obtains the following
332: decomposition of $\widehat{A}$:
333: \[
334: \widehat{A} = \widehat{A}_{\u(1)} + \widehat{A}_{\su(2)} +
335: \widehat{A}_{\mathbb{C}^2},
336: \]
337: with the notations
338: \[
339: \widehat{A}_{\u(1)} = \left(\begin{array}{cc}\widehat{\alpha}_{\u(1)} & 0\\0 &
340: 0\end{array}\right),\quad \widehat{A}_{\su(2)} =
341: \left(\begin{array}{cc}\widehat{\alpha}_{\su(2)} & 0\\0 & 0\end{array}\right),
342: \]
343: \[
344: \hbox{and}\quad \widehat{A}_{\mathbb{C}^2} = e^\rho\left(\begin{array}{cc} 0 &
345: \epsilon dz + \overline{\epsilon}d\bar{z}\\
346: -\,^t\!\left(\epsilon dz + \overline{\epsilon}d\bar{z}\right) &
347: 0\end{array}\right),
348: \]
349: where $\widehat{\alpha}_{\u(1)}$ is a 1-form on $\Omega$ with coefficients in
350: $\u(1)\simeq \mathbb{R}\left(\begin{array}{cc}i&0\\0&i\end{array}\right)$,
351: $\widehat{\alpha}_{\su(2)}$ a 1-form on $\Omega$ with coefficients in $\su(2)$,
352: $\epsilon:={1\over 2}\left(\begin{array}{c}1\\-i\end{array}\right)$ and
353: $\overline{\epsilon}:={1\over 2}\left(\begin{array}{c}1\\i\end{array}\right)$
354: (so that $\epsilon dz +
355: \overline{\epsilon}d\bar{z}=\left(\begin{array}{c}dx\\dy\end{array}\right)$).
356: Note that $\det F = e^{i\beta_{\widehat{u}}}$ implies $\widehat{\alpha}_{\u(1)}
357: = {d\beta_{\widehat{u}}\over
358: 2}\left(\begin{array}{cc}i&0\\0&i\end{array}\right)$.\\
359: \noindent Now we let $F:\Omega\longrightarrow U(3)$ be a Legendrian framing and
360: $A:= F^{-1}\cdot dF$.  The relation $F = \widehat{F}\cdot G^{-1}$ implies that
361: $A = G\cdot \widehat{A}\cdot G^{-1} - dG\cdot G^{-1}$.  Hence
362: \[
363: A = A_{\u(1)} + A_{\su(2)} + A_{\mathbb{C}^2},
364: \]
365: where, using the fact that $\u(1)$ commutes with $\su(2)$,
366: \[
367: A_{\u(1)} = \left(\begin{array}{cc}\alpha_{\u(1)} & 0\\0 & 0\end{array}\right) =
368: \left(\begin{array}{cc}\widehat{\alpha}_{\u(1)} & 0\\0 & 0\end{array}\right) =
369: \left(\begin{array}{ccc}i{d\beta_{\widehat{u}}\over
370: 2}&0&0\\0&i{d\beta_{\widehat{u}}\over 2}&0\\
371: 0&0&0\end{array}\right),
372: \]
373: \[
374: A_{\su(2)} = \left(\begin{array}{cc}\alpha_{\su(2)} & 0\\0 & 0\end{array}\right)
375: = \left(\begin{array}{cc}g\cdot \widehat{\alpha}_{\su(2)}\cdot g^{-1} -dg\cdot
376: g^{-1} & 0\\0 & 0\end{array}\right)
377: \]
378: and
379: \[
380: A_{\mathbb{C}^2} = e^\rho\left(\begin{array}{cc} 0 & g\cdot\left( \epsilon dz +
381: \overline{\epsilon}d\bar{z}\right)\\
382: -\left(g\cdot \left(\epsilon dz +
383: \overline{\epsilon}d\bar{z}\right)\right)^\dagger & 0\end{array}\right).
384: \]
385: We can further split the last term $A_{\mathbb{C}^2}$ along $dz$ and $d\bar{z}$
386: as $A_{\mathbb{C}^2} = A_{\mathbb{C}^2}' + A_{\mathbb{C}^2}''$ where
387: \[
388: A_{\mathbb{C}^2}':= e^\rho\left(\begin{array}{cc} 0 & g\cdot \epsilon \\
389: -\left(g\cdot \overline{\epsilon}\right)^\dagger & 0\end{array}\right) dz
390: \quad\hbox{and}\quad A_{\mathbb{C}^2}'':= e^\rho\left(\begin{array}{cc} 0 &
391: g\cdot \overline{\epsilon} \\
392: -\left(g\cdot \epsilon\right)^\dagger & 0\end{array}\right) d\bar{z}.
393: \]
394: 
395: \subsection{Interpretation in terms of an automorphism}
396: As expounded in \cite{HR1} and \cite{HR3} the key point in order to exploit the
397: structure of an integrable system is to observe that the splitting $A =
398: A_{\u(1)} + A_{\su(2)} + A_{\mathbb{C}^2}' + A_{\mathbb{C}^2}''$ corresponds to
399: a decomposition along the eigenspaces of the following automorphism in
400: $\u(3)^{\mathbb{C}}$, the complexification\footnote{we can define
401: $\u(3)^{\mathbb{C}}$ as the set $M(3,\mathbb{C})$ with its standard complex
402: structure and with the conjugation mapping $c:M\longmapsto -M^\dagger$; clearly
403: $c$ is a Lie algebra automorphism, an involution and the set of fixed points of
404: $c$ is $\u(3)$.\\
405: Similarly the complexification $U(3)^{\mathbb{C}}$ is the set $GL(3,\mathbb{C})$
406: with its standard complex structure and the conjugation map $C:G\longmapsto
407: \left(G^\dagger\right)^{-1}$.} of $\u(3)$.  We let $J:=
408: \left(\begin{array}{cc}0&-1\\1&0\end{array}\right)$ and
409: \[
410: \begin{array}{cccl}
411: \tau: & \u(3)^{\mathbb{C}} & \longrightarrow & \u(3)^{\mathbb{C}}\\
412:  & M & \longmapsto & - \left( \begin{array}{cc}-J & 0\\0 & 1\end{array}\right)
413:  \cdot \,^t\!M\cdot \left( \begin{array}{cc}J & 0\\0 & 1\end{array}\right).
414: \end{array}
415: \]
416: It is then straightforward that $\tau$ is a Lie algebra automorphism, that
417: $\u(3)$ is stable by $\tau$ and that $\tau^4 = Id$.  Hence we can diagonalize
418: the action of $\tau$ over $\u(3)^{\mathbb{C}}$ and in the following we denote by
419: $\u(3)^{\mathbb{C}}_a$ the eigenspace of $\tau$ for the eigenvalue $i^a$, for $a
420: = -1,0,1,2$.  We first point out that the eigenspaces $\u(3)^{\mathbb{C}}_0$ and
421: $\u(3)^{\mathbb{C}}_2$, with eigenvalues 1 and $-1$ respectively, are the
422: complexifications of $\u(3)_0$ and $\u(3)_2$ respectively, where
423: \[
424: \u(3)_0:= \left\{ \left(\begin{array}{cc}g & 0 \\0 & 0\end{array}\right) /g\in
425: \su(2)\right\} \hbox{ and } \u(3)_2:= \left\{ \left(\begin{array}{ccc}\lambda
426: i&0&0\\0&\lambda i&0\\0&0&\mu i\end{array}\right) / \lambda, \mu\in
427: \mathbb{R}\right\}.
428: \]
429: This can be obtained by first computing that
430: \[
431: \tau \left(\begin{array}{cc}A & X \\-\,^t\!Y & d\end{array}\right) =
432: \left(\begin{array}{cc}J\,^t\!AJ & -JY \\-\,^t\!XJ & -d\end{array}\right), \quad
433: \forall A\in M(2,\mathbb{C}),\forall X,Y\in \mathbb{C}^2,\forall d\in
434: \mathbb{C},
435: \]
436: and by using the fact that $\forall A\in \sl(2,\mathbb{C})$, $J\,^t\!AJ = A$.
437: Similarly the eigenspaces $\u(3)^{\mathbb{C}}_{1}$ and
438: $\u(3)^{\mathbb{C}}_{-1}$, with eigenvalues $i$ and $-i$ respectively, are found
439: to be
440: \[
441: \u(3)^{\mathbb{C}}_{1} = \left\{ \left(\begin{array}{cc}0 & X \\-\,^t\!Y &
442: 0\end{array}\right) /X,Y\in \mathbb{C}^2, JY = -iX\right\}
443: \]
444: and
445: \[
446: \u(3)^{\mathbb{C}}_{-1} = \left\{ \left(\begin{array}{cc}0 & X \\-\,^t\!Y &
447: 0\end{array}\right) /X,Y\in \mathbb{C}^2, JY = iX\right\}.
448: \]
449: Now we have the following
450: \begin{lemma}
451: The eigenspaces $\u(3)^{\mathbb{C}}_{1}$ and $\u(3)^{\mathbb{C}}_{-1}$ can be
452: characterized by
453: \[
454: \u(3)^{\mathbb{C}}_{1} = \left\{\lambda \left(\begin{array}{cc}0 & h\cdot
455: \overline{\epsilon} \\
456: -\left(h\cdot \epsilon\right)^\dagger & 0\end{array}\right) / \lambda\in
457: [0,\infty), h\in SU(2)\right\}
458: \]
459: and
460: \[
461: \u(3)^{\mathbb{C}}_{-1} = \left\{ \lambda\left(\begin{array}{cc}0 & h\cdot
462: \epsilon \\
463: -\left(h\cdot \overline{\epsilon}\right)^\dagger & 0\end{array}\right) /
464: \lambda\in [0,\infty), h\in SU(2)\right\}.
465: \]
466: \end{lemma}
467: \begin{proof} This can be proved either by adapting the argument in section 2.4
468: of \cite{HR1} or by a straightforward computation which exploits the fact that
469: $\forall h\in SU(2)$, $\overline{h}J = Jh$, $J\epsilon = i\epsilon$ and
470: $J\overline{\epsilon} = -i \overline{\epsilon}$.  \end{proof}
471: 
472: \noindent We conclude that if, using $\u(3)^{\mathbb{C}} =
473: \u(3)^{\mathbb{C}}_{-1}\oplus \u(3)^{\mathbb{C}}_0\oplus
474: \u(3)^{\mathbb{C}}_1\oplus \u(3)^{\mathbb{C}}_2$, we decompose $A$ as
475: \[
476: A = A_{-1} + A_0 + A_1 + A_2,
477: \]
478: where each $A_a$ is a 1-form with coefficients in the $\u(3)^{\mathbb{C}}_a$,
479: then we recover the previous splitting by setting $A_0 = A_{\su(2)}$, $A_2 =
480: A_{\u(1)}$, $A_{-1} = A_{\mathbb{C}^2}'$ and $A_1 = A_{\mathbb{C}^2}''$.  Note
481: that the two last conditions actually reflects the conformality of $u$.
482: 
483: \begin{remark}
484: Note that by the automorphism property $[\u(3)_a,\u(3)_b]\subset \u(3)_{a+b\bmod
485: 4}$.
486: \end{remark}
487: 
488: \subsection{Legendrian framings of Hamiltonian stationary Lagrangian immersions}
489: \noindent Given the Legendrian framing $F$ of a conformal Lagrangian immersion
490: $u$ in $\mathbb{C}P^2$, we define the family of deformations $A_\lambda$ of its
491: Maurer--Cartan form $A$, for $\lambda\in S^1\subset \mathbb{C}^*$ by
492: \begin{equation}\label{1.6.Alambda}
493: A_\lambda:= \lambda^{-2}A_2' + \lambda^{-1}A_{-1} + A_0 + \lambda A_1 +
494: \lambda^2A_2'',
495: \end{equation}
496: where $A_2':= A_2(\partial /\partial z)dz$ and $A_2'':= A_2(\partial /\partial
497: \bar{z})d\bar{z}$.  We then have the following:
498: 
499: \begin{theorem}\label{1.6.theo}
500: Given a conformal Lagrangian immersion $u:\Omega\longrightarrow \mathbb{C}P^2$
501: and a Legendrian framing $F$ of $u$, the Maurer--Cartan form of $F$ satisfies
502: \begin{equation}\label{1.6.lag}
503: A_{-1} = A_{-1}'= A_{\mathbb{C}^2}'\hbox{ and }A_1 = A_1''= A_{\mathbb{C}^2}'.
504: \end{equation}
505: Furthermore $u$ is Hamiltonian stationary if and only if, defining $A_\lambda$
506: as in (\ref{1.6.Alambda}),
507: \begin{equation}\label{1.6.curvature}
508: dA_\lambda + A_\lambda\wedge A_\lambda = 0,\quad \forall \lambda\in S^1.
509: \end{equation}
510: \end{theorem}
511: %
512: \begin{remark}
513: For $\lambda = 1$, $A_1 = A$ and equation (\ref{1.6.curvature}) is a consequence
514: of its definition $A:=F^{-1}\cdot dF$.
515: \end{remark}
516: %
517: \begin{proof} See \cite{HR1} and \cite{HR3}.\end{proof}
518: \noindent We remark that all the conditions that have been collected about the
519: components $A_a$ can be encoded by the following twisting condition on
520: $A_\lambda$:
521: \[
522: \forall \lambda\in S^1,\quad \tau(A_\lambda) = A_{i\lambda}.
