Minimal Lagrangian Connections on Compact Surfaces

5 Minimal Lagrangian connections

We will next compute the structure equations of a timelike/spacelike minimal Lagrangian connection \(\nabla\). As before, we work on the orthonormal coframe bundle \(F_g\) of the induced metric \(g=\pm\mathrm{Ric}(\nabla)\). The submanifold theory discussed in ยง3 tells us that the second fundamental form of \(o : \Sigma \to (T^*\Sigma,h_{\nabla})\) is described in terms of a cubic differential and a \((1,\! 0)\)-form. Using the definition (4.2) of the \((1,\! 0)\)-form, the expression (4.11) for the second fundamental form and the minimality conditions (4.16), one easily computes that on \(F_g\) the \((1,\! 0)\)-form is represented by the complex-valued function \[b=\pm\frac{3}{8}\left((R_{111}+R_{221})-\mathrm{i}(R_{112}+R_{222})\right).\] Hence comparing with (4.15), we conclude that \(b=b_1-\mathrm{i}b_2\). Note that if we define the \((1,\! 0)\)-form \[\beta^{1,0}:=\beta+\mathrm{i}\star_g\beta,\] then we have \(\upsilon^*\beta^{1,0}=(b_1-\mathrm{i}b_2)(\omega_1+\mathrm{i}\omega_2)\), thus the \((1,\! 0)\)-form obtained from the normal bundle valued quadratic differential \(Q_-\) by using the symplectic form \(\Omega_{\nabla}\) is \(\beta^{1,0}\). Likewise, \(Q_+\) gives the cubic differential \(C\) on \(\Sigma\) which is represented on \(F_g\) by the complex-valued function \[c=\mp\frac{1}{8}\left(\left(R_{111}-3R_{122}\right)+\mathrm{i}\left(-3R_{112}+R_{222} \right)\right).\] From Lemma 2.1 and \[\upsilon^*\left(\mathrm{Sym}_0\mathrm{Ric}(\nabla)\right)=\left(R_{ijk}-\frac{3}{2}\delta_{(ij}R_{k)lm}\delta^{lm}\right)\omega_i\otimes\omega_j\otimes \omega_k\] we easily compute that the cubic differential \(C\) satisfies \(\operatorname{Re}(C)=\mp\frac{1}{2}\mathrm{Sym}_0\nabla g\), where the subscript \(0\) denotes the trace-free part with respect to \(g\).

The structure equations can now be summarised as follows:

Proposition 5.1

Let \(\Sigma\) be an oriented surface and \(\nabla\) a timelike/spacelike minimal Lagrangian connection on \(T\Sigma\). Then we obtain a triple \((g,\beta,C)\) on \(\Sigma\) consisting of a Riemannian metric \(g=\pm\mathrm{Ric}(\nabla)\), a \(1\)-form \(\beta=\frac{3}{8}\operatorname{tr}_g \mathrm{Sym}\nabla g\) and a cubic differential \(C\) so that \(\mathrm{Re}(C)=\mp\frac{1}{2}\mathrm{Sym}_0\nabla g\). Furthermore, the triple \((g,\beta,C)\) satisfies the following equations \[\tag{5.1} K_g=\pm 1+2\,|C|_g^2+\delta_g\beta,\] \[\tag{5.2} \overline{\partial} C=\left(\beta-\mathrm{i}\star_g \beta\right)\otimes C,\] \[\tag{5.3} \mathrm{d}\beta=0.\]

