Extremal Conformal Structures on Projective Surfaces

4 The variational equations

By construction, a conformal structure \([g]\) on the oriented projective surface \((\Sigma,\mathfrak{p})\) is a section of \(Z \to \Sigma\). Here we will show that every conformal structure \([g]\) admits a natural lift \(\widetilde{[g]} : \Sigma \to Y\). In doing so we recover the functional \(\mathcal{E}_{\mathfrak{p}}\) from a different viewpoint, which simplifies the computation of its variational equations. We start with recalling the bundle of complex linear coframes of a Riemann surface.

4.1 The bundle of complex linear coframes

Let \(\Sigma\) be an oriented surface equipped with a conformal structure \([g]\), so that \(\Sigma\) inherits the structure of a Riemann surface whose integrable almost complex structure will be denoted by \(J\). The bundle of complex-linear coframes of \((\Sigma,[g])\) is the \(\mathrm{GL}(1,\mathbb{C})\)-subbundle \(F^+_{[g]}\) of \(F^+\) consisting of those coframes that are complex-linear with respect to \(J\) and the complex structure obtained on \(\mathbb{R}^2\) via the standard identification \(\mathbb{R}^2\simeq \mathbb{C}\). Of course, via the isomorphism \(\mathrm{CO}(2)\simeq \mathrm{GL}(1,\mathbb{C})\), we may equivalently think of \(F^+_{[g]}\) as consisting of those coframes in \(F^{+}\) that are angle preserving with respect to \([g]\) and the standard conformal inner product on \(\mathbb{R}^2\).

Recall that a principal \(\mathrm{CO}(2)\)-connection \(\varphi\) on \(F^+_{[g]}\) is called torsion-free if it satisfies \[\mathrm{d}\omega=-\varphi\wedge\omega,\] where here we think of the tautological \(\mathbb{R}^2\)-valued \(1\)-form \(\omega\) on \(F^+_{[g]}\) as taking values in \(\mathbb{C}\) and the connection taking values in the Lie algebra of \(\mathrm{CO}(2)\simeq\mathrm{GL}(1,\mathbb{C})\), that is, \(\mathbb{C}\). The curvature \(\Phi\) of \(\varphi\) is a \((1,\! 1)\)-form on \(\Sigma\) whose pullback to \(F^{+}_{[g]}\) can be written as \[\mathrm{d}\varphi=R\,\omega\wedge\overline{\omega}\] for some unique complex-valued function \(R\) on \(F^+_{[g]}\). By definition of \(F^{+}_{[g]}\), a complex-valued \(1\)-form on \(\Sigma\) is a \((1,\! 0)\)-form with respect to \(J\) if and only if its pullback to \(F^{+}_{[g]}\) is a complex multiple of \(\omega\). A consequence of this is the following elementary lemma whose proof we omit:

Lemma 4.1

A complex-valued function \(f\) on \(F^{+}_{[g]}\) represents a section of \(K_{\Sigma}^{m}\otimes \overline{K_{\Sigma}^{n}}\) if and only if there exist complex-valued functions \(f^{\prime}\) and \(f^{\prime\prime}\) on \(F^+_{[g]}\) so that \[\mathrm{d}f=f^{\prime}\omega+f^{\prime\prime}\overline{\omega}+fm\varphi+fn\overline{\varphi}.\]

Remark 4.2

Here \(K_{\Sigma}=T^*_{\mathbb{C}}\Sigma^{1,0}\) denotes the canonical bundle of \((\Sigma,J)\), \(K_{\Sigma}^m\) its \(m\)-th tensorial power and \(\overline{K^n_{\Sigma}}\) the conjugate bundle of the \(n\)-th tensorial power of \(K_{\Sigma}\). As usual, we we let \(\nabla_\varphi\) denote the connection induced by \(\varphi\) on \(K_{\Sigma}^{m}\otimes \overline{K_{\Sigma}^{n}}\) and by \(\nabla^{\prime}_\varphi\) its \((1,\! 0)\)-part and by \(\nabla^{\prime\prime}_\varphi\) its \((0,\! 1)\)-part. Of course, if \(s\) is the section of \(K_{\Sigma}^{m}\otimes \overline{K_{\Sigma}^{n}}\) represented by \(f\), then \(\nabla^{\prime}_\varphi s\) is represented by \(f^{\prime}\) and \(\nabla^{\prime\prime}_\varphi s\) is represented by \(f^{\prime\prime}\).

Lemma 4.1 implies that \(\varphi\) may also be thought of as the connection form of the connection induced by \(\varphi\) on \(K_{\Sigma}^{*}\). Therefore, the first Chern class of \(K_{\Sigma}^{*}\) is \[c_1(K_{\Sigma}^{*})=\left[\frac{\mathrm{i}}{2\pi}\Phi\right]\] and hence if \(\Sigma\) is compact, we obtain \[\tag{4.1} \int_{\Sigma}\mathrm{i}\, \Phi=2\pi\chi(\Sigma),\] where \(\chi(\Sigma)\) denotes the Euler-characteristic of \(\Sigma\).

4.2 Submanifold theory in the twistor space

We are interested in co-dimension two submanifolds of \(Z\) arising as images of sections of \(Z \to \Sigma\). The second order theory of such submanifolds is summarised in the following:

Lemma 4.3

Let \([g] : \Sigma \to Z\) be a conformal structure on \((\Sigma,\mathfrak{p})\). Then there exists a lift \(\widetilde{[g]} : \Sigma \to Y\) covering \([g]\) so that the pullback-bundle \(p : P_{[g]}^{\prime}=\widetilde{[g]}^*P\to \Sigma\) is isomorphic to the \(\mathrm{CO}(2)\)-bundle of complex linear coframes \(F^+_{[g]}\) of \((\Sigma,[g])\) and so that on \(P^{\prime}_{[g]}\simeq F^+_{[g]}\) we have \[\zeta_2=2\overline{a}\,\overline{\zeta_1}, \quad \zeta_3=k\zeta_1+2\overline{q}\overline{\zeta_1},\] for unique complex-valued functions \(a,k,q\) on \(P^{\prime}_{[g]}\).