523: \]
524: Thus we are led to define the following {\em twisted loop algebra}
525: \[
526: \Lambda \u(3)_\tau := \{S^1\ni\lambda \longmapsto \xi_\lambda\in \u(3)/\forall
527: \lambda\in S^1, \tau(\xi_\lambda) = \xi_{i\lambda}\},
528: \]
529: and $A_\lambda$ is a 1-form on $\Omega$ with coefficients in $\Lambda
530: \u(3)_\tau$.\\
531: 
532: \noindent Actually $\Lambda \u(3)_\tau$ is the Lie algebra of the following {\em
533: twisted loop group}
534: \[
535: \Lambda U(3)_\tau := \{S^1\ni\lambda \longmapsto g_\lambda\in U(3)/\forall
536: \lambda\in S^1, \tau(g_\lambda) = g_{i\lambda}\},
537: \]
538: where the Lie algebra automorphism $\tau :\u(3)^{\mathbb{C}}\longrightarrow
539: \u(3)^{\mathbb{C}}$ has been extended to the Lie group automorphism by
540: \[
541: \begin{array}{cccc}
542: \tau: & U(3)^{\mathbb{C}} & \longrightarrow & U(3)^{\mathbb{C}}\\
543:  & M & \longmapsto & \left( \begin{array}{cc}-J & 0\\0 & 1\end{array}\right)
544:  \cdot \,^t\!M^{-1}\cdot \left( \begin{array}{cc}J & 0\\0 & 1\end{array}\right).
545: \end{array}
546: \]
547: Now if we assume that $\Omega$ is simply connected, then relation
548: (\ref{1.6.curvature}) allows us to integrate $A_\lambda$, i.e.\,to find a map
549: $F_\lambda:\Omega\longrightarrow U(3)$ for any $\lambda\in S^1$ such that
550: $dF_\lambda = F_\lambda\cdot A_\lambda$.  Moreover if we choose some base point
551: $z_0\in \Omega$, then by requiring further that $F_\lambda(z_0) = Id$,
552: $F_\lambda$ is unique.  A key observation is then that $\tau(A_\lambda) =
553: A_{i\lambda}$ implies $\tau(F_\lambda) = F_{i\lambda}$, $\forall \lambda\in
554: S^1$.  Hence, a conformal Lagrangian immersion $u:\Omega\longrightarrow
555: \mathbb{C}P^2$ is Hamiltonian stationary if and only if any Legendrian lift $F$
556: of it can be deformed into a map $F_\lambda:\Omega\longrightarrow \Lambda
557: U(3)_\tau$, such that $F_\lambda^{-1}\cdot dF_\lambda$ has the form
558: (\ref{1.6.Alambda}).  Summarizing this result with the observations in the
559: previous section we have:
560: 
561: \begin{theorem}\label{1.6.theobis}
562: Given a simply connected domain $\Omega\subset \mathbb{C}$ and a base point
563: $z_0\in \Omega$, the set of Hamiltonian stationary conformal Lagrangian
564: immersions $u:\Omega\longrightarrow \mathbb{C}P^2$ such that $u(z_0) = [0:0:1]$
565: is in bijection with the set of maps $F_\lambda:\Omega\longrightarrow \Lambda
566: U(3)_\tau$, such that $F_\lambda(z_0)=\mathit{Id}$ and the Fourier decomposition
567: of $A_\lambda:=F_\lambda^{-1}\cdot dF_\lambda$, $A_\lambda = \sum_{k\in
568: \mathbb{Z}}\widehat{A}_k\lambda^k$ satisfies
569: \begin{equation}\label{1.6.-infty}
570: \forall k\in \mathbb{Z},\quad k\leq -3 \Longrightarrow \widehat{A}_k = 0,
571: \end{equation}
572: \begin{equation}\label{1.6.-2}
573: \widehat{A}_{-2}=
574: a(z)dz\left(\begin{array}{ccc}i&0&0\\0&i&0\\0&0&0\end{array}\right), \quad
575: \hbox{where }a\in{\cal C}^\infty(\Omega,\mathbb{C}),
576: \end{equation}
577: \begin{equation}\label{1.6.-1}
578: \widehat{A}_{-1}=\widehat{A}_{-1}(\partial /\partial z)dz,\quad \hbox{i.e. }
579: \widehat{A}_{-1}(\partial /\partial \bar{z}) = 0.
580: \end{equation}
581: \end{theorem}
582: \begin{proof} For any conformal Lagrangian Hamiltonian stationary immersion $u$
583: the existence of $F_\lambda$ and the properties (\ref{1.6.-infty}),
584: (\ref{1.6.-2}) and (\ref{1.6.-1}) are immediate consequences of Theorem
585: \ref{1.6.theo}.  Conversely for any map $F_\lambda$, conditions
586: (\ref{1.6.-infty}), (\ref{1.6.-2}) and (\ref{1.6.-1}) and the reality condition
587: $\overline{A_\lambda} = A_\lambda$ imply that $A_\lambda$ must satisfy
588: (\ref{1.6.Alambda}).  In particular we remark that condition (\ref{1.6.-2}) is a
589: reformulation of (\ref{A33=0}).  Thus by theorem \ref{1.6.theo} we deduce that
590: $F_1$ is the Legendrian lift of some Hamiltonian stationary conformal Lagrangian
591: immersion.\end{proof}
592: %
593: \begin{remark}
594: >From the analysis of the Maurer--Cartan form of a Legendrian lift we know that
595: actually the function $a$ in (\ref{1.6.-2}) is ${1\over 2}\partial
596: \beta/\partial z$, where $\beta$ is the Lagrangian angle function.  In
597: particular since $u$ is Hamiltonian stationary $\beta$ is harmonic and hence $a$
598: is holomorphic.
599: \end{remark}
600: 
601: \subsection{An alternative characterization}
602: We introduce here another construction using based loop groups for
603: characterizing Hamiltonian stationary Lagrangian conformal immersions.  Consider
604: \[
605: E_\lambda:= F_\lambda\cdot F^{-1}.
606: \]
607: We can observe that $E_\lambda$ is a map with values in the {\em based loop
608: group}
609: \[
610: \Omega U(3):= \{S^1\ni\lambda\longmapsto g_\lambda\in U(3)/g_{\lambda=1}=1\},
611: \]
612: since $F_{\lambda=1}=F$.  It is easy to check that $\Omega U(3)$ is a loop
613: group, the Lie algebra of which is
614: \[
615: \Omega \u(3):= \{S^1\ni\lambda\longmapsto \xi_\lambda\in
616: \u(3)/\xi_{\lambda=1}=0\}.
617: \]
618: Note that the (formal) Fourier expansion of an element $\xi_\lambda\in \Omega
619: \u(3)$ can be written $\xi_\lambda = \sum_{k\in\mathbb{Z}\setminus
620: \{0\}}\widehat{\xi}_k(\lambda^k-1)$.\\
621: 
622: \noindent The Maurer--Cartan form of $E_\lambda$ is
623: \[
624: \begin{array}{ccl}
625: \Gamma_\lambda & := & E_\lambda^{-1}\cdot dE_\lambda \\
626: & = & F\cdot\left( F_\lambda^{-1}\cdot dF_\lambda - F\cdot dF\right) \cdot
627: F^{-1} = F\cdot\left(A_\lambda - A\right) \cdot F^{-1}\\
628: & = & (\lambda^{-2}-1)\Gamma_2' + (\lambda^{-1}-1)\Gamma_{-1} +
629: (\lambda-1)\Gamma_1 + (\lambda^{2}-1)\Gamma_2'',
630:  \end{array}
631: \]
632: where $\Gamma_2':= F\cdot A_2'\cdot F^{-1}$, $\Gamma_{-1}:= F\cdot A_{-1}\cdot
633: F^{-1}$, $\Gamma_1:= F\cdot A_1\cdot F^{-1}$ and $\Gamma_2'':= F\cdot A_2''\cdot
634: F^{-1}$.  We can observe in particular that
635: \[
636: \Gamma_2' = ia \pi^\perp dz,\quad \hbox{where }\pi^\perp:=
637: F\cdot\left(\begin{array}{ccc}1&&\\&1&\\&&0\end{array}\right)\cdot F^{-1}.
638: \]
639: Note that $\pi^\perp$ is the Hermitian orthogonal projection in $\C^3$ onto the
640: plane $\widehat{u}^\perp$ (moreover $\pi^\perp$ is actually independent of the
641: lift $\widehat{u}$ chosen for $u$).\\
642: 
643: \noindent Lastly we point out the following equivariance property with respect
644: to the automorphism $\tau_u$ defined\footnote{ We can remark that the definition
645: of $\tau_u$ is independent from the choice of the Legendrian framing $F$ of $u$,
646: and depends only on $u$.  This means that for any pair of Legendrian framings
647: $F$ and $\widehat{F}$ such that $ \widehat{F} = F\cdot G$, where
648: $G=\left(\begin{array}{cc}g&\\&1\end{array}\right)$ and $g:\Omega\longrightarrow
649: SU(2)$, we have $\widehat{F}\cdot \tau(\widehat{F}^{-1}\cdot M\cdot
650: \widehat{F})\cdot \widehat{F}^{-1} = F\cdot \tau (F^{-1}\cdot M\cdot F)\cdot
651: F^{-1}$.  This can be checked by a computation using the fact that
652: $\left(\begin{array}{cc}\pm J&\\&1\end{array}\right)\cdot G =\overline{G}\cdot
653: \left(\begin{array}{cc}\pm J&\\&1\end{array}\right)$.} by
654: \[
655: \tau_u(M) = F\cdot \tau (F^{-1}\cdot M\cdot F)\cdot F^{-1}.
656: \]
657: We have obviously $\tau_u^4=1$.  Moreover, setting
658: \begin{eqnarray*}
659: \gamma_\lambda & := & \lambda^{-2}\Gamma_2' + \lambda^{-1}\Gamma_{-1} +
660: \lambda\Gamma_1+ \lambda^{2}\Gamma_2'' \\
661: & = & F\cdot\left( \lambda^{-2}A_2'+\lambda^{-1}A_{-1} + \lambda A_1+
662: \lambda^{2}A_2''\right)\cdot F^{-1}
663: \end{eqnarray*}
664: and $\gamma:= \gamma_{\lambda=1} = F\cdot\left( A_2'+A_{-1} + A_1+
665: A_2''\right)\cdot F^{-1}$, so that $\Gamma_\lambda = \gamma_\lambda -\gamma$, we
666: have
667: \[
668: \tau_u(\gamma_\lambda) = \gamma_{i\lambda}.
669: \]
670: 
671: 
672: \section{Finite type solutions}
673: \noindent In \cite{HR3} we showed how Theorem \ref{1.6.theobis} allows us to
674: adapt the theory of J. Dorfmeister, F. Pedit and H.Y. Wu \cite{DPW}, in order to
675: build a Weierstrass type representation theory of {\em all} conformal Lagrangian
676: Hamiltonian stationary immersions, i.e.\,using holomorphic data.  Here we want
677: to exploit Theorem \ref{1.6.theobis} in order to construct a particular class of
678: examples of solutions: the {\em finite type} ones.
679: 
680: \subsection{Definitions}
681: \noindent We invite the Reader to consult \cite{BFPP}, \cite{G} or \cite{H} for
682: more details.  We first observe that $U(3)_0:=\{g\in U(3)/ \tau(g)=g\}$, the
683: fixed set of $\tau$, is a subgroup of $U(3)$, the Lie algebra of which is
684: $\u(3)_0$ (same observation about $U(3)_0^{\C}$).  Actually $U(3)_0$ is
685: isomorphic to $SU(2)$ so that we make the identifications $U(3)_0\simeq SU(2)$
686: and $\u(3)_0\simeq \su(2)$.  We will need an Iwasawa decomposition of
687: $SU(2)^{\mathbb{C}}$ for our purpose: it will be a pair $(SU(2),\mathfrak{B})$
688: of {\em subgroups} of $SU(2)^{\mathbb{C}}$, such that $\forall g\in
689: SU(2)^{\mathbb{C}}$, $\exists !  (f,b)\in SU(2)\times \mathfrak{B}$ with
690: $g=f\cdot b$, a property that we summarize by writing $SU(2)^{\mathbb{C}} =
691: SU(2)\cdot \mathfrak{B}$.  Moreover $\mathfrak{B}$ is a solvable Borel subgroup.
692: We can choose for example
693: \[
694: \mathfrak{B}:=\left\{\left(\begin{array}{cc}T^1_1 &
695: 0\\T^2_1&T^2_2\end{array}\right)/ T^1_1,T^2_2\in (0,\infty), T^2_1\in
696: \mathbb{C}, T^1_1T^2_2 = 1\right\}.
697: \]
698: We denote by $\mathfrak{b}$ the Lie algebra of $\mathfrak{B}$.  The Iwasawa
699: decomposition $SU(2)^{\mathbb{C}} = SU(2)\cdot \mathfrak{B}$ immediately implies
700: the vector space decomposition $\su(2)^{\mathbb{C}} = \su(2)\oplus
701: \mathfrak{b}$, which leads to the definition of the two projection mappings
702: $(\cdot )_{\su}:\su(2)^{\mathbb{C}} \longrightarrow \su(2)$ and
703: $(\cdot)_\mathfrak{b}:\su(2)^{\mathbb{C}} \longrightarrow \mathfrak{b}$ such
704: that
705: \[
706: \forall \xi \in \su(2)^{\mathbb{C}}, \quad \xi =
707: (\xi)_{\su}+(\xi)_\mathfrak{b}\quad \hbox{with }(\xi)_{\su}\in \su(2)\hbox{ and
708: }(\xi)_\mathfrak{b} \in \mathfrak{b}.