Proof. In our frame adaption where \(S_{ij}=\pm\delta_{ij}\) on \(F_g\), we obtain from (4.13) \[0=\mathrm{d}S_{ij}=\left(R_{ijk}+\frac{2}{3}L_{(i}\varepsilon_{j)k}\right)\omega_k\pm\delta_{ik}\theta^k_j\pm\delta_{kj}\theta^k_i.\] Therefore, writing \(\theta_{ij}=\delta_{ik}\theta^k_j\), we have \[\tag{5.4} \theta_{(ij)}=\mp\frac{1}{2}\left(R_{ijk}+\frac{2}{3}L_{(i}\varepsilon_{j)k}\right)\omega_k.\] For later usage we introduce the notation \(c_1=\mp(\frac{1}{8}R_{111}-\frac{3}{8}R_{122})\) and \(c_2=\mp(-\frac{3}{8}R_{112}+\frac{1}{8}R_{222})\), so that \(c=c_1+\mathrm{i}c_2\). Equation (5.4) written out gives \[\begin{aligned} \theta_{11}&=\mp\frac{1}{2}R_{111}\omega_1\mp\left(\frac{3}{4}R_{112}+\frac{1}{4}R_{222}\right)\omega_2,\\ \frac{1}{2}(\theta_{12}+\theta_{21})&=\mp\left(\frac{3}{8}R_{112}-\frac{1}{8}R_{222}\right)\omega_1\mp\left(\frac{3}{8}R_{122}-\frac{1}{8}R_{111}\right)\omega_2,\\ \theta_{22}&=\mp\left(\frac{3}{4}R_{122}+\frac{1}{4}R_{111}\right)\omega_1\mp\frac{1}{2}R_{222}\omega_2.\\\end{aligned}\] Defining \[\varphi=\theta_{21}\mp\frac{1}{2}R_{222}\omega_1\pm\left(\frac{1}{4}R_{111}+\frac{3}{4}R_{122}\right)\omega_2,\] we compute \[\tag{5.5} \theta=\begin{pmatrix} -\beta & \star_g \beta-\varphi \\ \varphi-\star_g\beta & -\beta\end{pmatrix}+\begin{pmatrix} c_1\omega^1-c_2\omega^2 & -c_2\omega^1-c_1\omega^2\\ -c_2\omega^1-c_1\omega^2 & -c_1\omega^1+c_2 \omega^2\end{pmatrix}.\] The motivation for the definition of \(\varphi\) is that we have \[\mathrm{d}\omega_1=-\omega_2\wedge\varphi\quad \text{and}\quad \mathrm{d}\omega_2=-\varphi\wedge\omega_1,\] hence \(\varphi\) is the Levi-Civita connection form of \(g\). In particular, we see that timelike/spacelike minimal Lagrangian connections are twisted Weyl connections. Since \(\mathrm{Ric}(\nabla)=\pm g\), it follows that the curvature \(2\)-form of \(\theta\) must satisfy \[\tag{5.6} \Theta=\mathrm{d}\theta+\theta\wedge\theta=\begin{pmatrix} 0 & \pm\omega_1\wedge\omega_2 \\ \mp\omega_1\wedge\omega_2 & 0\end{pmatrix}.\] In order to evaluate this condition we first recall that we write \(\upsilon^*\beta=b_i\omega_i\) and \[\begin{aligned} \mathrm{d}b_1&=b_{11}\omega_1+b_{12}\omega_2+b_2\varphi,\\ \mathrm{d}b_2&=b_{21}\omega_1+b_{22}\omega_2-b_1\varphi,\end{aligned}\] for unique real-valued functions \(b_{ij}\) on \(F_g\). From (5.5) and (5.6) we obtain \[\mathrm{d}\beta=-\frac{1}{2}\left(\mathrm{d}\theta_{11}+\mathrm{d}\theta_{22}\right)=\frac{1}{2}\left(\theta_{12}\wedge\theta_{21}+\theta_{21}\wedge\theta_{12}\right)=0\] showing that \(\beta\) is closed, hence (5.3) is verified. Likewise, we also obtain \[\begin{aligned} \mathrm{d}\varphi&=\frac{1}{2}(\mathrm{d}\theta_{21}-\mathrm{d}\theta_{12})+\mathrm{d}\! \star_g\!\beta\\ &=(b_{11}+b_{22})\omega_1\wedge\omega_2+\frac{1}{2}\left((\theta_{11}-\theta_{22})\wedge(\theta_{21}+\theta_{12})\right)\mp \omega_1\wedge\omega_2\\ &=-\left(2\left((c_1)^2+(c_2)^2\right)-(b_{11}+b_{22})\pm 1\right)\omega_1\wedge\omega_2.\end{aligned}\] Writing \(K_g\) for the Gauss curvature of \(g\), this last equation is equivalent to \[K_g=\pm 1+2\,|C|_g^2+\delta_g\beta,\] which verifies (5.1).