Proof. First recall that in Lemma 3.3 we have defined \[\zeta_1=\theta^1_0+\mathrm{i}\theta^2_0,\quad \zeta_2=\left(\theta^1_1-\theta^2_2\right)+\mathrm{i}\left(\theta^1_2+\theta^2_1\right),\quad \zeta_3=\theta^0_1+\mathrm{i}\theta^0_2,\] where \(\theta=(\theta^i_j)\) is the Cartan connection of \((\Sigma,\mathfrak{p})\).

Let now \([g] : \Sigma \to Z\) be a conformal structure on \((\Sigma,\mathfrak{p})\) and let \(p : P_{[g]}=[g]^*P\to\Sigma\) denote the pullback of the bundle \(\mu : P\to Z\), that is, \[P_{[g]}=\left\{(p,u) \in \Sigma\times P\,|\, [g](p)=\mu(u)\right\}.\] Since \(P_{[g]}\) is \(6\)-dimensional, two of the components of \(\theta\) become linearly dependent when pulled back to \(P_{[g]}\). Clearly, these components must be among the \(1\)-forms that are semibasic for \(\mu\). Recall that these forms are spanned by \(\zeta_1,\zeta_2\) and their complex conjugates. However, since \([g]\) is a section of \(Z \to \Sigma\) and since the \(1\)-forms that are semibasic for the projection \(\pi : P\to \Sigma\) are spanned by \(\zeta_1,\overline{\zeta_1}\), it follows that \(\zeta_1\wedge\overline{\zeta_1}\) is non-vanishing on \(P_{[g]}\). Therefore, on \(P_{[g]}\) we have the relation \[\tag{4.2} \zeta_2=2\overline{a}\overline{\zeta_1}+c\zeta_1\] for unique complex-valued functions \(a,c\). From the equivariance properties of \(\zeta_1,\zeta_2\) under the \(\mathbb{R}_2\rtimes \mathrm{CO}(2)\)-right action (3.6), we obtain that for all \(u \in P_{[g]}\) and \(z \rtimes r\mathrm{e}^{\mathrm{i}\phi}\in \mathbb{R}_2\rtimes \mathrm{CO}(2)\) we have \[c(u\cdot z\rtimes r\mathrm{e}^{\mathrm{i}\phi})=r^3\mathrm{e}^{\mathrm{i}\phi}c(u)+ r^2 z\] and \[\tag{4.3} a(u\cdot z\rtimes r\mathrm{e}^{\mathrm{i}\phi})=r^3\mathrm{e}^{-3\mathrm{i}\phi}a(u).\] It follows that the equation \(c=0\) defines a locus that corresponds to a section \(\widetilde{[g]} : \Sigma \to Y\) covering \([g]\). On the pullback bundle \(P^{\prime}_{[g]}=\widetilde{[g]}^*P\), where \[P^{\prime}_{[g]}=\left\{(p,u) \in \Sigma\times P\,|\, \widetilde{[g]}(p)=\tau(u)\right\},\] we obtain \[\tag{4.4} \zeta_2=2\overline{a}\overline{\zeta_1}.\] Since \(P^{\prime}_{[g]}\) is \(4\)-dimensional, two of the remaining components of \(\theta\) become linearly dependent when pulled back to \(P^{\prime}_{[g]}\). Since the \(1\)-forms that are semibasic for the projection \(\tau : P\to Y\) are spanned by \(\zeta_1,\zeta_2,\zeta_3\) and their complex conjugates, it follows as before that \[\tag{4.5} \zeta_3=k\zeta_1+2\overline{q}\,\overline{\zeta_1}\] for unique complex-valued functions \(k,q\).

Now recall that Cartan’s bundle \(\pi : P\to \Sigma\) is isomorphic to \(F^+\times \mathbb{R}_2\to \Sigma\) equipped with the \(\mathrm{G}\)-right action (3.1). Therefore, \(P_{[g]} \to \Sigma\) is isomorphic to \(F^+_{[g]} \times \mathbb{R}_2 \to \Sigma\) and consequently, the bundle \(P^{\prime}_{[g]} \to \Sigma\) is isomorphic to \(F^+_{[g]} \to \Sigma\).

We also obtain:

Lemma 4.4

The functions \(a,k,q\) and the \(1\)-form \(\varphi\) satisfy the following structure equations on \(P^{\prime}_{[g]}\simeq F^+_{[g]}\) \[\tag{4.6} \mathrm{d}a=a^{\prime}\zeta_1-q \overline{\zeta_1}+2a\varphi-a\overline{\varphi},\] and \[\begin{aligned} \mathrm{d}k&=k^{\prime}\zeta_1+k^{\prime\prime}\overline{\zeta_1}+k\varphi+k\overline{\varphi},\\ \mathrm{d}q&=q^{\prime}\zeta_1+\frac{1}{2}\left(\overline{L}+\overline{k^{\prime\prime}}-2\overline{q}a\right)\overline{\zeta_1}+2q \varphi, \end{aligned}\] and \[\tag{4.7} \mathrm{d}\varphi=\left(|a|^2+\frac{1}{2}k-\overline{k}\right)\zeta_1\wedge\overline{\zeta_1}\] for unique complex-valued functions \(r^{\prime},k^{\prime},k^{\prime\prime}\) and \(q^{\prime}\) on \(P^{\prime}_{[g]}\).

Proof. We will only verify the structure equation for \(a\) as the other structure equations are derived in an entirely analogous fashion. The structure equations (3.11) and (4.4) gives \[\begin{aligned} \mathrm{d}\overline{\zeta_2}&=\mathrm{d}(2a\zeta_1)=2\mathrm{d}a\wedge\zeta_1+a\,\mathrm{d}\zeta_1=-\zeta_1\wedge \mathrm{d}a +2a\left(\zeta_1\wedge \varphi+\frac{1}{2}\overline{\zeta_1}\wedge\zeta_2\right)\\ &=-\overline{\zeta_1}\wedge\overline{\zeta_3}+\overline{\zeta_2}\wedge \overline{\varphi}-\overline{\zeta_2}\wedge \varphi\\ &=2q\zeta_1\wedge\overline{\zeta_1}+2a\zeta_1\wedge\overline{\varphi}-2a\zeta_1\wedge \varphi, \end{aligned}\] where we have used (4.5). Equivalently, we obtain \[0=\left(\mathrm{d}a+q\overline{\zeta_1}-2a \varphi+a\overline{\varphi}\right)\wedge\zeta_1,\] which implies (4.6). Finally, the structure equation (4.7) for \(\varphi\) is an immediate consequence of (3.11), (4.4) and (4.5).