709: \]
710: 
711: \noindent Then we define the following twisted loop algebras
712: \[
713: \Lambda \u(3)_\tau^{\mathbb{C}} := \{S^1\ni\lambda \longmapsto \xi_\lambda\in
714: \u(3)^{\mathbb{C}}/ \forall \lambda\in S^1, \tau(\xi_\lambda) =
715: \xi_{i\lambda}\},
716: \]
717: \[
718: \Lambda_{\mathfrak{b}}^+ \u(3)_\tau^{\mathbb{C}} := \{[\lambda \longmapsto
719: \xi_\lambda] \in \Lambda \u(3)_\tau^{\mathbb{C}}/\forall k\in \mathbb{Z}, k\leq
720: -1\Longrightarrow \widehat{\xi}_k = 0\hbox{ and }\widehat{\xi}_0\in
721: \mathfrak{b}\},
722: \]
723: where we use the Fourier decomposition $\xi_\lambda =\sum_{k\in
724: \mathbb{Z}}\widehat{\xi}_k\lambda^k$.\\
725: 
726: \noindent The decomposition $\su(2)^{\mathbb{C}} = \su(2)\oplus \mathfrak{b}$
727: can be extended to loop algebras, i.e. to the splitting $\Lambda
728: \u(3)_\tau^{\mathbb{C}} = \Lambda \u(3)_\tau\oplus \Lambda_{\mathfrak{b}}^+
729: \u(3)_\tau^{\mathbb{C}}$.  This can be checked by using the Fourier expansion of
730: an element $\xi_\lambda\in \Lambda \u(3)_\tau^{\mathbb{C}}$:
731: \[
732: \sum_{k\in \mathbb{Z}}\widehat{\xi}_k\lambda^k = \left(
733: \sum_{k<0}\widehat{\xi}_k\lambda^k + (\widehat{\xi}_0)_{\su} -
734: \sum_{k>0}\left(\widehat{\xi}_{-k}\right)^\dagger\lambda^k \right) + \left(
735: (\widehat{\xi}_0)_\mathfrak{b} + \sum_{k>0}\left(\widehat{\xi}_k
736: +\left(\widehat{\xi}_{-k}\right)^\dagger \right)\lambda^k \right).
737: \]
738: We will denote the corresponding projection mappings by $(\cdot
739: )_{\Lambda_{\su}}:\Lambda \u(3)_\tau^{\mathbb{C}}\longrightarrow \Lambda
740: \u(3)_\tau$ and $(\cdot)_{\Lambda_\mathfrak{b}^+}:\Lambda
741: \u(3)_\tau^{\mathbb{C}}\longrightarrow \Lambda_{\mathfrak{b}}^+
742: \u(3)_\tau^{\mathbb{C}}$.\\
743: 
744: \noindent We also introduce the following finite dimensional subspaces of
745: $\Lambda \u(3)_\tau$: for any $p\in \mathbb{N}$ we let
746: \[
747: \Lambda^{2+4p} \u(3)_\tau:= \left\{[\lambda \longmapsto \xi_\lambda]\in \Lambda
748: \u(3)_\tau/ \xi_\lambda =\sum_{k=-2-4p}^{2+4p}\widehat{\xi}_k\lambda^k\right\}.
749: \]
750: We can now define a pair of vector fields $X_1,X_2:\Lambda^{2+4p}
751: \u(3)_\tau\longrightarrow \Lambda \u(3)_\tau$ by
752: \begin{equation}\label{2.1.X}
753: X_1(\xi_\lambda):= [\xi_\lambda,(\lambda^{4p}\xi_\lambda)_{\Lambda_{\su}}],
754: \quad X_2(\xi_\lambda):=
755: [\xi_\lambda,(i\lambda^{4p}\xi_\lambda)_{\Lambda_{\su}}].
756: \end{equation}
757: Note that $\lambda^{4p}\xi_\lambda$ belongs to $\Lambda \u(3)_\tau^{\C}$, so
758: that $(\lambda^{4p}\xi_\lambda)_{\Lambda_{\su}}$ is well defined.
759: \begin{lemma}
760: Let $p\in \mathbb{N}$ and $X_1$ and $X_2$ defined by (\ref{2.1.X}).  Then
761: \begin{itemize}
762: \item $\forall \xi_\lambda\in \Lambda^{2+4p} \u(3)_\tau$, $X_1(\xi_\lambda),
763: X_2(\xi_\lambda)\in T_{\xi_\lambda}\Lambda^{2+4p} \u(3)_\tau \simeq
764: \Lambda^{2+4p} \u(3)_\tau$, so that $X_1$ and $X_2$ are tangent vector fields to
765: $\Lambda^{2+4p} \u(3)_\tau$.  \item $|\xi_\lambda|^2$ is preserved by $X_1$ and
766: $X_2$.  Hence the flow of these vector fields are defined for all time \item The
767: Lie bracket of $X_1$ and $X_2$ vanishes:
768: \begin{equation}\label{2.1.[,]}
769: [X_1,X_2] = 0.
770: \end{equation}
771: \end{itemize}
772: \end{lemma}
773: \begin{proof} This result follows by a straightforward adaptation of the
774: analogous results for harmonic maps in \cite{BFPP} (see e.g.\,\cite{G} and
775: \cite{H}).  Note that the proof of (\ref{2.1.[,]}) rests upon the crucial
776: property that $\Lambda \u(3)_\tau$ and $\Lambda_{\mathfrak{b}}^+
777: \u(3)_\tau^{\mathbb{C}}$ are Lie algebras (see e.g.\,\cite{BP},
778: \cite{H}).\end{proof}
779: 
780: \noindent This result allows us to integrate simultaneously $X_1$ and $X_2$.  So
781: for any $\xi_\lambda^0\in \Lambda^{2+4p} \u(3)_\tau$ there exists a unique map
782: $\xi_\lambda:\mathbb{R}^2\longrightarrow \Lambda^{2+4p} \u(3)_\tau$ such that
783: $\xi_\lambda(z_0) = \xi_\lambda^0$ and
784: \begin{equation}\label{2.1.flot}
785: {\partial \xi_\lambda\over \partial x}(x,y) = X_1\left(\xi_\lambda(x,y)\right)
786: \quad \hbox{and} \quad {\partial \xi_\lambda\over \partial y}(x,y) =
787: X_2\left(\xi_\lambda(x,y)\right).
788: \end{equation}
789: Denoting by $z=x+iy\in \C$, the system (\ref{2.1.flot}) can be rewritten
790: \[
791: \begin{array}{ccl}
792: d\xi_\lambda & = & \left[\xi_\lambda,
793: \left(\lambda^{4p}\xi_\lambda\right)_{\Lambda_{\su}}dx +
794: \left(i\lambda^{4p}\xi_\lambda\right)_{\Lambda_{\su}}dy\right]\\
795: & = & \left[\xi_\lambda, \left(\lambda^{4p}\xi_\lambda
796: dz\right)_{\Lambda_{\su}}\right].
797: \end{array}
798: \]
799: Let us denote by $A_\lambda:= (\lambda^{4p}\xi_\lambda dz)_{\Lambda_{\su}}$.
800: Since the system (\ref{2.1.flot}) is overdetermined, $A_\lambda$ should satisfy
801: a compatibility condition.  Indeed one can check that
802: \begin{equation}\label{2.1.curv}
803: dA_\lambda + A_\lambda\wedge A_\lambda = 0.
804: \end{equation}
805: This relation can be proved by a method similar to the proof of (\ref{2.1.[,]})
806: (see \cite{H}).  It implies that there exists a map $F_\lambda:\C\longrightarrow
807: \Lambda U(3)_\tau$ such that
808: \begin{equation}\label{2.1.dF}
809: dF_\lambda = F_\lambda\cdot A_\lambda.
810: \end{equation}
811: Now observe that $\lambda^{4p}\xi_\lambda =
812: \sum_{k=-2}^{8p+2}\hat{\xi}_{k-4p}\lambda^k$ implies
813: \[
814: A_\lambda = \lambda^{-2}\hat{\xi}_{-4p-2}dz + \lambda^{-1}\hat{\xi}_{-4p-1}dz +
815: \left(\hat{\xi}_{-4p}dz\right)_{\su} -
816: \lambda\left(\hat{\xi}_{-4p-1}\right)^\dagger d\bar{z} -
817: \lambda^{2}\left(\hat{\xi}_{-4p-2}\right)^\dagger d\bar{z}.
818: \]
819: We recall that $\hat{\xi}_{-4p-2}\in \u(3)_2^{\C}$ and so has the form
820: $\mathrm{diag}(ia,ia,ib)$.  Moreover we have the following result.
821: \begin{lemma}
822: If $\xi_\lambda\longrightarrow \Lambda^{2+4p} \u(3)_\tau$ and $A_\lambda:=
823: (\lambda^{4p}\xi_\lambda dz)_{\Lambda_{\su}}$ are solutions of $d\xi_\lambda =
824: [\xi_\lambda,A_\lambda]$, then $\hat{\xi}_{-4p-2}$ is constant.
825: \end{lemma}
826: \begin{proof} The relevant term in the Fourier expansion of $d\xi_\lambda =
827: [\xi_\lambda,A_\lambda]$ gives
828: \[
829: \begin{array}{ccl}
830: \displaystyle d\hat{\xi}_{-4p-2} & = & \displaystyle \left[ \hat{\xi}_{-4p-2},
831: \left(\hat{\xi}_{-4p}dz\right)_{\su}\right] + \left[ \hat{\xi}_{-4p-1},
832: \hat{\xi}_{-4p-1}\right]dz + \left[ \hat{\xi}_{-4p},
833: \hat{\xi}_{-4p-2}\right]dz\\
834:  & = & \displaystyle
835:  \left[\left(\hat{\xi}_{-4p}dz\right)_\mathfrak{b},\hat{\xi}_{-4p-2}\right].
836: \end{array}
837: \]
838: But since the coefficients of $\left(\hat{\xi}_{-4p}dz\right)_\mathfrak{b}$ are
839: in $\u(3)_0^{\C}$ and $\hat{\xi}_{-4p-2}$ takes values in $\u(3)_2$ we deduce
840: that $d\hat{\xi}_{-4p-2} = 0$, because $\u(3)_0^{\C}$ and $\u(3)_2^{\C}$
841: commute.  \end{proof}
842: 
843: We deduce from this result that if we choose the initial value $\xi_\lambda^0$
844: of $\xi_\lambda$ to be such that $\hat{\xi}^0_{-4p-2} = \mathrm{diag}(ia,ia,0)$
845: then $\hat{\xi}_{-4p-2}$ is equal to that value for all $(x,y)$.  So in this
846: case the map $F_\lambda$ obtained by integrating $A_\lambda$ satisfies all the
847: requirements of Theorem \ref{1.6.theobis}.  It implies that $F_\lambda$
848: represents a (conjugate family) of Hamiltonian stationary conformal Lagrangian
849: immersion(s).  The category of such $F_\lambda$'s are exactly characterized by
850: the following definition.
851: 
852: \begin{definition}\label{2.1.def-ft}
853: Let $F_\lambda$ be a family of Hamiltonian stationary conformal Lagrangian
854: immersions and let $A_\lambda:= F_\lambda^{-1}\cdot dF_\lambda$.  Then
855: $F_\lambda$ is called a family of {\em finite type solutions} if and only if
856: there exists $p\in \mathbb{N}$ and a map $\xi_\lambda:\mathbb{C}\longrightarrow
857: \Lambda^{2+4p} \u(3)_\tau$ such that $\hat{\xi}_{-4p-2} =
858: \mathrm{diag}(ia,ia,0)$, for some constant $a\in \C$, and
859: \begin{equation}\label{2.1.laxfn}
860: d\xi_\lambda = [\xi_\lambda,A_\lambda]
861: \end{equation}
862: \begin{equation}\label{2.1.laxfn2}
863: (\lambda^{4p}\xi_\lambda dz)_{\Lambda_{\su}} = A_\lambda.
864: \end{equation}
865: \end{definition}
866: We also need the following definition in which we introduce an a priori weaker
867: notion of finite type solution.
868: 
869: \begin{definition}
870: Let $F_\lambda$ be a family of Hamiltonian stationary conformal Lagrangian
871: immersions and let $A_\lambda:= F_\lambda^{-1}\cdot dF_\lambda$.  Then
872: $F_\lambda$ is called a family of {\em quasi-finite type solutions} if and only
873: if it satisfies the same requirements as in definition \ref{2.1.def-ft} excepted
874: that condition (\ref{2.1.laxfn2}) is replaced by
875: \begin{equation}\label{2.1.laxqfn2}
876: \exists B\in \Omega^1\otimes \u(3)^{\C}_0,\quad (\lambda^{4p}\xi_\lambda
877: dz)_{\Lambda_{\su}} = A_\lambda + B.
878: \end{equation}
879: \end{definition}
880: We shall see in Section 3.3 that both definitions are actually equivalent.
881: 
882: \subsection{An alternative description of quasi-finite type solutions}
883: We may as well characterize such finite type solutions in terms of $E_\lambda =
884: F_\lambda\cdot F^{-1}$.  For that purpose we need to introduce the untwisted
885: loop Lie algebra
886: \[
887: \Lambda^+\u(3)^{\C}:= \{S^1\ni\lambda\longmapsto \xi_\lambda\in \u(3)^{\C}/
888: \xi_\lambda = \sum_{k=0}^\infty \widehat{\xi}_k\lambda^k\}
889: \]
890: and observe that any $\xi_\lambda=\sum_{k=-\infty}^\infty
891: \widehat{\xi}_k\lambda^k\in \Lambda \u(3)^{\C}$ can be split as
892: \[
893: \xi_\lambda = \left(\sum_{k=-\infty}^{-1}\widehat{\xi}_k(\lambda^k-1) -
894: (\widehat{\xi}_k)^\dagger (\lambda^{-k}-1)\right) + \left(\sum_{k=0}^\infty
895: \widehat{\xi}_k\lambda^k + \sum_{k=1}^\infty \widehat{\xi}_{-k} +
896: (\widehat{\xi}_{-k})^\dagger (\lambda^{k}-1)\right)
897: \]
898: and hence $\Lambda \u(3)^{\C} = \Omega \u(3)\oplus \Lambda^+\u(3)^{\C}$.  This
899: defines a pair of projection mappings $(\cdot)_{\Omega}:\Lambda \u(3)^{\C}
900: \longrightarrow \Omega \u(3)$ and $(\cdot)_{\Lambda^+}:\Lambda \u(3)^{\C}
901: \longrightarrow \Lambda^+\u(3)^{\C}$.