In order to prove (5.2), we use \[\upsilon^*\left(\beta-\mathrm{i}\star_g\beta\right)=(b_1+\mathrm{i}b_2)(\omega^1-\mathrm{i}\omega^2).\] In light of (2.6) the condition (5.2) is equivalent to the condition \[\tag{5.7} \mathrm{d}c\wedge\omega=\overline{b}c\overline{\omega}\wedge\omega+3\mathrm{i}c\varphi\wedge\omega,\] where we use the complex notation \(b=b_1-\mathrm{i}b_2\), \(c=c_1+\mathrm{i}c_2\) and \(\omega=\omega_1+\mathrm{i}\omega_2\). Again, from (5.5) we compute \[c\omega=\frac{1}{2}\left[(\theta_{11}-\theta_{22})-\mathrm{i}\left(\theta_{12}+\theta_{21}\right)\right],\] hence \[\begin{gathered} \mathrm{d}c\wedge\omega=\mathrm{d}(c\omega)-c\mathrm{d}\omega=-\theta_{12}\wedge\theta_{21}\\ +\frac{\mathrm{i}}{2}\left(\theta_{11}\wedge(\theta_{12}-\theta_{21})+\theta_{22}\wedge(\theta_{21}-\theta_{12})\right)-(c_1+\mathrm{i}c_2)(\mathrm{d}\omega_1+\mathrm{i}\mathrm{d}\omega_2).\end{gathered}\] Using (5.5) and the structure equations (2.5) this gives \[\begin{gathered} \mathrm{d}c\wedge\omega=3c_2\omega_1\wedge\varphi+3c_1\omega_2\wedge\varphi-2(b_1c_2+b_2c_1)\omega_1\wedge\omega_2\\ +\mathrm{i}\left(-3c_1\omega_1\wedge\varphi+3c_2\omega_2\wedge\varphi+2(b_1c_1-b_2c_2)\omega_1\wedge\omega_2\right),\end{gathered}\] which is equivalent to \[\begin{gathered} \mathrm{d}c\wedge\omega=(b_1+\mathrm{i}b_2)(c_1+\mathrm{i}c_2)(\omega_1-\mathrm{i}\omega_2)\wedge(\omega_1+\mathrm{i}\omega_2)\\ +3\mathrm{i}(c_1+\mathrm{i}c_2)\varphi\wedge(\omega_1+\mathrm{i}\omega_2),\end{gathered}\] that is, equation (5.7). This completes the proof.

Conversely, unravelling our computations backwards, we also get:

Proposition 5.2

Suppose a triple \((g,\beta,C)\) on an oriented surface \(\Sigma\) satisfies the equations (5.1),(5.2),(5.3). Then the connection form (5.5) on \(F_g\) defines a timelike/spacelike minimal Lagrangian connection \(\nabla\) on \(T\Sigma\) with \(\mathrm{Ric}(\nabla)=\pm g\).

We immediately obtain:

Corollary 5.3

Let \(\Sigma\) be an oriented surface. Then there exists a one-to-one correspondence between timelike/spacelike minimal Lagrangian connections on \(T\Sigma\) and triples \((g,\beta,C)\) satisfying (5.1),(5.2),(5.3).

Proof. Clearly, the map sending a torsion-free minimal Lagrangian connection \(\nabla\) into the set of triples \((g,\beta,C)\) satisfying the above structure equations, is surjective. Now suppose the two triples \((g_1,\beta_1,C_1)\) and \((g_2,\beta_2,C_2)\) on \(\Sigma\) satisfy the above structure equations and define the same torsion-free spacelike minimal Lagrangian connection \(\nabla\) on \(T\Sigma\). Then \(g_1=\pm \mathrm{Ric}(\nabla)=g_2\) and consequently we obtain \(\beta_1=\beta_2\) as well as \(C_1=C_2\), since these quantities are defined in terms of \(\nabla \mathrm{Ric}(\nabla)\) by using the metric \(g_1=g_2\).

Remark 5.4

Remark 4.10 immediately implies that a minimal Lagrangian connection is projectively flat if and only if the cubic differential \(C\) is holomorphic.

Another consequence of the structure equation is:

Proposition 5.5

Let \(\nabla\) be a timelike/spacelike Lagrangian connection that is totally geodesic. Then \(\nabla\) is the Levi-Civita connection of a metric \(g\) of Gauss curvature \(K_g=\pm 1\).

Proof. The Lagrangian connection \(\nabla\) is totally geodesic if and only if the second fundamental form vanishes identically or equivalently, if \(\beta\) and \(C\) vanish identically. In this case (5.5) implies that \(\theta\) is the Levi-Civita connection of \(g\) and the structure equation (5.1) gives that \(g\) has Gauss curvature \(\pm 1\).

Remark 5.6

As we have mentioned previously in Remark 3.6, a twisted Weyl connection on \((\Sigma,[g])\) defines an AH structure \((\mathfrak{p}(\nabla),[g])\). Moreover, a twisted Weyl connection arising from a triple \((g,\beta,C)\) satisfying \(\overline{\partial} C=(\beta-\mathrm{i}\star_g\beta)\otimes C\) defines an associated AH structure \((\mathfrak{p}(\nabla),[g])\) which is naive Einstein in the terminology of [16, 17, 18]. Therefore, every minimal Lagrangian connection defines a naive Einstein AH structure.