As we will see next, the functions \(a,q,k\) on \(P^{\prime}_{[g]}\) satisfy certain equivariance properties with respect to the \(\mathrm{CO}(2)\)-right action on \(P^{\prime}_{[g]}\) and hence represent sections of complex line bundles associated to \(p : P^{\prime}_{[g]}\to \Sigma\).

Proposition 4.5

The choice of a conformal structure \([g]\) on \((\Sigma,\mathfrak{p})\) determines the following objects:

  • A torsion-free connection \(\varphi\) on the bundle of complex-linear coframes of \((\Sigma,[g])\);

  • A section \(\alpha\) of \(K^2_{\Sigma}\otimes \overline{K_{\Sigma}^{*}}\) that is represented by \(a\).

  • A quadratic differential \(Q\) on \(\Sigma\) that is represented by \(q\);

  • A \((1,\! 1)\)-form \(\kappa\) on \(\Sigma\) that is represented by \(k\).

Moreover, the quadratic differential \(Q\) satisfies \[\tag{4.8} Q=-\nabla^{\prime\prime}_\varphi\alpha.\]

Proof. By construction of \(P^{\prime}_{[g]}\simeq F^+_{[g]}\), a complex-valued \(1\)-form on \(\Sigma\) is a \((1,\! 0)\)-form for the complex structure \(J\) induced by \([g]\) and the orientation if and only if its \(p\)-pullback to \(P^{\prime}_{[g]}\) is a complex multiple of \(\zeta_1\). Since \[\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\zeta_1=\frac{1}{r^3}\mathrm{e}^{\mathrm{i}\phi}\zeta_1\] it follows that the sections of \(K_{\Sigma}^2\) are in one-to-one correspondence with the complex-valued functions \(f\) on \(P^{\prime}_{[g]}\) satisfying \[\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*f=r^3\mathrm{e}^{-\mathrm{i}\phi}r^3\mathrm{e}^{-\mathrm{i}\phi}f=r^6\mathrm{e}^{-2\mathrm{i}\phi} f.\] Likewise, it follows that the sections of \(K_{\Sigma}^2\otimes \overline{K_{\Sigma}^{*}}\) are in one-to-one correspondence with the complex-valued functions \(f\) on \(P^{\prime}_{[g]}\) satisfying \[\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*f=r^3\mathrm{e}^{-\mathrm{i}\phi}r^3\mathrm{e}^{-\mathrm{i}\phi}\overline{r^{-3}\mathrm{e}^{\mathrm{i}\phi}} f=r^3\mathrm{e}^{-3\mathrm{i}\phi}f\] and that the sections of \(K_{\Sigma}\otimes \overline{K_{\Sigma}}\) are in one-to-one correspondence with the complex valued functions \(f\) on \(P^{\prime}_{[g]}\) satisfying \[\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*f=r^3\mathrm{e}^{-\mathrm{i}\phi}r^3\mathrm{e}^{\mathrm{i}\phi}f=r^6f.\] From (4.5) and (3.9) we obtain that for all \(u \in P^{\prime}_{[g]}\) and \(r\mathrm{e}^{\mathrm{i}\phi}\in \mathrm{CO}(2)\) \[\begin{aligned} k(u\cdot r\mathrm{e}^{\mathrm{i}\phi})&=r^6k(u),\\ q(u\cdot r\mathrm{e}^{\mathrm{i}\phi})&=r^6\mathrm{e}^{-2\mathrm{i}\phi}q(u). \end{aligned}\] These equations imply that there exists a unique quadratic differential \(Q\) on \(\Sigma\) that is represented by \(q\) and a unique \((1,\! 1)\)-form \(\kappa\) on \(\Sigma\) that is represented by \(k\). Furthermore, (4.3) implies that there exists a unique section \(\alpha\) of \(K_{\Sigma}^2\otimes \overline{K_{\Sigma}^{*}}\) that is represented by \(a\).

It follows from the properties (ii) and (iii) of the Cartan connection that \(\varphi\) is a connection \(1\)-form on the \(\mathrm{CO}(2)\)-bundle \(P^{\prime}_{[g]} \to \Sigma\). Its pushforward under the bundle isomorphism \(P^{\prime}_{[g]} \to F^+_{[g]}\) is then a \(\mathrm{CO}(2)\)-connection on \(F^+_{[g]}\) which – by abuse of notation – we denote by \(\varphi\) as well. The structure equation (4.7) implies that \(\varphi\) is torsion-free.

Finally, the identity \(Q=-\nabla^{\prime\prime}_\varphi\alpha\) is an immediate consequence of the structure equation (4.6) and Lemma 4.1.

We call a map \(\psi : (M,g) \to (N,h)\) between two pseudo-Riemannian manifolds weakly conformal if \(\psi^*h=f g\) for some smooth function \(f\) on \(M\). Note that we do not require \(f\) to be positive. Two immediate consequences of Proposition 4.5 are:

Corollary 4.6

Let \([g]\) be a conformal structure on \((\Sigma,\mathfrak{p})\). Then the lift \(\widetilde{[g]} : (\Sigma,[g]) \to (Y,h_{\mathfrak{p}})\) is weakly conformal if and only if \(Q\equiv 0\). Furthermore, the image of \([g] : \Sigma \to Z\) is a holomorphic curve if and only if \(\alpha\equiv 0\). In particular, if \([g](\Sigma) \subset Z\) is a holomorphic curve, then \(\widetilde{[g]}(\Sigma)\subset Y\) is a holomorphic contact curve.