902: 
903: Now consider a family $F_\lambda$ of quasi-finite type, let $A_\lambda:=
904: F_\lambda^{-1}\cdot dF_\lambda$, $A:= F^{-1}\cdot dF$ (where $F=F_{\lambda=1}$)
905: and $\xi_\lambda$ be a solution of (\ref{2.1.laxfn}).  We let
906: \[
907: \eta_\lambda:= F\cdot \xi_\lambda\cdot F^{-1} = \sum_{k=-2-4p}^{2+4p}F\cdot
908: \widehat{\xi}_k\cdot F^{-1}\lambda^k .
909: \]
910: Then (\ref{2.1.laxfn}) implies by a straightforward computation that
911: \[
912: \begin{array}{ccl}
913: d\eta_\lambda & = & F\cdot (d\xi_\lambda + [A,\xi_\lambda])\cdot F^{-1}\\
914: & = & F\cdot ([\xi_\lambda,A_\lambda] - [\xi_\lambda,A])\cdot F^{-1} =
915: [\eta_\lambda, \Gamma_\lambda],
916: \end{array}
917: \]
918: where $\Gamma_\lambda:= E_\lambda^{-1}\cdot dE_\lambda$.  Now setting
919: $R_\lambda:= \sum_{k=-4p}^{2+4p}F\cdot \widehat{\xi}_k\cdot F^{-1}\lambda^k$, we
920: have
921: \begin{eqnarray*}
922: \displaystyle \left(\lambda^{4p}\eta_\lambda dz\right)_\Omega & = &
923: \displaystyle \left( \lambda^{-2}F\cdot \widehat{\xi}_{-2-4p}\cdot F^{-1}dz +
924: \lambda^{-1}F\cdot \widehat{\xi}_{-1-4p}\cdot F^{-1}dz + \lambda^{4p}R_\lambda
925: dz\right)_\Omega\\
926: & = & \displaystyle (\lambda^{-2}-1)F\cdot \widehat{\xi}_{-2-4p}\cdot F^{-1}dz +
927: (\lambda^{-1}-1)F\cdot \widehat{\xi}_{-1-4p}\cdot F^{-1}dz\\
928: && \displaystyle - (\lambda-1)\left(F\cdot \widehat{\xi}_{-1-4p}\cdot
929: F^{-1}\right)^\dagger d\bar{z} - (\lambda^{2}-1)\left(F\cdot
930: \widehat{\xi}_{-2-4p}\cdot F^{-1}\right)^\dagger d\bar{z}.
931: \end{eqnarray*}
932: But relation (\ref{2.1.laxqfn2}) implies in particular that
933: $\widehat{\xi}_{-2-4p} = A_2'(\partial /\partial z)$ and $\widehat{\xi}_{-1-4p}
934: = A_{-1}(\partial /\partial z)$.  So we deduce that
935: \[
936: \displaystyle \left(\lambda^{4p}\eta_\lambda dz\right)_\Omega = F\cdot
937: \left(A_\lambda - A\right) \cdot F^{-1} = \Gamma_\lambda.
938: \]
939: Hence $E_\lambda$ can be constructed by solving a system analogous to
940: (\ref{2.1.laxfn}), (\ref{2.1.laxfn2}), i.e.
941: \begin{equation}\label{2.2.laxfn}
942: d\eta_\lambda + \left[\Gamma_\lambda,\eta_\lambda\right] = 0 \quad \hbox{and}
943: \quad \Gamma_\lambda = \left(\lambda^{4p}\eta_\lambda dz\right)_\Omega.
944: \end{equation}
945: Conversely a similar computation shows that a solution of (\ref{2.2.laxfn})
946: gives rise to a quasi-finite type family of solutions by an inverse
947: transformation, but we shall prove more in the next section.\\
948: 
949: \noindent Note that system (\ref{2.2.laxfn}) can also be interpreted as a pair
950: of commuting ordinary differential equations in the finite dimensional space
951: $\Lambda^{2+4p}\u(3):= \{S^1\ni\lambda\longmapsto\eta_\lambda \in \u(3)/
952: \eta_\lambda =\sum_{k=-2-4p}^{2+4p}\widehat{\eta}_k\lambda^k\}$.  It is the
953: analogue of the definition of a finite type solution according to \cite{BFPP}.\\
954: 
955: 
956: \subsection{Quasi-finite type solutions are actually finite type}
957: 
958: \noindent We show here the following
959: \begin{theorem}\label{3.3.qtf=tf}
960: For any family $F_\lambda$ of Hamiltonian stationary Lagrangian conformal
961: immersions of quasi-finite type, i.e.\,such that there exists
962: $\xi_\lambda:\Omega\longrightarrow \Lambda^{2+4p}\u(3)_\tau$ which satisfies
963: (\ref{2.1.laxfn}) and (\ref{2.1.laxqfn2}), there exists a gauge transformation
964: $F_\lambda\longmapsto F_\lambda^G:=F_\lambda\cdot G$, where $G\in {\cal
965: C}^\infty(\Omega,U(3)_0)$, such that $F_\lambda^G$ is of finite type.  More
966: precisely, denoting by $A_\lambda^G:=G^{-1}\cdot A_\lambda\cdot G + G^{-1}\cdot
967: dG$ and $\xi_\lambda^G:=G^{-1}\cdot \xi_\lambda\cdot G$, then $d\xi_\lambda^G
968: +\left[A_\lambda^G,\xi_\lambda^G\right] = G^{-1}\cdot \left(d\xi_\lambda
969: +[A_\lambda,\xi_\lambda]\right)\cdot G = 0$ and $(\lambda^{4p}\xi_\lambda^G
970: dz)_{\Lambda_{\su}} =A_\lambda^G$.
971: \end{theorem}
972: \begin{proof} We set $E_\lambda:=F_\lambda\cdot F^{-1}$,
973: $\Gamma_\lambda:=E_\lambda^{-1}\cdot dE_\lambda$ and $\eta_\lambda:= F\cdot
974: \xi_\lambda\cdot F^{-1}$ and will use the results of the previous section.\\
975: 
976: {\em A constant in $\Lambda^{2+4p}\u(3)_\tau$ associated to the quasi-finite
977: type family} --- First (\ref{2.2.laxfn}), which is a reformulation of
978: (\ref{2.1.laxfn}), implies
979: \[
980: d\left(E_\lambda\cdot \eta_\lambda\cdot E_\lambda^{-1}\right) = E_\lambda\cdot
981: \left( d\eta_\lambda +[\Gamma_\lambda,\eta_\lambda]\right)\cdot E_\lambda^{-1}
982: =0.
983: \]
984: Hence
985: \[
986: \eta_\lambda^0:= E_\lambda\cdot \eta_\lambda\cdot E_\lambda^{-1}
987: \]
988: is a constant in $\Lambda \u(3)$.  Moreover
989: \[
990: \eta_\lambda^0 = E_\lambda(z_0)\cdot \eta_\lambda(z_0)\cdot E_\lambda^{-1}(z_0)
991: = \eta_\lambda(z_0) = F(z_0)\cdot \xi_\lambda(z_0)\cdot F^{-1}(z_0)
992: =\xi_\lambda(z_0),
993: \]
994: which proves that $\eta_\lambda^0\in \Lambda^{2+4p}\u(3)_\tau$.\\
995: 
996: {\em An auxiliary map into $\Lambda^+U(3)^{\C}$} --- We let
997: \[
998: \Theta_\lambda:= \left(\lambda^{4p}\eta_\lambda dz\right)_{\Lambda^+} =
999: \lambda^{4p}\eta_\lambda dz - \left(\lambda^{4p}\eta_\lambda dz\right)_\Omega.
1000: \]
1001: Then using (\ref{2.2.laxfn}) we have $\Theta_\lambda = \lambda^{4p}\eta_\lambda
1002: dz - \Gamma_\lambda$ and so
1003: \[
1004: d\Gamma_\lambda + \Gamma_\lambda\wedge \Gamma_\lambda + d\Theta_\lambda -
1005: \Theta_\lambda\wedge \Theta_\lambda = -\lambda^{4p} \left(d\eta_\lambda
1006: +[\Gamma_\lambda,\eta_\lambda]\right) \left({\partial \over \partial
1007: \bar{z}}\right)dz\wedge d\bar{z} =0.
1008: \]
1009: But since $d\Gamma_\lambda + \Gamma_\lambda\wedge \Gamma_\lambda = 0$ this
1010: implies that $d\Theta_\lambda - \Theta_\lambda\wedge \Theta_\lambda = 0$.  Hence
1011: $\exists!  V_\lambda:\Omega\longrightarrow \Lambda^+U(3)^{\C}$ such that
1012: \[
1013: dV_\lambda = \Theta_\lambda\cdot V_\lambda \quad \hbox{and}\quad V_\lambda (z_0)
1014: = 1.
1015: \]
1016: Now, starting from $\lambda^{4p}\eta_\lambda dz =
1017: \Gamma_\lambda+\Theta_\lambda$, we deduce that
1018: \[
1019: \begin{array}{ccl}
1020: \displaystyle \lambda^{4p}\eta_\lambda^0dz & = & \displaystyle E_\lambda\cdot
1021: \Gamma_\lambda\cdot E_\lambda^{-1} + E_\lambda\cdot \Theta_\lambda\cdot
1022: E_\lambda^{-1}\\
1023: & = & \displaystyle dE_\lambda\cdot E_\lambda^{-1} + E_\lambda\cdot
1024: dV_\lambda\cdot V_\lambda^{-1}\cdot E_\lambda^{-1}\\
1025: & = & \displaystyle d\left(E_\lambda\cdot V_\lambda\right)\left(E_\lambda\cdot
1026: V_\lambda\right)^{-1},
1027: \end{array}
1028: \]
1029: which can be integrated into the relation
1030: \[
1031: E_\lambda\cdot V_\lambda = e^{\lambda^{4p}(z-z_0)\eta_\lambda^0}.
1032: \]
1033: 
1034: {\em An Iwasawa decomposition of $e^{\lambda^{4p}(z-z_0)\eta_\lambda^0}$} ---
1035: The latter implies
1036: \[
1037: e^{\lambda^{4p}(z-z_0)\eta_\lambda^0} = F_\lambda\cdot F^{-1}\cdot V_\lambda.
1038: \]
1039: >From this relation and the fact that $\eta_\lambda^0$ and $F_\lambda$ are
1040: twisted we deduce that $W_\lambda:= F^{-1}\cdot V_\lambda$ is twisted.  It is
1041: also a map with values in $\Lambda^+U(3)^{\C}_\tau$.  However it may not be not
1042: in $\Lambda^+_{\mathfrak{B}}U(3)^{\C}_\tau$ in general, because in the
1043: development
1044: \[
1045: F^{-1}\cdot V_\lambda = \widehat{W}_0 + \sum_{k=1}^\infty
1046: \widehat{W}_k\lambda^k,
1047: \]
1048: we are not sure that $\widehat{W}_0$ takes values in $\mathfrak{B}$.  But it
1049: takes values in $U(3)^{\C}_0$, so by using the Iwasawa decomposition
1050: $U(3)^{\C}_0 = U(3)_0\cdot \mathfrak{B}$ we know that $\exists !  G\in U(3)_0$,
1051: $\exists !\widehat{B}_0\in \mathfrak{B}$, $\widehat{W}_0 = G\cdot
1052: \widehat{B}_0$.  Hence
1053: \[
1054: G^{-1}\cdot F^{-1}\cdot V_\lambda = \widehat{B}_0 + \sum_{k=1}^\infty
1055: G^{-1}\cdot \widehat{W}_k\lambda^k
1056: \]
1057: takes values in $\Lambda^+_{\mathfrak{B}}U(3)^{\C}_\tau$.  So the splitting
1058: \[
1059: e^{\lambda^{4p}(z-z_0)\eta_\lambda^0} = \left(F_\lambda\cdot G\right)
1060: \left(G^{-1}\cdot F^{-1}\cdot V_\lambda\right)
1061: \]
1062: exactly reproduces the Iwasawa decomposition $\Lambda U(3)_\tau^{\C} = \Lambda
1063: U(3)_\tau \cdot \Lambda^+_{\mathfrak{B}}U(3)^{\C}_\tau$ proved in \cite{DPW}.\\
1064: 
1065: \noindent {\em Conclusion} --- Let us denote by $F_\lambda^G:= F_\lambda\cdot
1066: G$, $A_\lambda^G:= \left(F_\lambda^G\right)^{-1}\cdot dF_\lambda^G = G^{-1}\cdot
1067: A_\lambda\cdot G + G^{-1}\cdot dG$ and $B_\lambda^G:= G^{-1}\cdot F^{-1}\cdot
1068: V_\lambda$ and let us introduce
1069: \[
1070: \xi_\lambda^G:= \left(F_\lambda^G\right)^{-1}\cdot \eta_\lambda^0\cdot
1071: F_\lambda^G.
1072: \]
1073: (These definitions imply immediately $d\xi_\lambda^G +
1074: [A_\lambda^G,\xi_\lambda^G] = 0$.)  The first main observation is that the
1075: relation $\eta_\lambda^0 = E_\lambda\cdot \eta_\lambda\cdot E_\lambda^{-1} =
1076: F_\lambda\cdot F^{-1}\cdot \eta_\lambda\cdot F\cdot F_\lambda^{-1} =
1077: F_\lambda\cdot \xi_\lambda\cdot F_\lambda^{-1}$ implies
1078: \begin{equation}\label{2.3.tf1}
1079: \xi_\lambda^G = G^{-1}\cdot F_\lambda^{-1}\cdot \eta_\lambda^0\cdot
1080: F_\lambda\cdot G = G^{-1}\cdot \xi_\lambda\cdot G.
1081: \end{equation}
1082: Second, from the relation
1083: \[
1084: \begin{array}{ccl}
1085: \displaystyle \lambda^{4p}\eta^0_\lambda dz & = & \displaystyle
1086: d\left(e^{\lambda^{4p}(z-z_0)\eta_\lambda^0}\right) \cdot
1087: e^{-\lambda^{4p}(z-z_0)\eta_\lambda^0}\\
1088:  & = & \displaystyle d\left(F_\lambda^G\cdot B_\lambda^G\right)\cdot
1089:  \left(F_\lambda^G\cdot B_\lambda^G\right)^{-1}\\
1090: & = & \displaystyle dF_\lambda^G\cdot \left(F_\lambda^G\right)^{-1} +
1091: F_\lambda^G\cdot dB_\lambda^G\cdot \left(B_\lambda^G\right)^{-1}\cdot
1092: \left(F_\lambda^G\right)^{-1},
1093: \end{array}
1094: \]
1095: we deduce that
1096: \[
1097: \lambda^{4p}\xi_\lambda^G dz = \left(F_\lambda^G\right)^{-1} \cdot \left(
1098: \lambda^{4p}\eta_\lambda^0 dz\right)\cdot F_\lambda^G =
1099: \left(F_\lambda^G\right)^{-1}\cdot dF_\lambda^G + dB_\lambda^G\cdot
1100: \left(B_\lambda^G\right)^{-1}.
1101: \]
1102: Hence, since $B_\lambda^G$ takes values in
1103: $\Lambda^+_{\mathfrak{B}}U(3)^{\C}_\tau$,
1104: \begin{equation}\label{2.3.tf2}
1105: A_\lambda^G = \left(F_\lambda^G\right)^{-1}\cdot dF_\lambda^G = \left(
1106: \lambda^{4p}\xi_\lambda^Gdz\right)_{\Lambda_{\su}}.
1107: \end{equation}
1108: And relations (\ref{2.3.tf1}) and (\ref{2.3.tf2}) lead to the conclusion.
1109: \end{proof}
1110: 
1111: 
1112: 
1113: \section{All Hamiltonian stationary Lagrangian tori are of finite type}
1114: The subject of this section is to prove the following:
1115: \begin{theorem}\label{4.MainTheo}
1116: Let $u:\mathbb{C}\longrightarrow \mathbb{C}P^2$ be a doubly periodic Hamiltonian
1117: stationary Lagrangian conformal immersion.  Then $u$ is of finite type.
1118: \end{theorem}
1119: We will actually prove a slightly more general result, since we can replace the
1120: doubly periodicity assumption by the hypothesis that the Maurer--Cartan form of
1121: any {\em Legendrian framing} of $u$ is doubly periodic.  This result of course
1122: implies immediately that Hamiltonian stationary Lagrangian tori are of finite
1123: type, since they always can be covered conformally by the plane.\\
1124: 
1125: \noindent Note also that the study of Hamiltonian stationary Lagrangian tori
1126: splits into exactly two subcases: the {\em minimal} Lagrangian tori and the {\em
1127: non minimal} Hamiltonian stationary Lagrangian ones.  The first case occurs when
1128: the Lagrangian angle function along any Legendrian lift is locally constant, the
1129: second one when this function is harmonic and non constant.  In the case of
1130: minimal Lagrangian surfaces, Theorem \ref{4.MainTheo} is a special case of the
1131: result in \cite{BFPP}, since in this case $u$ is a harmonic map into
1132: $\mathbb{C}P^2$, as discussed in \cite{McI1}, \cite{McI2}, \cite{McI3} and
1133: \cite{J}.  The non minimal case however is not covered by the theory in
1134: \cite{BFPP} and is the subject of this section.  \\
1135: 
1136: \noindent Let $F:\mathbb{C}\longrightarrow U(3)$ be a Legendrian framing of $u$,
1137: $A:=F^{-1}\cdot dF$ its Maurer--Cartan form and $A_\lambda$ the family of
1138: deformations of $A$ as defined by (\ref{1.6.Alambda}).  The first basic
1139: observation is that $A_2\left({\partial \over \partial z}\right)$ is holomorphic
1140: and doubly periodic on $\C$, hence constant.  Thus two cases occur: either
1141: $A_2\left({\partial \over \partial z}\right)=0$, which corresponds to the
1142: minimal case that we exclude here, or $A_2\left({\partial \over \partial
1143: z}\right)$ is a constant different from 0, the case that we consider next.\\
1144: 
1145: \noindent In order to show Theorem \ref{4.MainTheo} we need to prove that there
1146: exists some $p\in \mathbb{N}$ and a map $\xi_\lambda:\mathbb{C}\longrightarrow
1147: \Lambda^{2+4p}\u(3)_\tau$ such that $d\xi_\lambda = [\xi_\lambda,A_\lambda]$ and
1148: $A_\lambda = (\lambda^{4p}\xi_\lambda dz)_{\Lambda_{\su}}$.  But thanks to
1149: Theorem \ref{3.3.qtf=tf} it will enough to prove that $A_\lambda -
1150: (\lambda^{4p}\xi_\lambda dz)_{\Lambda_{\su}}$ is a 1-form with coefficients in
1151: $\u(3)_0$.  Our proof here follows a strategy inspired from \cite{BFPP}: a first
1152: step consists in building a formal series $Y_\lambda = \sum_{k=-2}^\infty
1153: \widehat{Y}_k\lambda^k$ which is a solution of $dY_\lambda =
1154: [Y_\lambda,A_\lambda]$.  Such a series is called a {\em formal Killing field}.
1155: We will also require $Y_\lambda$ to be {\em quasi-adapted}, i.e.\,is such that
1156: \begin{equation}\label{4.adapted}
1157: (Y_\lambda dz)_{\Lambda_{\su}} = A_\lambda + B,\quad \hbox{where the
1158: coefficients of }B\hbox{ are in } \u(3)_0.
1159: \end{equation}
1160: This is achieved through a recursion procedure.\\
1161: 
1162: \noindent In a second step we will show that the coefficients of $Y_\lambda$
1163: form a countable collection of doubly periodic functions satisfying an elliptic
1164: PDE and hence, by using a compactness argument, we conclude that they are
1165: contained in a finite dimensional space.  Then we deduce the existence of
1166: $\xi_\lambda$ using linear algebra.
1167: 
1168: 
1169: 
1170: 
1171: \subsection{Construction of an adapted formal Killing field}
1172: 
1173: We first introduce some notations.  We denote by
1174: \[
1175: \pi_0^\perp := \left(\begin{array}{ccc} 1 &&\\ & 1& \\ && 0 \end{array}\right)
1176: \]
1177: and $a := {1\over 2}{\partial \beta\over \partial z}$ (here a constant different
1178: from 0).  Then $A_2' = ia\pi_0^\perp dz$.  We will also set
1179: $X:=A_{-1}\left({\partial \over \partial z}\right)$ and $C:=A_0\left({\partial
1180: \over \partial z}\right)$, so that
1181: \[
1182: A_\lambda = \lambda^{-2}ia\pi_0^\perp dz + \lambda^{-1}Xdz + Cdz -C^\dagger
1183: d\bar{z} -\lambda X^\dagger d\bar{z} + \lambda^2i\overline{a}\pi_0^\perp
1184: d\bar{z}.
1185: \]
1186: We also introduce the linear map $\ad \pi_0^\perp : \u(3)^{\C}\longrightarrow
1187: \u(3)^{\C}$, acting by $\xi\longmapsto [\pi_0^\perp ,\xi]$\footnote{Actually the
1188: map $\xi\longmapsto i[\pi_0^\perp ,\xi]$ corresponds to the complex structure on
1189: the Legendrian distribution.}.  We observe that $\pi_0^\perp $ commutes with the
1190: elements in $\u(3)_0^{\C}$ and $\u(3)_2^{\C}$.  Moreover
1191: \[
1192: \forall a,b\in \C, \; \left[\pi_0^\perp ,\left(\begin{array}{ccc}&& a \\ && b \\
1193: \mp ib &\pm ia \end{array}\right)\right] = \left(\begin{array}{ccc} && a \\ && b
1194: \\ \pm ib & \mp ia& \end{array}\right),
1195: \]
1196: that is $\ad \pi_0^\perp $ maps $\u(3)_{\mp 1}^{\C}$ to $\u(3)_{\pm 1}^{\C}$.
1197: >From that we deduce that $V:= \Ker \ad \pi_0^\perp $ coincides with
1198: $\u(3)_0^{\C}\oplus \u(3)_2^{\C}$ and $V^\perp:= \Im \ad \pi_0^\perp $ coincides
1199: with $\u(3)_{-1}^{\C}\oplus \u(3)_1^{\C}$ (note that $V^\perp$ is actually the
1200: orthogonal subspace to $V$ in $\u(3)^{\C}$).  In our construction we will use
1201: extensively the following properties:
1202: \begin{itemize}
1203: \item the map $\ad \pi_0^\perp \at{V^\perp \to V^\perp}$ is a vector space
1204: isomorphism (it is actually a involution on $V^\perp$), \item the inclusions $VV
1205: \subset V$, $VV^\perp \subset V^\perp$, $V^\perp V \subset V^\perp$ and $V^\perp
1206: V^\perp \subset V$.  These properties can be checked by a direct computation
1207: using the fact that matrices in $V$ are diagonal by blocks and the matrices in
1208: $V^\perp$ are off-diagonal by blocks.  (The three first properties can also be
1209: deduced from the definition of $V$ and $V^\perp$ and the fact that $\ad$ is a
1210: derivation).
1211: \end{itemize}
1212: We look for a formal Killing field $Y_\lambda$, i.e.\,a solution of the equation
1213: \begin{equation} \label{eq:champ_formel}
1214: dY_\lambda = [Y_\lambda,A_\lambda],
1215: \end{equation}
1216: of the form $Y_\lambda = (1+W_\lambda)^{-1}\lambda^{-2}ia\pi_0^\perp
1217: (1+W_\lambda)$, where $W_\lambda = \sum_{k=0}^\infty \widehat{W}_k\lambda^k$ as
1218: in \cite{BFPP}.  In order to have a well-posed problem (and in particular to
1219: guarantee the existence of an unique solution of this type) we assume that
1220: $W_\lambda$ takes values in $V^\perp$.  We start by evaluating
1221: (\ref{eq:champ_formel}) along $\partial/\partial z$.  It gives, after
1222: conjugation by $1+W_\lambda$:
1223: \begin{multline} \label{eq:W}
1224: \lambda^{-2}\frac{\partial a}{\partial z} \pi_0^\perp + \lambda^{-2}
1225: a\big[\pi_0^\perp , \\
1226:  \frac{\partial W_\lambda}{\partial z}(1+W_\lambda)^{-1}
1227:  -(1+W_\lambda)(\lambda^{-2}ia\pi_0^\perp
1228:  +\lambda^{-1}X+C)(1+W_\lambda)^{-1}\big] =0
1229: \end{multline}
1230: Here the fact that $a$ is a constant leads to an immediate simplification,
1231: namely that the bracket in the left hand side of (\ref{eq:W}) is 0.  Thus
1232: equation (\ref{eq:W}) implies that $\frac{\partial W_\lambda}{\partial z}
1233: (1+W_\lambda)^{-1}-(1+W_\lambda)(\lambda^{-2}ia\pi_0^\perp +\lambda^{-1}X+C)
1234: (1+W_\lambda)^{-1}$ lies in $V$, hence there exists a map
1235: $\varphi_\lambda:\C\longrightarrow V$ such that
1236: \[
1237: \pa{W_\lambda}{z}(1+W_\lambda)^{-1} -(1+W_\lambda)(\lambda^{-2}ia\pi_0^\perp
1238: +\lambda^{-1}X+C)(1+W_\lambda)^{-1} = \varphi_\lambda
1239: \]
1240: or
1241: \[
1242: \pa{W_\lambda}{z}-(1+W_\lambda)(\lambda^{-2}ia\pi_0^\perp +\lambda^{-1}X+C) =
1243: \varphi_\lambda(1+W_\lambda),
1244: \]
1245: which can be projected according to the splitting $V \oplus V^\perp$ as
1246: \[
1247: \left\{ \begin{array}{l}\displaystyle \lambda^{-2}ia\pi_0^\perp
1248: +\lambda^{-1}W_\lambda X+C = -\varphi_\lambda \in V
1249: \\
1250: \displaystyle \pa{W_\lambda}{z} - \lambda^{-2}ia W_\lambda \pi_0^\perp -
1251: \lambda^{-1}X - W_\lambda C = \varphi_\lambda W_\lambda \in V^\perp.
1252: \end{array} \right.
1253: \]
1254: Substituting $\varphi_\lambda$,
1255: \[
1256: \pa{W_\lambda}{z} - \lambda^{-2}ia W_\lambda \pi_0^\perp - \lambda^{-1}X-
1257: W_\lambda C + \lambda^{-2}ia\pi_0^\perp W_\lambda+\lambda^{-1}W_\lambda X
1258: W_\lambda +C W_\lambda = 0
1259: \]
1260: or
1261: \[
1262: [ia\pi_0^\perp ,W_\lambda] + \lambda(W_\lambda X W_\lambda-X) +
1263: \lambda^2[C,W_\lambda] + \lambda^2 \pa{W_\lambda}{z} = 0
1264: \]
1265: or
1266: \begin{multline*}
1267: ia\sum_{n\geq 0} [\pi_0^\perp ,\widehat{W}_n]\lambda^n + \sum_{n\geq
1268: 1}\left(\sum_{k=0}^{n-1} \widehat{W}_k X \widehat{W}_{n-1-k}\right) \lambda^n -
1269: \lambda X \\
1270: + \sum_{n\geq 2} \left([C,\widehat{W}_{n-2}] + \pa{\widehat{W}_{n-2}}{z} \right)
1271: \lambda^n = 0.