Remark 4.7

Here we call a holomorphic curve \(\Sigma\subset Y\) a contact curve if its tangent bundle is contained in the (holomorphic) contact structure of \(Y\).

Proof of Corollary 4.6. By construction, the metric \(h_{\mathfrak{p}}\) has the property that its pullback to \(P\) is \[\tau^*h_{\mathfrak{p}}=\frac{1}{2}\left(\zeta_1\circ \overline{\zeta_3}+\zeta_3\circ\overline{\zeta_1}+\zeta_2\circ\overline{\zeta_2}\right).\] Therefore, from (4.4) and (4.5) it follows that \[\tag{4.9} p^*\left(\widetilde{[g]}^*h_{\mathfrak{p}}\right)=\frac{1}{2}\left(4|a|^2+(k+\overline{k})\right)\zeta_1\circ\overline{\zeta_1}+q\,\zeta_1\circ\zeta_1+\overline{q}\,\overline{\zeta_1}\circ\overline{\zeta_1}.\] Since a complex-valued \(1\)-form on \(\Sigma\) is a \((1,\! 0)\)-form for the complex structure defined by \([g]\) and the orientation if and only if its \(p\)-pullback to \(P^{\prime}_{[g]}\) is a complex multiple of \(\zeta_1\), equation (4.9) implies that \(\widetilde{[g]}^*h_{\mathfrak{p}}\) is weakly conformal to \([g]\) if and only if \(q\) vanishes identically. The first claim follows.

The second part of the claim is an immediate consequence of (4.4) and the characterisation of the complex structures on \(Z,Y\) in terms of \(\zeta_1,\zeta_2,\zeta_3\) and the characterisation of the holomorphic contact structure in terms of \(\zeta_2=0\).

Remark 4.8

Recall that if \(\psi : (\Sigma,[g]) \to (N,h)\) is a map from a Riemann surface into a (pseudo-)Riemannian manifold, then the \((2,\! 0)\)-part of the pulled back metric \(\psi^*h\) is called the Hopf differential of \(\psi\). Therefore (4.9) implies that quadratic differential \(Q\) is the Hopf differential of \(\widetilde{[g]}\).

Proposition 4.5 shows that for every choice of a conformal structure \([g]\) on \(\Sigma\) we obtain a section \(\alpha\) of \(K_{\Sigma}^2\otimes \overline{K_{\Sigma}^{*}}\), as well as a connection \(\varphi\) on the principal \(\mathrm{GL}(1,\mathbb{C})\)-bundle of complex-linear coframes of \((\Sigma,[g])\). Since \(K_{\Sigma}^2\otimes \overline{K_{\Sigma}^{*}}\) is a subbundle of \(T^*_{\mathbb{C}}\Sigma^2\otimes T_{\mathbb{C}}\Sigma\), we may use the canonical real structure of the latter bundle to take the real part of \(\alpha\). Consequently, the real part of \(\alpha\) is a \(1\)-form on \(\Sigma\) with values in \(\mathrm{End}(T\Sigma)\). We have already encountered an endomorphism valued \(1\)-form \(A_{[g]}\) whose properties we discussed in Theorem 2.4. In Corollary 2.6 we have also seen that the choice of a conformal structure \([g]\) on \((\Sigma,\mathfrak{p})\) determines a unique \([g]\)-conformal connection \({}^{[g]}\nabla\) so that \({}^{[g]}\nabla+A_{[g]}\) defines \(\mathfrak{p}\). On the other hand, \(\varphi\) also induces a \([g]\)-conformal connection on \(TM\) which we denote by \(\nabla_{\varphi}\).

Proposition 4.9

We have: \[\tag{4.10} \nabla_{\varphi}={}^{[g]}\nabla,\] \[\tag{4.11} 2\operatorname{Re}(\alpha)=A_{[g]},\] \[\tag{4.12} p^*\left(|A_{[g]}|^2_g\,d\mu_g\right)=2\mathrm{i}|a|^2\zeta_1\wedge\overline{\zeta_1}=-\frac{\mathrm{i}}{2}\zeta_2\wedge\overline{\zeta_2}.\]

Since a \([g]\)-conformal connection \({}^{[g]}\nabla\) has holonomy in \(\mathrm{CO}(2)\), it corresponds to a unique torsion-free principal \(\mathrm{CO}(2)\)-connection \(\varphi\) on \(F^+_{[g]}\), see for instance [5]. Before proving Proposition 4.9 it is helpful to see explicitly how the principal connection \(\varphi\) is constructed from \({}^{[g]}\nabla\). The \([g]\)-conformal connection \({}^{[g]}\nabla\) can be written as \[\tag{4.13} {}^{(g,\beta)}\nabla={}^g\nabla+g\otimes \beta^{\sharp}-\beta\otimes\mathrm{Id}-\mathrm{Id}\otimes \beta,\] where \(g\in [g]\) and \(\beta\) is a \(1\)-form on \(M\) with \(g\)-dual vector field \(\beta^{\sharp}\). Let \(g_{ij}=g_{ji}\) be the unique real-valued functions on \(F^+\) so that \(\upsilon^*g=g_{ij}\omega^i\otimes \omega^j\). Let \(\psi=(\psi^i_j)\) denote the Levi-Civita connection form of \(g\), so that we have the structure equations. \[\begin{aligned} \mathrm{d}\omega^i&=-\psi^i_j\wedge\omega^j,\\ \mathrm{d}g_{ij}&=g_{ik}\psi^k_j+g_{kj}\psi^k_i \end{aligned}\] as well as \[\tag{4.14} \mathrm{d}\psi^i_j+\psi^i_k\wedge\psi^k_j=g_{jk}K_g\omega^i\wedge\omega^k,\] where the real-valued function \(K_g\) on \(F^+\) is (the pullback of) the Gauss curvature of \(g\). Therefore, writing \(\upsilon^*\beta=b_i\omega^i\) for real-valued functions \(b_i\) on \(F^+\), the connection \(1\)-form of (4.13) is \[\eta^i_j=\psi^i_j+\left(b_kg^{ki}g_{jl}-\delta^i_jb_l-\delta^i_lb_j\right)\omega^l,\] where the real-valued functions \(g^{ij}=g^{ji}\) on \(F^+\) satisfy \(g^{ik}g_{kj}=\delta^i_j\). The equivariance properties of the functions \(b_i\) imply that there exist unique real-valued functions \(b_{ij}\) on \(F\) so that \[\tag{4.15} \mathrm{d}b_i=b_j\psi^j_i+b_{ij}\omega^j.\] From the equivariance properties of the functions \(g_{ij}\) it follows that the conditions \(g_{11}= g_{22}\) and \(g_{12}=0\) define a reduction of \(\upsilon : F^+ \to \Sigma\) to the \(\mathrm{CO}(2)\)-subbundle of complex linear coframes of \(F^+_{[g]} \to \Sigma\) of \((\Sigma,[g])\). On \(F^+_{[g]}\) we obtain \[0=\mathrm{d}g_{12}=g_{11}\psi^1_2+g_{12}\psi^2_2+g_{12}\psi^1_1+g_{22}\psi^2_1=g_{11}(\psi^1_2+\psi^2_1)\] and hence \(\psi^2_1=-\psi^1_2\). Likewise, we have \[\begin{aligned} 0&=\mathrm{d}g_{11}-\mathrm{d}g_{22}=2\left(g_{11}\psi^1_1+g_{12}\psi^2_1\right)-2\left(g_{12}\psi^1_2+g_{22}\psi^2_2\right)\\ &=2g_{11}\left(\psi^1_1-\psi^2_2\right) \end{aligned}\] so that \(\psi^1_1=\psi^2_2\). Idenfifying \(\mathbb{R}^2\simeq \mathbb{C}\), we may think of \(\omega=(\omega^i)\) as taking values in \(\mathbb{C}\). If we define \(\varphi:=\frac{1}{2}\left(\eta^1_1+\eta^2_2\right)+\frac{\mathrm{i}}{2}\left(\eta^2_1-\eta^1_2\right)\), we obtain \[\tag{4.16} \varphi=\left(\psi^1_1-b_1\omega^1-b_2\omega^2\right)+\mathrm{i}\left(\psi^2_1+b_2\omega^1-b_1\omega^2\right)\] Using this notation the first structure equation can be written in complex form \[\mathrm{d}\omega=-\varphi\wedge\omega,\] hence \(\varphi\) defines a torsion-free principal \(\mathrm{CO}(2)\)-connection on \(F^+_{[g]}\).

Proof of Proposition 4.9. Without losing generality, we can assume that the projective structure \(\mathfrak{p}\) is defined by \({}^{[g]}\nabla+A_{[g]}\) for some \([g]\)-conformal connection \({}^{[g]}\nabla\) and some \(1\)-form \(A_{[g]}\) having all the properties of Theorem 2.4. Recall (3.3) that the choice of a representative connection \(\nabla \in \mathfrak{p}\) gives an identification \(P\simeq F^+\rtimes\mathbb{R}_2\) of Cartan’s bundle so that the Cartan connection form becomes \[\tag{4.17} \theta=\begin{pmatrix} -\frac{1}{3}\operatorname{tr}\eta-\xi \omega & \mathrm{d}\xi-\xi\eta-(S\omega)^t-\xi\omega\xi\\ \omega & \eta-\frac{1}{3}\mathrm{I}\operatorname{tr}\eta+\omega\xi \end{pmatrix}.\] We will construct Cartan’s connection for the representative connection \[\tag{4.18} {}^{(g,\beta)}\nabla+A_{[g]}={}^g\nabla+g\otimes \beta^{\sharp}-\beta\otimes\mathrm{Id}-\mathrm{Id}\otimes \beta+A_{[g]}.\] Let \(A^i_{jk}\) denote the real-valued functions on \(F^+\) representing \(A_{[g]}\). In particular, we have \[\tag{4.19} A^i_{jk}=A^i_{kj}\quad \text{and} \quad A^l_{il}=0.\] On \(F^+\) the connection form of (4.18) is given by \[\tag{4.20} \eta^i_j=\psi^i_j+\left(b_kg^{ki}g_{jl}-\delta^i_jb_l-\delta^i_lb_j+A^i_{jl}\right)\omega^l,\] By definition, the pullback bundle \(P_{[g]}\) is the subbundle of \(F^+\times \mathbb{R}_2\) defined by the equations \(g_{11}=g_{22}\) and \(g_{12}=0\). Now on \(P_{[g]}\simeq F^+_{[g]}\times \mathbb{R}_2\) we have \(\psi^2_1=-\psi^1_2\) and \(\psi^1_1=\psi^2_2\). Using (4.17), (4.19) and (4.20) we compute \[\begin{aligned} \zeta_2&=(\theta^1_1-\theta^2_2)+\mathrm{i}\left(\theta^1_2+\theta^2_1\right)\\ &=\psi^1_1-\psi^2_2+\left(\xi_1+2A^1_{11}\right)\omega^1+\left(-\xi_2-2A^2_{22}\right)\omega^2\\ &\phantom{=}+\mathrm{i}\left(\psi^1_2+\psi^2_1+\left(\xi_2-2A^2_{22}\right)\omega^1+\left(\xi_1-2A^1_{11}\right)\omega^2\right)\\ &=2\overline{a}\overline{\zeta_1}+c\zeta_1, \end{aligned}\] where \[\tag{4.21} \begin{aligned} a&=A^1_{11}+\mathrm{i}A^2_{22},\\ c&=\xi_1+\mathrm{i}\xi_2 \end{aligned}\] and we have used that on \(F^+_{[g]}\) \[\delta_{il}A^l_{jk}=\delta_{jl}A^l_{ik},\] which follows from Theorem 2.4 (vi). Recall that \(P^{\prime}_{[g]}\) was defined by the equation \(c=0\). Hence on \(P^{\prime}_{[g]}\simeq F^+_{[g]}\) the function \(\xi\) vanishes identically. Using this we compute \[\varphi=-\frac{1}{2}\left(3\theta^0_0+\mathrm{i}\left(\theta^1_2-\theta^2_1\right)\right)=\psi^1_1-b_1\omega^1-b_2\omega^2+\mathrm{i}\left(\psi^2_1+b_2\omega^1-b_1\omega^2\right).\] This is precisely (4.16). It follows that the connection defined by \(\varphi\) is the same as the induced torsion-free connection on \(F^+_{[g]}\) by \({}^{[g]}\nabla\). This proves (4.10).

Suppose \(x=(x^i) : U \to \mathbb{R}^2\) are local orientation preserving \([g]\)-isothermal coordinates on \(\Sigma\) and write \(z=(x^1+\mathrm{i}x^2)\). Applying the exterior derivative to \(z\) we obtain a local section \(\tilde{z} : U \to F^+_{[g]}\) so that \[A_{[g]}=\tilde{z}^*A^i_{jk}\mathrm{d}x^j\otimes \mathrm{d}x^k\otimes \frac{\partial}{\partial x^i}.\] By definition of \(\alpha\) we have \[\alpha=\tilde{z}^*a\,\mathrm{d}z\otimes\mathrm{d}z\otimes \otimes \frac{\partial}{\partial \overline{z}},\] hence (4.11) is an immediate consequence of (4.21).

Finally, in our coordinates we obtain \[|A_{[g]}|_g^2\,d\mu_g=4|a|^2\mathrm{d}x^1\wedge \mathrm{d}x^2,\] so that \(p^*\left(|A_{[g]}|_g^2\,d\mu_g\right)=2\mathrm{i}|a|^2\zeta_1\wedge\overline{\zeta_1}=-\frac{\mathrm{i}}{2}\zeta_2\wedge\overline{\zeta_2}\), as claimed.

Note that \(A_{[g]}\) vanishes identically if and only if \(\alpha\) vanishes identically. Therefore, as an immediate consequence of Proposition 4.9, Corollary 2.6 and Corollary 4.6, we obtain an alternative proof of [37] (see also [36] for a ‘generalisation’ to higher dimensions):

Theorem 4.10

A conformal structure \([g]\) on \((\Sigma,\mathfrak{p})\) is preserved by a conformal connection defining \(\mathfrak{p}\) if and only if the image of \([g] : \Sigma \to Z\) is a holomorphic curve.

Remark 4.11

Locally the bundle \(Z \to \Sigma\) always admits sections having holomorphic image and therefore every torsion-free connection on \(T\Sigma\) is locally projectively equivalent to a conformal connection (see [37] for additional details).

4.3 Derivation of the variational equations

Applying a technique from [4], we compute the variational equations for the functional \(\mathcal{E}_{\mathfrak{p}}\). For a compact domain \(\Omega\subset \Sigma\) and a section \([g] : \Sigma \to Z\) we write \[\mathcal{E}_{\mathfrak{p},\Omega}([g])=\int_{\Omega}|A_{[g]}|^2_gd\mu_g.\]

Definition 4.12

We say \([g]\) is an \(\mathcal{E}_{\mathfrak{p}}\)-critical point or that \([g]\) is extremal for the projective structure \(\mathfrak{p}\) if for every compact \(\Omega\subset \Sigma\) and for every smooth variation \([g]_t : \Sigma \to Z\) with support in \(\Omega\), we have \[\frac{\mathrm{d}}{\mathrm{d}t}\bigg|_{t=0}\mathcal{E}_{\mathfrak{p},\Omega}([g]_t)=0.\]

Using this definition we obtain:

Theorem A

Let \((\Sigma,\mathfrak{p})\) be an oriented projective surface. A conformal structure \([g]\) on \(\Sigma\) is extremal for \(\mathfrak{p}\) if and only if \(\widetilde{[g]} : (\Sigma,[g]) \to (Y,h_{\mathfrak{p}})\) is weakly conformal.

Proof. Let \([g] : \Sigma \to Z\) be a conformal structure and \([g]_t : \Sigma \to Z\) a smooth variation of \([g]\) with support in some compact set \(\Omega\subset \Sigma\) and with \(|t|<\varepsilon\). We consider the submanifold of \(\Sigma\times P\times (-\varepsilon,\varepsilon)\) defined by \[P^{\prime}_{[g]_t}=\left\{(p,u,t_0) \in \Sigma\times P\times (-\varepsilon,\varepsilon)\,|\,(p,u)\in P^{\prime}_{[g]_{t_0}} \right\}\] and denote by \(\iota_{[g]_t} : P^{\prime}_{[g]_t} \to \Sigma \times P\times (-\varepsilon,\varepsilon)\) the inclusion map. On \(\Sigma\times P\times (-\varepsilon,\varepsilon)\) we define the real-valued \(2\)-form \[\mathsf{A}=-\frac{\mathrm{i}}{2}\zeta_2\wedge\overline{\zeta_2},%.-\d t\wedge\left(\partial_t \inc\left(\zeta_2\wedge\ov{\zeta_2}\right)\right)\Bigg],\] where, by abuse of notation, we write \(\zeta_2\) for the pullback of \(\zeta_2\) to \(\Sigma\times P\times (-\varepsilon,\varepsilon)\). Using the structure equations (3.11), we compute \[\tag{4.22} \mathrm{d}\mathsf{A}=\frac{\mathrm{i}}{2}\left(\zeta_1\wedge\zeta_3\wedge\overline{\zeta_2}-\zeta_2\wedge\overline{\zeta_1}\wedge\overline{\zeta_3}\right).\]

Now Proposition 4.9 implies \[f(t_0):=\left.\mathcal{E}_{\mathfrak{p},\Omega}([g]_t)\right|_{t=t_0}=\int_{\Omega}\left.\left(\left(\iota_{[g]_t}\right)^*\mathsf{A}\right)\right|_{t=t_0}.\] Therefore \[f^{\prime}(0)=\int_{\Omega}\left.\left(\mathrm{L}_{\partial_t}(\iota_{[g]_t})^*\mathsf{A}\right)\right|_{t=0}=\int_{\Omega}\left.\left(\partial_t \hspace{3pt}\rule[0.5pt]{2mm}{0.5pt}\rule[0.5pt]{0.5pt}{4.5pt}\hspace{3pt}(\iota_{[g]_t})^*\mathrm{d}\mathsf{A}\right)\right|_{t=0},\] where \(\mathrm{L}_{\partial_t}\) denotes the Lie-derivative with respect to the vector field \(\partial_t\). It follows from the proof of Lemma 4.3 that on \(P^{\prime}_{[g]_t}\) there exist complex-valued functions \(a,k,q,B,C\) such that \[\tag{4.23} \zeta_2=2\overline{a}\overline{\zeta_1}+B\mathrm{d}t\quad \text{and}\quad \zeta_3=k\zeta_1+2\overline{q}\overline{\zeta_1}+C \mathrm{d}t\] where we now write \(\zeta_i\) instead of \((\iota_{[g]_t})^*\zeta_i\). Combining (4.22) with (4.23) gives \[(\iota_{[g]_t})^*\mathrm{d}\mathsf{A}=\mathrm{i}\left(qB+\overline{q} \overline{B}\right) \mathrm{d}t\wedge\zeta_1\wedge\overline{\zeta_1}\] so that \[\tag{4.24} f^{\prime}(0)=\mathrm{i}\int_{\Omega}\left.\left(qB+\overline{q}\overline{B}\right)\zeta_1\wedge\overline{\zeta_1}\right|_{t=0}.\] Recall that \(\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\zeta_2=\mathrm{e}^{2\mathrm{i}\phi}\zeta_2\) and therefore, by definition, the complex-valued function \(B|_{t=0}\) satisfies \[\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\left(B|_{t=0}\right)=\mathrm{e}^{2\mathrm{i}\phi}\left(B|_{t=0}\right).\] Since \(\left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\zeta_1=r^{-3}\mathrm{e}^{\mathrm{i}\phi}\zeta_1\) it follows that \(B|_{t=0}\) represents a section of \(\overline{K_{\Sigma}}\otimes K_{\Sigma}^{*}\) with support in \(\Omega\). Here \(K_{\Sigma}\) denotes the canonical bundle of \(\Sigma\) with respect to the complex structure induced by the orientation and \([g]=[g]_t|_{t=0}\).

It remains to show that every such section in (4.23) with support in \(\Omega\) can be realised via some variation of \([g]\). We fix a representative metric \(g \in [g]\). Let \(g_{ij}=g_{ji}\) be the real-valued functions on Cartan’s bundle \(P\) so that \(\pi^*g=g_{ij}\theta^i_0\otimes \theta^j_0\). In particular, from the equivariance properties (ii) of the Cartan connection \(\theta\) it follows that \[\left(R_{b\rtimes a}\right)^*\begin{pmatrix} g_{11} & g_{12} \\ g_{21} & g_{22}\end{pmatrix}=(\det a)^2a^t\begin{pmatrix} g_{11} & g_{12} \\ g_{21} & g_{22}\end{pmatrix}a.\] Applying property (iii) of the Cartan connection this implies the existence of unique real-valued functions \(g_{ijk}=g_{jik}\) so that \[\tag{4.25} \mathrm{d}g_{ij}=-2g_{ij}\theta^0_0+g_{kj}\theta^k_i+g_{ik}\theta^k_j+g_{ijk}\theta^k_0.\] Consider the following conformally invariant functions \[G=\frac{(g_{11}-g_{22})+2\mathrm{i}g_{12}}{\sqrt{g_{11}g_{22}-(g_{12})^2}}, \quad H=\frac{g_{11}+g_{22}}{\sqrt{g_{11}g_{22}-(g_{12})^2}}.\] Translating (4.25) into complex form gives the following structure equation \[\tag{4.26} \mathrm{d}G=G^{\prime}\zeta_1+G^{\prime\prime}\overline{\zeta_1}+H\zeta_2+\overline{G}\left(\overline{\varphi}-\varphi\right),\] for unique complex-valued functions \(G^{\prime},G^{\prime\prime}\) on \(P\). Clearly, the complex-valued functions \(G^{\prime}\) and \(G^{\prime\prime}\) can be expressed in terms of the functions \(g_{ijk}\), as \(\zeta_1=\theta^1_0+\mathrm{i}\theta^2_0\). In order to verify (4.26) it is thus sufficient to plug in the definitions of the functions \(G,H\), the definitions of the forms \(\zeta_2,\varphi\) and to use \[\mathrm{d}g_{ij}=-2g_{ij}\theta^0_0+g_{kj}\theta^k_i+g_{ik}\theta^k_j \quad \text{mod}\quad \theta^1_0,\theta^2_0.\] While this is somewhat tedious, it is straightforward, so we omit the computation.

Fix a section of \(\overline{K_{\Sigma}}\otimes K_{\Sigma}^{*}\) with respect to \([g]\) having support in \(\Omega\). Such sections are well-known to correspond to endomorphisms of \(T\Sigma\) that are trace-free and symmetric with respect to \([g]\). In particular, on \(P\) there exist real-valued functions \((B^i_j)\) representing the corresponding endomorphism. The functions satisfy \[B^i_i=0 \quad \text{and} \quad g_{ij}B^j_k=g_{kj}B^j_i.\] as well as the equivariance property \[\left(R_{b\rtimes a}\right)^*\begin{pmatrix} B^1_1 & B^1_2 \\ B^2_1 & B^2_2\end{pmatrix}=a^{-1}\begin{pmatrix} B^1_1 & B^1_2 \\ B^2_1 & B^2_2\end{pmatrix}a.\] We define \(B=\frac{1}{2}(B^1_1-B^2_2)+\frac{\mathrm{i}}{2}\left(B^1_2+B^2_1\right)\), then \(B\) satisfies \((R_{z\rtimes r\mathrm{e}^{\mathrm{i}\phi}})^*B=\mathrm{e}^{2\mathrm{i}\phi}B\), hence for sufficiently small \(t\) we may vary \([g]\) by defining \([g]_t\) via the zero-locus of the function \[G_t=G-tB H.\] Consequently, on \[P_{[g]_t}=\left\{(p,u,t_0)\in \Sigma\times P\times (-\varepsilon,\varepsilon)\,|\,(p,u) \in P_{[g]_{t_0}}\right\}\] we get \[\begin{aligned} 0&=\mathrm{d}G_t=\mathrm{d}G-\mathrm{d}tB H-t\mathrm{d}\left(B H\right)\\ &=G^{\prime}\zeta_1+G^{\prime\prime}\overline{\zeta_1}+H\zeta_2+\overline{G}\left(\overline{\varphi}-\varphi\right)-\mathrm{d}tB H-t\mathrm{d}\left(B H\right)\\ &=G^{\prime}\zeta_1+G^{\prime\prime}\overline{\zeta_1}+H\zeta_2+t\overline{B}H\left(\overline{\varphi}-\varphi\right)-\mathrm{d}tB H-t\mathrm{d}\left(B H\right) \end{aligned}\] In particular, if we evaluate this last equation on \(\left.P_{[g]_t}\right|_{t=0}\), we obtain \[0=G^{\prime}\zeta_1+G^{\prime\prime}\overline{\zeta_1}+H\zeta_2-\mathrm{d}t B H\] Since \(H\) is non-vanishing on \(\left.P_{[g]_t}\right|_{t=0}\) we must have \[\zeta_2=-\frac{G^{\prime}}{H}\zeta_1-\frac{G^{\prime\prime}}{H}\overline{\zeta_1}+B\mathrm{d}t.\] Since \(P_{[g]_t}^{\prime}\) arises by reducing \(P_{[g]_t}\), it follows that on \(\left.P^{\prime}_{[g]_t}\right|_{t=0}\) we obtain \[\zeta_2=-\frac{G^{\prime\prime}}{H}\overline{\zeta_1}+B \mathrm{d}t,\] as desired. Finally, we now know that (4.24) must vanish where \(B\) is any complex-valued function representing an arbitrary section of \(\overline{K_{\Sigma}}\otimes K_{\Sigma}^{*}\) with support in \(\Omega\). This is only possible if \(q|_{t=0}\) vanishes identically. Applying Corollary 4.6 proves the claim.

Remark 4.13

Clearly, if \([g](\Sigma) \subset Z\) is a holomorphic curve, then \(\widetilde{[g]} : \Sigma \to Y\) is weakly conformal. Using the structure equations this can be seen as follows. The image \([g](\Sigma)\subset Z\) is a holomorphic curve if and only if \(\alpha\) vanishes identically. However, if \(\alpha\) vanishes identically, then so does \(a\) and hence (4.6) implies that \(q\) vanishes identically as well. Consequently, every projective structure \(\mathfrak{p}\) locally admits a conformal structure \([g]\) so that \(\widetilde{[g]}\) is weakly conformal.

We conclude this section by showing that in the compact case \(\mathcal{E}_{\mathfrak{p}}([g])\) is – up to a topological constant – just the Dirichlet energy of \(\widetilde{[g]} : (\Sigma,[g])\to (Y,h_{\mathfrak{p}})\).

Lemma 4.14

Let \((\Sigma,\mathfrak{p})\) be a compact oriented projective surface. Then for every conformal structure \([g] : \Sigma \to Z\) we have \[\int_{\Sigma}|A_{[g]}|^2_g d\mu_g=2\pi\chi(\Sigma)+\frac{1}{2}\int_{\Sigma}\operatorname{tr}_g \widetilde{[g]}^*h_{\mathfrak{p}}\, d\mu_g,\] where \(\chi(\Sigma)\) denotes the Euler-characteristic of \(\Sigma\).

Proof. Recall from (4.9) that \[p^*\left(\widetilde{[g]}^*h_{\mathfrak{p}}\right)=\frac{1}{2}\left(4|a|^2+(k+\overline{k})\right)\zeta_1\circ\overline{\zeta_1}+q\,\zeta_1\circ\zeta_1+\overline{q}\,\overline{\zeta_1}\circ\overline{\zeta_1}.\] Hence we obtain \[\frac{1}{2}\int_{\Sigma}\operatorname{tr}_g \widetilde{[g]}^*h_{\mathfrak{p}}\, d\mu_g=\frac{1}{2}\int_{\Sigma}\left(4|a|^2+(k+\overline{k})\right)\frac{\mathrm{i}}{2}\zeta_1\wedge\overline{\zeta_1}.%=\int_{\Sigma}|A_{[g]}|^2_gd\mu_g\\ \] Since \[\mathrm{d}\varphi=\left(|a|^2+\frac{1}{2}k-\overline{k}\right)\zeta_1\wedge\overline{\zeta_1},\] we get \[\frac{\mathrm{i}}{2}\left(\mathrm{d}\varphi-\mathrm{d}\overline{\varphi}\right)=\frac{1}{2}\left(4|a|^2-(k+\overline{k})\right)\frac{\mathrm{i}}{2}\zeta_1\wedge\overline{\zeta_1}\] and thus \[\begin{aligned} \frac{1}{2}\int_{\Sigma}\operatorname{tr}_g \widetilde{[g]}^*h_{\mathfrak{p}}\, d\mu_g&=\int_{\Sigma}2\mathrm{i}|a|^2\zeta_1\wedge\overline{\zeta_1}-\int_{\Sigma}\frac{\mathrm{i}}{2}(\mathrm{d}\varphi-\mathrm{d}\overline{\varphi})\\ &=\int_{\Sigma}|A_{[g]}|^2_gd\mu_g-2\pi\chi(\Sigma), \end{aligned}\] where we have used (4.1) and (4.12).

As an obvious consequence of Lemma 4.14 and Theorem 2.4 we have the lower bound:

Theorem B

Let \((\Sigma,\mathfrak{p})\) be a compact oriented projective surface. Then for every conformal structure \([g] : \Sigma \to Z\) we have \[\frac{1}{2}\int_{\Sigma}\operatorname{tr}_g \widetilde{[g]}^*h_{\mathfrak{p}}\, d\mu_g\geqslant -2\pi\chi(\Sigma),\] with equality if and only if \(\mathfrak{p}\) is defined by a \([g]\)-conformal connection.