1272: \end{multline*}
1273: Hence
1274: \[
1275: \left\{ \begin{array}{ll} \displaystyle n=0, & \displaystyle ia[\pi_0^\perp
1276: ,\widehat{W}_0] = 0
1277: \\
1278: \displaystyle n=1, & \displaystyle ia[\pi_0^\perp ,\widehat{W}_1]+\widehat{W}_0
1279: X \widehat{W}_0 - X = 0
1280: \\
1281: \displaystyle n \geq 2, & \displaystyle ia[\pi_0^\perp
1282: ,\widehat{W}_n]+\sum_{k=0}^{n-1} \widehat{W}_k X \widehat{W}_{n-1-k} +
1283: [C,\widehat{W}_{n-2}] + \pa{\widehat{W}_{n-2}}{z} =0 \end{array} \right.
1284: \]
1285: and thus
1286: \[
1287: \left\{ \begin{array}{l} \widehat{W}_0 = 0
1288: \\
1289: \widehat{W}_1 = -ia^{-1}[\pi_0^\perp ,X]
1290: \\
1291: \displaystyle \widehat{W}_n = ia^{-1}\left[\pi_0^\perp ,\sum_{k=0}^{n-1}
1292: \widehat{W}_k X \widehat{W}_{n-1-k} + [C,\widehat{W}_{n-2}] +
1293: \pa{\widehat{W}_{n-2}}{z}\right] \end{array} \right.
1294: \]
1295: We observe that the formal Killing field is {\em quasi-adapted} in the sense
1296: that the two first coefficients are the right ones:
1297: \begin{eqnarray*}
1298: Y_\lambda &=& ia\lambda^{-2} (1+W_\lambda)^{-1}\pi_0^\perp (1+W_\lambda) = ia
1299: \lambda^{-2} \left( \pi_0^\perp - \lambda [\widehat{W}_1,\pi_0^\perp ] +
1300: \O(\lambda^2) \right) \\
1301: &=& \lambda^{-2} i a \pi_0^\perp + \lambda^{-1} X + \O(1).
1302: \end{eqnarray*}
1303: Another pleasant property is that this formal field is automatically twisted (as
1304: in the case of $\C^2$, see \cite{HR1}).  Indeed using the fact that $\tau$ is an
1305: automorphism for the product of matrices as well as for the Lie bracket (and so
1306: $[\u(3)^{\C}_a,\u(3)^{\C}_b] \subset \u(3)^{\C}_{a+b}$ and $\u(3)^{\C}_a
1307: \u(3)^{\C}_b \subset \u(3)^{\C}_{a+b}$), we obtain that
1308: \[
1309: \tau(Y_\lambda) = \tau(1+W_\lambda)^{-1} (-\lambda^2)ia\pi_0^\perp
1310: \tau(1+W_\lambda).
1311: \]
1312: Thus it is enough to show that $1+W_\lambda$ is twisted, i.e.\,$\tau(W_\lambda)
1313: = W_{i\lambda}$.  In terms of the Fourier decomposition of $W_\lambda$ this is
1314: equivalent to proving that $\widehat{W}_n$ belongs to $\u(3)^{\C}_n$.  Let us
1315: prove it by recursion.  We already know that $\widehat{W}_1 =
1316: -ia^{-1}[\pi_0^\perp ,X]$ is in $\u(3)^{\C}_1$.  Assume that the result is true
1317: up to $n-1$,then
1318: \[
1319: \sum_{p=0}^{n-1} \widehat{W}_p X \widehat{W}_{n-1-p} + [C,\widehat{W}_{n-2}] +
1320: \pa{\widehat{W}_{n-2}}{z}
1321: \]
1322: belongs to $\u(3)^{\C}_{n-2}$.  And since $\pi_0^\perp \in \u(3)^{\C}_2$,
1323: $\widehat{W}_n$ is in $\gC_n$.
1324: 
1325: We now prove that (\ref{eq:champ_formel}) is also true along $\partial/\partial
1326: \bar{z}$.  We follow here the same kind of arguments as in \cite{BFPP} slightly
1327: simplified\footnote{essentially the simplifications occur because the
1328: semi-simple term $B$ of \cite{BFPP} is here $ia\pi_0^\perp $ which is constant
1329: for $d$, so we do not need to introduce a flat connection.}.  We want to show
1330: that
1331: \[
1332: \frac{\partial Y_\lambda}{\partial \bar{z}} +
1333: \left[A_\lambda\left(\frac{\partial }{\partial \bar{z}}\right), Y_\lambda\right]
1334: = 0
1335: \]
1336: and for that purpose we rather consider the conjugate of the left hand side
1337: $\zeta_\lambda = (1+W_\lambda) (\frac{\partial Y_\lambda}{\partial \bar{z}} +
1338: [A_\lambda\left(\frac{\partial }{\partial \bar{z}}\right),
1339: Y_\lambda])(1+W_\lambda)^{-1}$.  We then prove two facts
1340: \begin{itemize}
1341: \item $\zeta_\lambda$ takes its values in $V^\perp$: this follows from the
1342: identity
1343: \[
1344: \zeta_\lambda = \left[\lambda^{-2}ia\pi_0^\perp , \frac{\partial
1345: W_\lambda}{\partial \bar{z}} (1+W_\lambda)^{-1} -
1346: (1+W_\lambda)A_\lambda\left(\frac{\partial }{\partial \bar{z}}\right)
1347: (1+W_\lambda)^{-1}\right].
1348: \]
1349: Note that since $\zeta_\lambda$ is twisted the fact that $\zeta_\lambda\in
1350: V^\perp$ implies also that $\zeta_\lambda$ is an odd function of $\lambda$
1351: and so that
1352: \begin{equation}\label{3.1.oddzeta}
1353: \zeta_\lambda = \sum_{k=0}^\infty \widehat{\zeta}_{2k-1}\lambda^{2k-1}.
1354: \end{equation}
1355: 
1356: \item the relation
1357: \begin{equation}\label{3.1.dzeta}
1358: {\partial \zeta_\lambda\over \partial z} = [\varphi_\lambda,\zeta_\lambda].
1359: \end{equation}
1360: Indeed $d+\ad A_\lambda$ is a flat connection and in particular
1361: $\frac{\partial }{\partial z} +\ad A_\lambda(\frac{\partial }{\partial z})$
1362: commutes with $\frac{\partial }{\partial \bar{z}} +A_\lambda(\frac{\partial
1363: }{\partial \bar{z}})$.  Hence
1364: \[
1365: \left( \frac{\partial }{\partial z}+\ad A_\lambda\left(\frac{\partial
1366: }{\partial z}\right) \right) \left( \frac{\partial }{\partial \bar{z}} +\ad
1367: A_\lambda\left(\frac{\partial }{\partial \bar{z}}\right) \right) Y_\lambda
1368: = 0
1369: \]
1370: i.e.
1371: \[
1372: \left( \frac{\partial }{\partial z}+\ad A_\lambda\left(\frac{\partial
1373: }{\partial z}\right) \right)
1374: \left((1+W_\lambda)^{-1}\zeta_\lambda(1+W_\lambda)\right) = 0
1375: \]
1376: Thus (\ref{3.1.dzeta}) follows from a computation which uses
1377: $\varphi_\lambda = {\partial W_\lambda\over\partial z}(1+W_\lambda)^{-1} -
1378: (1+W_\lambda) A_\lambda\left(\frac{\partial }{\partial
1379: z}\right)(1+W_\lambda)^{-1}$.
1380: 
1381: \end{itemize}
1382: Now assume by contradiction that $\zeta_\lambda\neq 0$: in view of
1383: (\ref{3.1.oddzeta}) there exists an integer $k\in \mathbb{N}$ such that
1384: $\widehat{\zeta}_{2k-1}\neq 0$ and $\widehat{\zeta}_{2k-3} = 0$.  By
1385: substituting the Fourier decompositions in (\ref{3.1.dzeta}) and observing that
1386: the Fourier series expansion of $\varphi_\lambda$ starts by $\lambda^{-2}
1387: ia\pi_0^\perp$, we deduce that $0 = \partial \widehat{\zeta}_{2k-3}/\partial z =
1388: [ia\pi_0^\perp , \widehat{\zeta}_{2k-1}]$; but $\ad \pi_0^\perp $ is invertible
1389: on $V^\perp$ and hence $\widehat{\zeta}_{2k-1} = 0$.  So we get a contradiction.
1390: 
1391: \subsection{Polynomial Killing fields}
1392: We now deduce the existence of a non-trivial polynomial Killing field.
1393: 
1394: A first easy consequence of the results of the previous section is that, for all
1395: $n\in \N$ and for all polynomial of the form $P(\lambda)=a_n\lambda ^{-4n} +
1396: a_{n-1}\lambda ^{-4(n-1)} + \cdots + a_0$, where $a_0,a_1,\cdots ,a_n\in \C$ and
1397: $a_n\neq 0$, then $Z_\lambda:=P(\lambda)Y_\lambda$ is again formal Killing
1398: field.  Moreover it is quasi-adapted (modulo the multiplicative factor
1399: $a_n\lambda^{-4n}$), i.e.\,the lower degree terms are $a_n\lambda^{-4n}\left(ia
1400: \lambda^{-2} \pi_0^\perp + \lambda^{-1} X + O(\lambda ^0)\right)$).  Let us
1401: consider
1402: \[
1403: Z_\leq := \sum_{k=-2-4n}^0\widehat{Z}_k\lambda^k,\quad \hbox{and}\quad Z_>:=
1404: \sum_{k=1}^\infty \widehat{Z}_k\lambda^k,
1405: \]
1406: so that $Z_\lambda= Z_\leq + Z_>$.  We study
1407: \begin{equation}\label{dZ-}
1408: R_\lambda:= dZ_\leq +[A_\lambda,Z_\leq].
1409: \end{equation}
1410: We first remark that $R_\lambda$ is necessarily of the form
1411: $R_\lambda=\sum_{k=-4-4n}^2\widehat{R}_k\lambda^k$.  But because of $dZ_\lambda
1412: + [A_\lambda,Z_\lambda]= 0$, we also have
1413: \begin{equation}\label{dZ+}
1414: R_\lambda = -dZ_> - [A_\lambda,Z_>],
1415: \end{equation}
1416: which implies $R_\lambda=\sum_{k=-1}^\infty \widehat{R}_k\lambda^k$.  Hence
1417: finally
1418: \[
1419: R_\lambda = \lambda^{-1}\widehat{R}_{-1} + \widehat{R}_{0} +
1420: \lambda^{1}\widehat{R}_{1} + \lambda^{2}\widehat{R}_{2}.
1421: \]
1422: Each term $\widehat{R}_{k}$ can be evaluated through two different ways: by
1423: using (\ref{dZ-}) or (\ref{dZ+}).  From (\ref{dZ-}) we obtain
1424: \begin{equation}\label{dZ-z}
1425: \left\{ \begin{array}{l} \widehat{R}_{-1}(\partial _z) = \partial _z
1426: \widehat{Z}_{-1} + [A_{-1}(\partial _z), \widehat{Z}_0] + [A_{0}(\partial _z),
1427: \widehat{Z}_{-1}] \\
1428: \widehat{R}_{0}(\partial _z) = \partial _z \widehat{Z}_{0} + [A_{0}(\partial
1429: _z), \widehat{Z}_0] \\
1430: \widehat{R}_{1}(\partial _z) = 0\\
1431: \widehat{R}_{2}(\partial _z) = 0 \end{array}\right.
1432: \end{equation}
1433: and
1434: \begin{equation}\label{dZ-zbar}
1435: \left\{ \begin{array}{l} \widehat{R}_{-1}(\partial _{\bar{z}}) = \partial
1436: _{\bar{z}} \widehat{Z}_{-1} + [A_{0}(\partial_{\bar{z}}), \widehat{Z}_{-1}] +
1437: [A_{1}(\partial_{\bar{z}}), \widehat{Z}_{-2}] + [A_{2}''(\partial _{\bar{z}}),
1438: \widehat{Z}_{-3}] \\
1439: \widehat{R}_{0}(\partial _{\bar{z}}) = \partial_{\bar{z}} \widehat{Z}_{0} +
1440: [A_{0}(\partial _{\bar{z}}), \widehat{Z}_{0}] + [A_{1}(\partial _{\bar{z}}),
1441: \widehat{Z}_{-1}] + [A_{2}''(\partial _{\bar{z}}), \widehat{Z}_{-2}] \\
1442: \widehat{R}_{1}(\partial _{\bar{z}}) = [A_{1}(\partial _{\bar{z}}),
1443: \widehat{Z}_{0}] + [A_{2}''(\partial _{\bar{z}}), \widehat{Z}_{-1}] \\
1444: \widehat{R}_{2}(\partial _{\bar{z}}) = [A_{2}''(\partial _{\bar{z}}),
1445: \widehat{Z}_{0}].  \end{array}\right.
1446: \end{equation}
1447: >From (\ref{dZ+}) we get
1448: \begin{equation}\label{dZ+z}
1449: \left\{ \begin{array}{l} \widehat{R}_{-1}(\partial _z) = - [A_{2}'(\partial _z),
1450: \widehat{Z}_{1}] \\
1451: \widehat{R}_{0}(\partial _z) = - [A_{2}'(\partial _z), \widehat{Z}_{2}] -
1452: [A_{-1}(\partial _z), \widehat{Z}_{1}] \\
1453: \widehat{R}_{1}(\partial _z) = - \partial _z\widehat{Z}_{1} - [A_{2}'(\partial
1454: _z), \widehat{Z}_{3}] - [A_{-1}(\partial _z), \widehat{Z}_{2}] - [A_{0}(\partial
1455: _z), \widehat{Z}_{1}] \\
1456: \widehat{R}_{2}(\partial _z) = - \partial _z\widehat{Z}_{2} - [A_{2}'(\partial
1457: _z), \widehat{Z}_{4}] - [A_{-1}(\partial _z), \widehat{Z}_{3}] - [A_{0}(\partial
1458: _z), \widehat{Z}_{2}] \end{array}\right.
1459: \end{equation}
1460: and
1461: \begin{equation}\label{dZ+zbar}
1462: \left\{ \begin{array}{l} \widehat{R}_{-1}(\partial _{\bar{z}}) = 0 \\
1463: \widehat{R}_{0}(\partial _{\bar{z}}) = 0 \\
1464: \widehat{R}_{1}(\partial _{\bar{z}}) = - \partial _{\bar{z}} \widehat{Z}_{1} -
1465: [A_{0}(\partial _{\bar{z}}), \widehat{Z}_{1}] \\
1466: \widehat{R}_{2}(\partial _{\bar{z}}) = - \partial _{\bar{z}} \widehat{Z}_{2} -
1467: [A_{0}(\partial _{\bar{z}}), \widehat{Z}_{2}] - [A_{1}(\partial _{\bar{z}}),
1468: \widehat{Z}_{1}] .  \end{array}\right.
1469: \end{equation}
1470: Thus in order to obtain an expression of $R_\lambda$ which does depend only on
1471: $\widehat{Z}_{-1}$ and $\widehat{Z}_{0}$, we exploit (\ref{dZ-z}) and the two
1472: last equations in (\ref{dZ-zbar}).  But instead of using the two first equations
1473: of (\ref{dZ-zbar}) we take the two first ones of (\ref{dZ+zbar}).  This gives us
1474: \begin{equation}\label{Rz}
1475: R_\lambda(\partial _z) = \lambda^{-1}\left( \partial _z\widehat{Z}_{-1} +
1476: [A_{-1}(\partial _z),\widehat{Z}_0] + [A_{0}(\partial
1477: _z),\widehat{Z}_{-1}]\right) + \left( \partial _z\widehat{Z}_{0} +
1478: [A_{0}(\partial _z),\widehat{Z}_0] \right),
1479: \end{equation}
1480: \begin{equation}\label{Rzbar}
1481: R_\lambda(\partial _{\bar{z}}) = \lambda \left( [A_{1}(\partial
1482: _{\bar{z}}),\widehat{Z}_0] + [A_{2}''(\partial
1483: _{\bar{z}}),\widehat{Z}_{-1}]\right) + \lambda^2[A_{2}''(\partial
1484: _{\bar{z}}),\widehat{Z}_{0}].
1485: \end{equation}
1486: These relations will imply that $\widehat{Z}_{-1}$ and $\widehat{Z}_{0}$ satisfy
1487: a second order elliptic equation.  In order to prove that we need to establish
1488: another relation between $R_\lambda(\partial _z)$ and $R_\lambda(\partial
1489: _{\bar{z}})$.  For that purpose recall that $dA_\lambda + A_\lambda\wedge
1490: A_\lambda = 0$, which means that the connection $d+\ad A_\lambda$ has a
1491: vanishing curvature.  In particular
1492: \[
1493: 0= \left( d+\ad A_\lambda\right) \circ \left( d+\ad A_\lambda\right)Z_\leq =
1494: dR_\lambda + [A_\lambda \wedge R_\lambda].
1495: \]
1496: This implies
1497: \begin{equation}\label{fermeture}
1498: {\partial R_\lambda(\partial _z)\over \partial \bar{z}} - {\partial
1499: R_\lambda(\partial _{\bar{z}}) \over \partial z} = [A_\lambda(\partial
1500: _z),R_\lambda(\partial _{\bar{z}})] - [A_\lambda(\partial
1501: _{\bar{z}}),R_\lambda(\partial _z)].
1502: \end{equation}
1503: A substitution of (\ref{Rz}) and (\ref{Rzbar}) in (\ref{fermeture}) gives a
1504: system of linear elliptic equations on $\widehat{Z}_{-1}$ and $\widehat{Z}_{0}$.
1505: Since the space of solutions to this system which are periodic is finite
1506: dimensional, it turns out that $\widehat{Z}_{-1}$ and $\widehat{Z}_{0}$ belong
1507: to a finite dimensional vector space.  Hence relations (\ref{Rz}) and
1508: (\ref{Rzbar}) force $R_\lambda(\partial _z)$ and $R_\lambda(\partial
1509: _{\bar{z}})$ to stay in a finite dimensional vector space.
1510: 
1511: We can conclude: let us consider
1512: \[
1513: {\cal R}:= \{ R_\lambda / R_\lambda(\partial _z), R_\lambda(\partial
1514: _{\bar{z}})\hbox{ are given by (\ref{Rz}) and (\ref{Rzbar}) and satisfy
1515: (\ref{fermeture})}\}.
1516: \]
1517: It is a complex finite dimensional vector space.  Let us also denote by ${\cal
1518: P}_n:=\{P(\lambda)=a_n\lambda ^{-4n} + a_{n-1}\lambda ^{-4(n-1)} + \cdots + a_0/
1519: (a_0,\cdots , a_n)\in \C ^{n+1}\}$ and ${\cal P}_\infty :=\cup_{n\in \N}{\cal
1520: P}_n$.
1521: 
1522: The linear map ${\cal P}_\infty \ni P(\lambda)\longmapsto dZ_\leq
1523: +[A_\lambda,Z_\leq]$, where $Z_\leq = \left( P(\lambda)Y_\lambda\right)_\leq$
1524: takes values in ${\cal R}$ and so has a finite rank, say $n$.  Then since
1525: $\hbox{dim}_\mathbb{C} {\cal P}_n=n+1$, the map ${\cal P}_n \ni
1526: P(\lambda)\longmapsto dZ_\leq +[A_\lambda,Z_\leq]$ has a non trivial kernel: let
1527: $P(\lambda)=\sum_{k=0}^na_k\lambda^{-4k}$ be a non trivial polynomial in this
1528: kernel.  Let $4p$ be the degree of $P$ in $\lambda^{-1}$, i.e.\,such that
1529: $P(\lambda)=\sum_{k=0}^pa_k\lambda^{-4k}$ and $a_p\neq 0$.  Without loss of
1530: generality we can assume that $a_p = 1$.  Then $\xi_\lambda:=
1531: \left(P(\lambda)Y_\lambda\right)_\leq -
1532: \left(P(\lambda)Y_\lambda\right)_\leq^\dagger$ is a solution of
1533: (\ref{2.1.laxfn}) and (\ref{2.1.laxqfn2}).
1534: 
1535: 
1536: \section{Homogeneous tori in $\mathbb{C} P^2$}
1537: 
1538: We describe here the simplest examples of Hamiltonian stationary Lagrangian tori
1539: in $\mathbb{C}P^2$: the {{\em homogeneous}} Hamiltonian stationary Lagrangian
1540: tori, i.e. immersions $u$ of $S^1 \times S^1$ into $\mathbb{C} P^2$ such that $u
1541: ( x + t, y ) = e^{t A} u ( x, y )$ and $u ( x, y + t ) = e^{t B} u ( x, y )$ for
1542: some skew-Hermitian matrices $A$ and $B$.  Notice that $A$ and $B$ are only
1543: defined up to addition with a multiple of $i \mathit{Id}$.  The simplest example
1544: is the Clifford torus, namely the image by the Hopf map $\pi$ of the product
1545: torus $\{ z = ( z^1, z^2, z^3 ) ; | z^1 | = | z^2 | = | z^3 | = 1 / \sqrt{3}
1546: \}$.  This torus is minimal.  The main result states that all homogeneous
1547: Hamiltonian stationary Lagrangian tori are similar to the Clifford torus.
1548: 
1549: \begin{theorem} \label{homogeneous}
1550: Any homogeneous Hamiltonian stationary Lagrangian torus in $\mathbb{C} P^2$ is
1551: the image by the Hopf map of some Cartesian product $r_1 S^1 \times r_2 S^1
1552: \times r_3 S^1 = \{ z = ( z^1, z^2, z^3 ) ; | z^1 | = r_1, | z^2 | = r_2, | z^3
1553: | = r_3 \}$ where $r_1^2 + r_2^2 + r_3^2 = 1$, up to $U(3)$ congruence.
1554: Moreover, the torus is special Lagrangian if and only if $r_1 = r_2 = r_3 =
1555: \sqrt{3}$.
1556: \end{theorem}
1557: %
1558: \begin{proof}
1559: Let us first see why $\pi ( T )$ is a Hamiltonian stationary Lagrangian torus in
1560: $\mathbb{C} P^2$, where $T = r_1 S^1 \times r_2 S^1 \times r_3 S^1$.  Indeed it
1561: suffices to show that $\pi ( T )$ admits a Legendrian preimage.  Let \[ f ( x, y
1562: ) : = \left( r_1 e^{i (( 1 - r_1^2 ) x - r_2^2 y )}, r_2 e^{i ( - r_1^2 x + ( 1
1563: - r_2^2 ) y )}, r_3 e^{i ( - r_1^2 x - r_2^2 y )} \right) .
1564: \]
1565: Then the orbit under the Hopf action of the image of $f$ is exactly the 3-torus
1566: $T$ above and $\pi \circ f$ is doubly periodic with periods $(2\pi,0)$ and
1567: $(0,2\pi)$.  Note that this immersion is not conformal but there exists an
1568: orthonormal Hermitian moving frame $(e_1,e_2)$ such that ${\partial f\over
1569: \partial x} = r_1\sqrt{1-r_1^2}\, e_1$ and ${\partial f\over \partial y} =
1570: \frac{r_2}{\sqrt{1-r_1^2}} (r_3 e_2 - r_1r_2 e_1)$.  And it is easy to check
1571: that $f$ is Legendrian (and flat).  Its Lagrangian angle function is
1572: \[
1573: \beta ( x, y ) = x ( 1 - 3 r_1^2 ) + y ( 1 - 3 r_2^2 ) + \pi
1574: \]
1575: and since the metric is flat, $\beta$ is clearly harmonic, and constant if and
1576: only if $r_1 = r_2 = r_3 = 1 / \sqrt{3}$.  Notice that many of these tori do not
1577: lift up to $S^5$ as Legendrian tori (they do not close up).  Indeed the Maslov
1578: class is not always an integer: for the implicit homology basis, $t \longmapsto
1579: ( 2 \pi t, 0 )$ and $t \longmapsto ( 0, 2 \pi t )$, it is $( 1 - 3 r_1^2, 1 - 3
1580: r_2^2 )$.  However, if all $r_i^2$ are rational, the torus in $\mathbb{C} P^2$
1581: possesses a Legendrian toric multiple cover.
1582: 
1583: Suppose now that $u : S^1 \times S^1 \longrightarrow \mathbb{C} P^2$ is a
1584: homogeneous Lagrangian immersion.  According to our definition $u$ has a lift
1585: $\hat{u}$ such that $\pi ( \hat{u} ( x + t, y )) = \pi ( e^{t A} \hat{u} ( x, y
1586: ))$ and $\pi ( \hat{u} ( x, y + t )) = \pi ( e^{t B} \hat{u} ( x, y ))$.  In
1587: particular $\pi ( e^{x A} e^{y B} p ) = \pi ( e^{y B} e^{x A} p )$, for any $p
1588: \in S^5$ in the image.  However the image is never contained in a complex
1589: subspace of $\mathbb{C}^3$, hence $[ A, B ] \in i\mathbb{R} \mathit{Id}$.  Since
1590: $[ A, B ]$ is traceless, $A$ and $B$ commute.
1591: 
1592: The obvious (non Legendrian) lift in $S^5$ is $( x, y ) \longmapsto e^{x A} e^{y
1593: B} p$ where now $p = ( p_1, p_2, p_3 )$ is a fixed point mapped by the Hopf map
1594: $\pi$ to $u ( 0, 0 )$.  A Legendrian lift $\hat{u}$ takes the following form:
1595: $\hat{u} ( x, y ) = e^{i \theta ( x, y )} e^{x A} e^{y B} p$ for some function
1596: $\theta$.  The horizontality condition implies $\langle ( i\frac{\partial
1597: \theta}{\partial x} \mathit{Id} + A ) p, p \rangle_{\mathbb{C}^3} = 0$ so that
1598: $\frac{\partial \theta}{\partial x} = i \frac{\langle A p, p
1599: \rangle_{\mathbb{C}^3}}{| p |^2}$ is a constant.  The same holds in the $y$
1600: direction so we can define the lift $\hat{u} ( x, y ) = e^{x \hat{A} + y
1601: \hat{B}} p$ where $\hat{A} = A + i\frac{\partial \theta}{\partial x}
1602: \mathit{Id}$ and $\hat{B} = B + i\frac{\partial \theta}{\partial y} \mathit{Id}$
1603: are two commuting skew-symmetric matrices.  (Notice that $\hat{u}$ is only
1604: defined on the universal cover $\mathbb{R}^2$.)  The base point $p$ depends of
1605: course on the choice of origin and is only defined up to multiplication by a
1606: complex unit number.  Nevertheless it plays an important role.
1607: 
1608: Consider now the metric induced by $\hat{u}$.  Due to homogeneity, it is a
1609: constant metric on the $( x, y )$-plane.  By doing a simple change in variables,
1610: we may as well assume that the metric is the standard plane metric, in other
1611: words the immersion is isometric.  (Of course that will change the matrices
1612: $\hat{A}$ and $\hat{B}$, but since they are replaced by some real linear
1613: combination of themselves, the properties mentioned above still hold.)
1614: Henceforth we suppose that $\hat{u}$ is an isometric homogeneous Legendrian
1615: immersion of the plane.
1616: 
1617: Up to a unitary rotation in $\mathbb{C}^3$ we may suppose that $\hat{A}$ is
1618: diagonal, and write $\hat{A} = i \,\hbox{diag} ( a_1, a_2, a_3 )$ with real
1619: coefficients $a_1, a_2, a_3$.  We will now consider three cases and show that
1620: only case (i) is possible.
1621: \begin{enumerate}
1622: \item Suppose $B = i \hbox{diag} ( b_1, b_2, b_3 )$ is diagonal.  Then the
1623: surface lies inside the three torus $T = | p_1 | S^1 \times | p_2 | S^1
1624: \times | p_3 | S^1$.  Necessarily it lifts $\pi ( T )$.  Isometry will
1625: constrain the coefficients to be as above.
1626: 
1627: \item One and only one of the off-diagonal coefficients of $B$ is non zero.
1628: We can assume it is $b_{12}$ up to permutation of the coordinates.
1629: Commutation of $\hat{A}$ and $\hat{B}$ forces $a_1 = a_2$, while $a_3 \neq
1630: a_1$, otherwise we would get a contradiction: $A$ cannot be a multiple of
1631: $i\mathit{Id}$.  Let us first look at equations involving $A$.  The
1632: immersion being isometric in $S^5$, $| p | = | A p | = 1$
1633: \[
1634: 1 = | p_1 |^2 + | p_2 |^2 + | p_3 |^2 = a_1^2 ( | p_1 |^2 + | p_2 |^2 ) +
1635: a_3^2 | p_3 |^2 \hspace{0.75em}
1636: \]
1637: but it is also Legendrian, so \[ \omega ( A p, p ) = \langle i A p, p
1638: \rangle_{\mathbb{C}^3} = a_1 ( | p_1 |^2 + | p_2 |^2 ) + a_3 | p_3 |^2 =
1639: 0 .
1640: \]
1641: Hence \[ | p_1 |^2 + | p_2 |^2 = \frac{a_3}{a_3 - a_1} \hspace{0.25em},
1642: \hspace{0.75em} | p_3 |^2 = - \frac{a_1}{a_3 - a_1} \hbox{ and } a_1 a_3
1643: = - 1
1644: \]
1645: excluding thus $a_1 = 0$, and finally \[ | p_1 |^2 + | p_2 |^2 =
1646: \frac{1}{1 + a_1^2} \hspace{0.75em}, \hspace{0.75em} | p_3 |^2 =
1647: \frac{a_1^2}{1 + a_1^2} \hspace{0.75em} .
1648: \]
1649: Take now into account the Legendrian constraints on $B$: \[ B =
1650: \left(\begin{array}{ccc} i b_1 & b_{12} & 0\\
1651: - \overline{b_{12}} & i b_2 & 0\\
1652: 0 & 0 & b_3 \end{array}\right)
1653: \]
1654: \[ 0 = \langle B p, p \rangle_{\mathbb{C}^3} = i \left( \sum_1^3 b_j |
1655: p_j |^2 + 2 \mathrm{Im} ( b_{12} \overline{p_1} p_2 ) \right) \] \[ 0 =
1656: \langle B p, A p \rangle_{\mathbb{C}^3} = \sum_1^3 a_j b_j | p_j |^2 + 2
1657: a_1 \mathrm{Im} ( b_{12} \overline{p_1} p_2 ).
1658: \]
1659: Uniting both, we deduce \[ a_1 b_1 | p_1 |^2 + a_1 b_2 | p_2 |^2 + a_1 b_3
1660: | p_3 |^2 = a_1 b_1 | p_1 |^2 + a_1 b_2 | p_2 |^2 + a_3 b_3 | p_3 |^2.
1661: \]
1662: Since $| p_3 | \neq 0$ and $a_1 \neq a_3$, $b_3$ vanishes.  The remaining
1663: equations are: \[ | p_1 |^2 + | p_2 |^2 = \frac{1}{1 + a_1^2}
1664: \]
1665: \begin{equation}
1666: \label{bmatrix1} b_1 | p_1 |^2 + b_2 | p_2 |^2 + 2 \mathrm{Im} ( b_{12}
1667: \overline{p_1} p_2 ) = 0
1668: \end{equation}
1669: \begin{equation}
1670: \label{bmatrix2} ( b_1^2 + | b_{12} |^2 ) | p_1 |^2 + ( b_2^2 + |
1671: b_{12}|^2 ) | p_2 |^2 + 2 ( b_1 + b_2 ) \mathrm{Im} ( b_{12}
1672: \overline{p_1} p_2 ) = 1.
1673: \end{equation}
1674: Finally we infer a contradiction: indeed equation (\ref{bmatrix1}) amounts to
1675: the existence of an isotropic vector $( p_1, p_2 )$ for the skew-hermitian
1676: matrix $\left( \begin{array}{cc} i b_1 & b_{12}\\
1677: - \overline{b_{12}} & i b_2 \end{array} \right)$, and that requires its
1678: determinant $|b_{12}|^2 - b_1 b_2$ to vanish.  Plugging this into
1679: (\ref{bmatrix2}), we obtain \[ ( b_1 + b_2 ) \big( b_1 | p_1 |^2 + b_2 |
1680: p_2 |^2 + 2 \mathrm{Im} ( b_{12} \overline{p_1} p_2 ) \big) = 1,
1681: \]
1682: obviously contradicting (\ref{bmatrix1}).
1683: 
1684: \item If at least two off-diagonal coefficients of $B$ are non-zero, then $A =
1685: ia_1 \mathit{Id}$.  But that contradicts $( A p | p ) = 0$.  So that case is
1686: also excluded.
1687: \end{enumerate}
1688: \medskip
1689: 
1690: Notice that in the language of integrable systems, homogeneous tori correspond
1691: to vacuum solutions and are of finite type for $p=0$.
1692: \end{proof}
1693: 
1694: \begin{thebibliography}{99}
1695: %
1696: \bibitem{Br}
1697: {\bibname R. Bryant}, `Minimal lagrangian submanifolds of K\"ahler-Einstein
1698: manifolds', Lecture Notes in Math. Vol. 1255, Springer, New York, 1987, 1--12.
1699: 
1700: \bibitem{BP}
1701: {\bibname F. E. Burstall \and F. Pedit}, `Harmonic maps via
1702: Adler--Kostant--Symes theory', \emph{Harmonic maps and integrable systems}, (eds
1703: A. Fordy and J. C. Wood), Aspects of Mathematics E23 (Vieweg 1994); also in
1704: http://www.amsta.leeds.ac.uk/Pure/staff/wood/FordyWood/contents.html~.
1705: 
1706: \bibitem{BFPP}
1707: {\bibname F. Burstall, D. Ferus, F. Pedit \and U. Pinkall}, `Harmonic tori in
1708: symmetric spaces and commuting Hamiltonian systems on loop algebras', \emph{Ann.
1709: of Maths.} (II) 138 (1993), 173--212.
1710: 
1711: %\bibitem{CaUr}
1712: %{\bibname I. Castro \and F. Urbano}, `Examples of unstable Hamiltonian-minimal
1713: %Lagrangian tori in $\C^2$', \emph{Compositio Mathematica} 111 (1998), 1-14.
1714: 
1715: %\bibitem{CM}
1716: %{\bibname B.-Y. Chen \and J.-M. Morvan}, `Deformations of isotropic submanifolds
1717: %in K\"ahler manifolds', \emph{J. of Geometry and Physics} 13 (1994), 79-104.
1718: 
1719: \bibitem{Da}
1720: {\bibname P. Dazard}, `Sur la g\'eom\'etrie des sous-fibr\'es et des feuilletages
1721: lagrangiens', \emph{Ann. Sci. Ecole Norm. Sup}. (4), 14 (1981) 465--480.
1722: 
1723: \bibitem{DPW}
1724: {\bibname J. Dorfmeister, F. Pedit \and Hong-You Wu}, `Weierstrass type
1725: representation of harmonic maps into symmetric spaces', \emph{Comm.  in Analysis
1726: and Geometry} Vol 6, Number 4 (1998), 633-668.
1727: 
1728: \bibitem{G}
1729: {\bibname M. Guest}, {\em Harmonic maps, loop groups and integrable systems},
1730: (Cambridge University Press, Cambridge, 1997).
1731: 
1732: \bibitem{Has}
1733: {\bibname M. Haskins}, `Special Lagrangian Cones', arXiv:math.DG/0005164.
1734: 
1735: \bibitem{H}
1736: {\bibname F. H\'elein}, {\em Harmonic maps, constant mean curvature surfaces and
1737: integrable systems}, Lecture in Mathematics, ETH Z\"urich, (Birkh\"auser, 2001).
1738: 
1739: \bibitem{HR2}
1740: {\bibname F. H\'elein \and P. Romon}, `Weierstrass representation of Lagrangian
1741: surfaces in four-dimensional space using spinors and quaternions',
1742: \emph{Commentarii Mathematici Helvetici} 75 (2000), 668--680.
1743: 
1744: \bibitem{HR1}
1745: {\bibname F. H\'elein \and P. Romon}, `Hamiltonian stationary Lagrangian
1746: surfaces in $\C^2$', \emph{Communications in Analysis Geometry}, Vol.  10, N. 1
1747: (2002), 79--126; arXiv:math.DG/0009202.
1748: 
1749: \bibitem{HR3}
1750: {\bibname F. H\'elein \and P. Romon}, `Hamiltonian stationary Lagrangian
1751: surfaces in Hermitian symmetric spaces', \emph{Differential geometry and
1752: integrable systems (Tokyo, 2000)}, (eds Martin Guest, Reiko Miyaoka, and
1753: Yoshihiro Ohnita), \emph{Contemp.  Math.} 308, (Amer.  Math.  Soc., Providence,
1754: RI, 2002), pp.  161--178.
1755: 
1756: \bibitem{J}
1757: {\bibname D. Joyce}, `Special Lagrangian 3-folds and integrable systems', Proc.
1758: of the 9th Math.  Soc.  Japan Internat.  Res.  Inst.; arXiv:math.DG/0101249.
1759: 
1760: \bibitem{MM}
1761: {\bibname H. Ma \and Y. Ma}, `Totally real minimal tori in $\C P^2$',
1762: arXiv:math.DG/0106141.
1763: 
1764: \bibitem{McI1}
1765: {\bibname I. McIntosh}, `The construction of all non-isotropic harmonic tori in
1766: complex projective space', \emph{Internat.  J. Math.} 6 (1995), 831--879.
1767: 
1768: \bibitem{McI2}
1769: {\bibname I. McIntosh}, 
1770: `Two remarks on the construction of harmonic tori in $\C P^2$', 
1771: \emph{Internat.  J. Math.} 7 (1996), 515--521.
1772: 
1773: \bibitem{McI3}
1774: {\bibname I. McIntosh}, 
1775: `Special Lagrangian cones in $\C^3$ and primitive harmonic maps', 
1776: \emph{J. London Math. Soc., Second Series} 67 (2003), number 3, pp. 769--789;
1777: arXiv:math.DG/0201157.
1778: 
1779: %\bibitem{PrS}
1780: %{\bibname A. Pressley \and G. Segal}, {\em Loop groups}, Oxford Mathematical
1781: %Monographs, (Clarendon Press, Oxford, 1986).
1782: 
1783: \bibitem{Re1}
1784: {\bibname H. Reckziegel}, `Horizontal lifts of isometric immersions into the
1785: bundle space of a pseudo-Riemannian submersion', \emph{Global differential
1786: geometry and global analysis (1984)}, Lecture Notes in Mathematics 1156,
1787: (Springer, 1985), 264-279.
1788: 
1789: \bibitem{Re2}
1790: {\bibname H. Reckziegel}, `A correspondence between horizontal submanifolds of
1791: Sasakian manifolds and totally real submanifolds of K\"ahlerian submanifolds',
1792: \emph{Pap.  Colloq., Hajduszoboszlo/Hung.  1984, Vol.  2}, Colloq.  Math.  Soc.
1793: Janos Bolyai 46, 1063-1081 (1988).
1794: 
1795: \bibitem{ScWo}
1796: {\bibname R. Schoen \and J. Wolfson}, `Minimizing volume among Lagrangian
1797: submanifolds', \emph{Proc.  Sympos.  Pure Math., 65}, Amer.  Math.  Soc.
1798: (1999).
1799: 
1800: \bibitem{Sh}
1801: {\bibname R. Sharipov}, `Minimal tori in the five dimensional sphere in $\C^3$',
1802: \emph{Theor.  and Math.  Phys.} 87 (1991), 363--369.
1803: 
1804: \bibitem{SYZ}
1805: {\bibname A. Strominger, S.-T. Yau \and E. Zaslov}, `Mirror symmetry is
1806: T-duality', \emph{Nucl.  Phys.  B} 479 (1996), 243--259.
1807: 
1808: \bibitem{T}
1809: {\bibname C. L. Terng}, `Geometries and symmetries of soliton equations and
1810: integrable elliptic equations', 
1811: to appear in \emph{Surveys on Geometry and Integrable Systems}, 
1812: Advanced Studies in Pure Mathematics, Mathematical Society of Japan; 
1813: arXiv:math.DG/0212372.
1814: 
1815: \bibitem{U}
1816: {\bibname K. Uhlenbeck}, `Harmonic maps into Lie groups', \emph{Journal of
1817: Differential Geometry} 30 (1989), 1-50.
1818: 
1819: \end{thebibliography}
1820: 
1821: \affiliationone{ Fr\'ed\'eric H\'elein\\
1822: Universit\'e Denis Diderot (Paris 7)\\
1823: Institut de Math\'ematiques de Jussieu -- UMR 7586, Case 7012\\
1824: 2, place Jussieu\\
1825: 75251 Paris Cedex 05\\
1826: France \email{helein@math.jussieu.fr}}
1827: % Important: Do not put any empty line here.
1828: \affiliationtwo{ Pascal Romon\\
1829: Universit\'e de Marne-la-Vall\'ee\\
1830: 5, bd Descartes, Champs-sur-Marne\\
1831: 77454 Marne-la-Vall\'ee Cedex 2 \\
1832: France \email{romon@univ-mlv.fr}}
1833: 
1834: \end{document}
1835: