Extremal Conformal Structures on Projective Surfaces

3 Projective surfaces and associated bundles

A natural case to consider is \(n=2\), where \(\mathcal{F}\) is just the square of the \(L^2\)-norm of \(A_{[g]}\) taken with respect to \([g]\). We will henceforth consider the surface case only.

There are several natural geometric spaces fibering over an oriented projective surface which we will discuss next. Before doing so, we recall a result of Cartan [9], which canonically associates a principal bundle together with a “connection” to every projective manifold. The reader interested in a description of Cartan’s construction using modern language may also consult [26]. For additional background on Cartan geometries the reader may also consult [8].

3.1 Cartan’s normal projective connection

Let \(\Sigma\) be an oriented surface and let \(\mathrm{G}\simeq \mathbb{R}_2\rtimes \mathrm{GL}^+(2,\mathbb{R})\) denote the two-dimensional orientation preserving affine group which we think of as the subgroup of \(\mathrm{SL}(3,\mathbb{R})\) consisting of matrices of the form \[b\rtimes a=\begin{pmatrix} \det a^{-1} & b \\ 0 & a\end{pmatrix},\] for \(b \in \mathbb{R}_2\) and \(a \in \mathrm{GL}^+(2,\mathbb{R})\). We denote by \(\upsilon : F^+\to \Sigma\) the principal right \(\mathrm{GL}^+(2,\mathbb{R})\)-bundle of coframes that are orientation preserving with respect to the chosen orientation on \(\Sigma\) and the standard orientation on \(\mathbb{R}^2\). We define a right \(\mathrm{G}\)-action on \(F^+\times \mathbb{R}_2\) by the rule \[\tag{3.1} (u,\xi)\cdot (b\rtimes a)=\left(\det a^{-1} a^{-1} \circ u,\xi a \det a + b\det a\right),\] for all \(b\rtimes a \in \mathrm{G}\). Here \(\xi : F^+ \times \mathbb{R}_2 \to \mathbb{R}_2\) denotes the projection onto the latter factor. This action turns \(\pi : F^+\times \mathbb{R}_2 \to \Sigma\) into a principal right \(\mathrm{G}\)-bundle over \(\Sigma\), where \(\pi : F^+\times \mathbb{R}_2\to \Sigma\) denotes the natural basepoint projection. Suppose \(\nabla\) is a torsion-free connection on \(T\Sigma\) with connection \(1\)-form \(\eta=(\eta^i_j)\) on \(F^+\) so that we have the structure equations2 \[\begin{aligned} \mathrm{d}\omega^i&=-\eta^i_j\wedge\omega^j,\\ \mathrm{d}\eta^i_j&=-\eta^i_k\wedge\eta^k_j+(\delta^i_{[k}S_{l]j}-S_{[kl]}\delta^i_j)\omega^k\wedge\omega^l, \end{aligned}\] where \(S=(S_{ij})\) represents the projective Schouten tensor \(\mathrm{Schout}(\nabla)\) of \(\nabla\) and \(\omega^i\) the components of the tautological \(\mathbb{R}^n\)-valued \(1\)-form \(\omega\) on \(F\). Recall that the Schouten tensor is defined as \[\tag{3.2} \mathrm{Schout}(\nabla)=\mathrm{Ric}^{+}(\nabla)-\frac{1}{3}\mathrm{Ric}^{-}(\nabla),\] where \(\mathrm{Ric}^{\pm}(\nabla)\) denote the symmetric and anti-symmetric part of the Ricci curvature of \(\nabla\). On \(P=F^+\times \mathbb{R}_2\) we define the \(\mathfrak{sl}(3,\mathbb{R})\)-valued \(1\)-form \[\tag{3.3} \theta=\begin{pmatrix} -\frac{1}{3}\operatorname{tr}\eta-\xi \omega & \mathrm{d}\xi-\xi\eta-(S\omega)^t-\xi\omega\xi\\ \omega & \eta-\frac{1}{3}\mathrm{I}\operatorname{tr}\eta+\omega\xi \end{pmatrix}.\] The reader may check that the pair \((\pi : P\to \Sigma,\theta)\) defines a Cartan geometry of type \((\mathrm{SL}(3,\mathbb{R}),\mathrm{G})\), that is, \(\pi : P\to \Sigma\) is a principal right \(\mathrm{G}\)-bundle and \(\theta\) is an \(\mathfrak{sl}(3,\mathbb{R})\)-valued \(1\)-form on \(P\) satisfying the following properties:

Moreover, writing \(\theta=(\theta^i_j)_{i,j=0,1,2}\), the Cartan geometry \((\pi : P\to \Sigma,\theta)\) also satisfies:

Remark 3.1

Cartan’s bundle is unique in the following sense: If \((\hat{\pi} : \hat{P} \to \Sigma,\hat{\theta})\) is another Cartan geometry of type \(\left(\mathrm{SL}(3,\mathbb{R}),\mathrm{G}\right)\) so that the properties (iv),(v) and (vi) hold, then there exists a \(\mathrm{G}\)-bundle isomorphism \(\psi : P\to \hat{P}\) satisfying \(\psi^*\hat{\theta}=\theta\).

A projective structure \(\mathfrak{p}\) on \(\Sigma\) is called flat if every point \(p \in \Sigma\) has a neighbourhood \(U_p\) which is diffeomorphic to a subset of \(\mathbb{RP}^2\) in such a way that the geodesics of \(\mathfrak{p}\) contained in \(U_p\) are mapped onto (segments) of projective lines \(\mathbb{RP}^1\subset \mathbb{RP}^2\). Furthermore, a torsion-free connection \(\nabla\) on \(T\Sigma\) is called projectively flat if \(\mathfrak{p}(\nabla)\) is flat. Using Cartan’s connection, one can show that a projective structure \(\mathfrak{p}\) is flat if and only if the functions \(L_1\) and \(L_2\) vanish identically. Another consequence of Cartan’s result is that there exists a unique \(1\)-form \(\lambda\in\Omega^1(\Sigma,\Lambda^2(T^*\Sigma))\) so that \[\pi^*\lambda=(L_1\theta^1_0+L_2\theta^2_0)\otimes \theta^1_0\wedge\theta^2_0.\] The \(1\)-form \(\lambda\) was first discovered by R. Liouville [32], hence we call \(\lambda\) the Liouville curvature of \(\mathfrak{p}\). In particular, the Liouville curvature is the complete obstruction to flatness of a two-dimensional projective structure.

Example 3.2

Note that the left action of \(\mathrm{SL}(3,\mathbb{R})\) on \(\mathbb{R}^3\) by matrix multiplication descends to define a transitive left action on the projective \(2\)-sphere \(\mathbb{S}^2\). The stabiliser subgroup of the element \([(1\;0\;0)^t]\) is the group \(\mathrm{G}\subset \mathrm{SL}(3,\mathbb{R})\) so that \(\mathbb{S}^2\simeq \mathrm{SL}(3,\mathbb{R})/\mathrm{G}\). Taking \(\theta\) to be the Maurer-Cartan form of \(\mathrm{SL}(3,\mathbb{R})\), the pair \((\pi : \mathrm{SL}(3,\mathbb{R})\to \mathbb{S}^2,\theta)\) is a Cartan geometry of type \((\mathrm{SL}(3,\mathbb{R}),\mathrm{G})\) defining an orientation and projective structure \(\mathfrak{p}_{\mathrm{can}}\) on \(\mathbb{S}^2\) whose geodesics are the “great circles”. Since \(\mathrm{d}\theta+\theta\wedge\theta=0\), this projective structure is flat. We call \(\mathfrak{p}_{\mathrm{can}}\) the canonical flat projective structure on \(\mathbb{S}^2\).

3.2 The twistor space

Inspired by Hitchin’s twistorial description of holomorphic projective structures on complex surfaces [19], it was shown in [13, 42] how to construct a “twistor space” for smooth projective structures. For what follows it will be convenient to construct the twistor space in the smooth category by using the Cartan geometry of a projective surface.

Let therefore \((\Sigma,\mathfrak{p})\) be an oriented projective surface with Cartan geometry \((\pi : P\to \Sigma,\theta)\). By construction, the quotient of \(P\) by the normal subgroup \(\mathbb{R}_2\rtimes\mathrm{\{Id\}}\subset \mathrm{G}\) is isomorphic to the bundle \(\upsilon : F^+ \to \Sigma\) of orientation preserving coframes of \(\Sigma\). In particular, the choice of a conformal structure \([g]\) on \(\Sigma\) corresponds to a section of the fibre bundle \(\mathrm{C}(\Sigma)\simeq P/\left(\mathbb{R}_2\rtimes\mathrm{CO}(2)\right) \to \Sigma\). Here \(\mathrm{CO}(2)=\mathbb{R}^+\times\mathrm{SO}(2)\) is the linear orientation preserving conformal group. By construction, the typical fibre of the bundle \(\mathrm{C}(\Sigma) \to \Sigma\) is diffeomorphic to \(\mathrm{GL}^+(2,\mathbb{R})/\mathrm{CO}(2)\simeq\mathrm{SL}(2,\mathbb{R})/\mathrm{SO}(2)\), that is, the open unit disk \(D^2\subset \mathbb{C}\).

We write the elements of the group \(\mathbb{R}_2\rtimes \mathrm{CO}(2)\) in the following form \[z\rtimes r\mathrm{e}^{\mathrm{i}\phi}=\begin{pmatrix} r^{-2} & \operatorname{Re}(z) & \operatorname{Im}(z)\\ 0& r\cos\phi & r \sin\phi \\ 0 & -r \sin\phi & r\cos\phi\end{pmatrix}, \quad z \in \mathbb{C},r\mathrm{e}^{\mathrm{i}\phi}\in \mathbb{C}^*.\] Property (iii) of the Cartan geometry \((\pi : P\to \Sigma,\theta)\) implies that the (real – or complex-valued) \(1\)-forms on \(P\) that are semibasic3 for the quotient projection \(\mu : P\to \mathrm{C}(\Sigma)\) are complex linear combinations of the complex-valued \(1\)-forms \[\tag{3.5} \zeta_1=\theta^1_0+\mathrm{i}\theta^2_0, \qquad \zeta_2=\left(\theta^1_1-\theta^2_2\right)+\mathrm{i}\left(\theta^1_2+\theta^2_1\right)\] and their complex conjugates. The equivariance property (ii) of the Cartan geometry gives\[\tag{3.6} \left(R_{z\rtimes r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\begin{pmatrix}\zeta_1 \\ \zeta_2\end{pmatrix}=\begin{pmatrix} \frac{1}{r^3} \mathrm{e}^{\mathrm{i}\phi} & 0\\ \frac{z}{r}\mathrm{e}^{\mathrm{i}\phi} & \mathrm{e}^{2\mathrm{i}\phi} \end{pmatrix}\begin{pmatrix}\zeta_1 \\ \zeta_2\end{pmatrix}.\] It follows that there exists a unique almost complex structure \(\mathfrak{J}\) on \(\mathrm{C}(\Sigma)\) having the property that a complex-valued \(1\)-form on \(P\) is the pullback of a \((1,\! 0)\)-form on \(\mathrm{C}(\Sigma)\) if and only if it is a complex linear combination of \(\zeta_1\) and \(\zeta_2\). Indeed, locally we may use a section \(s\) of the bundle \(\mu : P\to \mathrm{C}(\Sigma)\) to pull down the forms \(\zeta_1,\zeta_2\) onto the domain of definition \(U\subset \mathrm{C}(\Sigma)\) of \(s\). Since \(\zeta_1,\zeta_2\) are semi-basic for the projection \(\mu : P\to \mathrm{C}(\Sigma)\), it follows that the pulled down forms are linearly independent over \(\mathbb{C}\) at each point of \(U\). Hence we obtain a unique almost complex structure \(\mathfrak{J}\) on \(U\) whose \((1,\! 0)\)-forms are \(s^*\zeta_1,s^*\zeta_2\). The equivariance (3.6) implies that \(\mathfrak{J}\) is independent of the choice of the section \(s\) and extends to all of \(\mathrm{C}(\Sigma)\). Using property (vi) of the Cartan geometry the reader may easily verify that \[\mathrm{d}\zeta_1=\mathrm{d}\zeta_2=0, \quad \text{mod}\quad \zeta_1,\zeta_2.\] It follows from the Newlander-Nirenberg theorem that \(\mathfrak{J}\) is integrable, thus giving \(\mathrm{C}(\Sigma)\) the structure of a complex surface which we will denote by \(Z\) and which – abusing language – we call the twistor space of the projective surface \((\Sigma,\mathfrak{p})\).

3.3 An indefinite Kähler-Einstein 3-fold

From (3.6) it follows that the holomorphic cotangent bundle \(T^*_{\mathbb{C}}Z^{1,0}\to Z\) is the bundle associated to \(\mu : P\to Z\) via the complex two-dimensional representation \(\rho : \mathbb{R}_2\rtimes \mathrm{CO}(2) \to \mathrm{GL}(2,\mathbb{C})\) defined by the rule \[\tag{3.7} \rho(z \rtimes r\mathrm{e}^{\mathrm{i}\phi})(w_1\;w_2)=(w_1\;w_2)\begin{pmatrix} \frac{1}{r^3} \mathrm{e}^{\mathrm{i}\phi} & 0\\ \frac{z}{r}\mathrm{e}^{\mathrm{i}\phi} & \mathrm{e}^{2\mathrm{i}\phi} \end{pmatrix}\] for all \((w_1\;w_2) \in \mathbb{C}_2\). In particular, the form \(\zeta_1\) is well defined on \(Z\) up to complex-scale and hence may be thought of as a section of the projective holomorphic cotangent bundle \(\mathbb{P}(T^*_{\mathbb{C}}Z^{1,0}) \to Z\). Abusing notation, we write \(\zeta_1(Z) \subset \mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\) to denote the image of \(Z\) under this section. We now have:

Lemma 3.3

There exists a unique integrable almost complex structure on the quotient \(P/\mathrm{CO}(2)\) having the property that its \((1,\! 0)\)-forms pull back to \(P\) to become linear combinations of the forms \[\tag{3.8} \zeta_1=\theta^1_0+\mathrm{i}\theta^2_0,\quad \zeta_2=\left(\theta^1_1-\theta^2_2\right)+\mathrm{i}\left(\theta^1_2+\theta^2_1\right),\quad \zeta_3=\theta^0_1+\mathrm{i}\theta^0_2.\] Furthermore, with respect to this complex structure \(P/\mathrm{CO}(2)\) is biholomorphic to \(Y=\mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\setminus \zeta_1(Z)\) in such a way that the standard holomorphic contact structure on \(Y\) is identified with the subbundle of \(T_{\mathbb{C}}(P/\mathrm{CO}(2))^{1,0}\) defined by the equation \(\zeta_2=0\).

Proof. Again, it follows from the property (iii) of the Cartan connection \(\theta\) that the \(1\)-forms that are semibasic for the quotient projection \(\tau :P\to P/\mathrm{CO}(2)\) are linear combinations of the forms \(\zeta_1,\zeta_2,\zeta_3\) and their complex conjugates. Here \(\mathrm{CO}(2)\subset \mathrm{G}\) is the subgroup consisting of elements of the form \(0\rtimes r\mathrm{e}^{\mathrm{i}\phi}\). Writing \(r\mathrm{e}^{\mathrm{i}\phi}\) instead of \(0\rtimes r\mathrm{e}^{\mathrm{i}\phi}\) and \(\zeta=(\zeta_i)\), we compute from the equivariance property (ii) of \(\theta\) that we have \[\tag{3.9} \left(R_{r\mathrm{e}^{\mathrm{i}\phi}}\right)^*\begin{pmatrix}\zeta_1\\ \zeta_2\\ \zeta_3\end{pmatrix}=\begin{pmatrix}\frac{1}{r^3}\mathrm{e}^{\mathrm{i}\phi} & 0 & 0\\ 0 & \mathrm{e}^{2\mathrm{i}\phi} & 0 \\ 0 & 0 & r^3\mathrm{e}^{\mathrm{i}\phi}\end{pmatrix}\begin{pmatrix}\zeta_1\\ \zeta_2\\ \zeta_3\end{pmatrix}.\] It follows as before that there exists a unique almost complex structure \(\mathfrak{J}\) on the quotient \(P/\mathrm{CO}(2)\) having the property that its \((1,\! 0)\)-forms pull back to \(P\) to become linear combinations of the forms \(\zeta_1,\zeta_2,\zeta_3\). Suppose there exists a \(1\)-form \(\gamma=(\gamma_{ij})\) on \(P\) with values in \(\mathfrak{gl}(3,\mathbb{C})\), so that \(\mathrm{d}\zeta=-\gamma\wedge\zeta\), then it follows again from the Newlander–Nirenberg theorem that \(\mathfrak{J}\) is integrable. Clearly, if such a \(\gamma\) exists, then it is not unique. Defining \(\hat{\gamma}=(\hat{\gamma}_{ij})\), with \(\hat{\gamma}_{ij}=\gamma_{ij}+T_{ijk}\zeta_k\) for some complex-valued functions satisfying \(T_{ijk}=T_{ikj}\) on \(P\) will also work. We can exploit this freedom and make \(\gamma\) take values in the Lie algebra \[\mathfrak{u}(2,1)=\left\{\begin{pmatrix} w_1 & -\overline{w_2} & \mathrm{i}x_1\\ -w_3 & \mathrm{i}x_2 & w_2 \\ \mathrm{i}x_3 & \overline{w_3} & -\overline{w_1} \end{pmatrix}\,: \, w_1,w_2,w_3 \in \mathbb{C}\;\text{and}\; x_1,x_2,x_3 \in \mathbb{R}\right\}\] of the indefinite unitary group \(\mathrm{U}(2,\! 1)\), where the model of \(\mathrm{U}(2,\! 1)\) being used is the subgroup of \(\mathrm{GL}(3,\mathbb{C})\) that fixes the Hermitian form in \(3\)-variables \[H(z)=z_1\overline{z_3}+z_3\overline{z_1}+z_2\overline{z_2}.\] Indeed, writing \[\tag{3.10} L=-\frac{1}{2}\left(L_2-\mathrm{i}L_1\right)\quad\text{and}\quad \varphi=-\frac{1}{2}\left(3\theta^0_0+\mathrm{i}(\theta^1_2-\theta^2_1)\right),\] we have \[\tag{3.11} \mathrm{d}\zeta=-\gamma\wedge\zeta,\] where \[\gamma=\begin{pmatrix} \varphi & -\frac{1}{2}\overline{\zeta_1} & 0 \\-\frac{1}{2}\zeta_3 & \varphi-\overline{\varphi} & \frac{1}{2}\zeta_1\\ L\overline{\zeta_1}-\overline{L}\zeta_1 &\frac{1}{2}\overline{\zeta_3} &-\overline{\varphi}\end{pmatrix},\] as the reader can verify by using the definitions (3.8),(3.10) and the structure equations (3.4). It follows that \(\mathfrak{J}\) is integrable. Likewise, the reader may verify that \[\tag{3.12} \mathrm{d}\varphi=\frac{1}{2}\zeta_3\wedge\overline{\zeta_1}-\frac{1}{4}\zeta_2\wedge\overline{\zeta_2}-\zeta_1\wedge\overline{\zeta_3},\] simply by plugging in the definitions of the involved forms and by using the structure equations (3.4).

Now consider the map \[\tilde{\psi} : P\to P\times \mathbb{C}_2\setminus\{0\}, \quad u \mapsto \left(u,\begin{pmatrix} 0 & 1\end{pmatrix}\right)\] and let \(q : P\times \mathbb{C}_2\setminus\{0\} \to \mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\) denote the natural quotient projection induced by (the projectivisation of) \(\rho\). Then \(q \circ \tilde{\psi} : P\to \mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\) is a submersion onto \(Y\) whose fibres are the \(\mathrm{CO}(2)\)-orbits. Indeed, let \((u,w)\) be a representative of an element \([\nu]\in\mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\) which lies in the complement of \(\zeta_1(Z)\). Then using (3.7) it follows that we might transform with the affine part of the right action of \(\mathbb{R}_2\rtimes \mathrm{CO}(2)\) to ensure that \(w\) is of the form \((0\; w_2)\) for some non-zero complex number \(w_2\). It follows that the element \(u \in P\) is mapped onto \([\nu]\) showing that \(q \circ \tilde{\psi}\) is surjective onto \(Y\). Clearly \(q \circ \tilde{\psi}\) is smooth and a submersion. Furthermore, suppose the two points \(u,u^{\prime} \in P\) are mapped to the same element of \(Y\). Then, there exists an element \(z\rtimes r\mathrm{e}^{\mathrm{i}\phi}\in \mathbb{R}_2\rtimes \mathrm{CO}(2)\) and a non-zero complex number \(s\) so that \[\rho\left((z\rtimes r\mathrm{e}^{\mathrm{i}\phi})^{-1}\right)\begin{pmatrix}0 & 1\end{pmatrix}=\begin{pmatrix}-zr^2\mathrm{e}^{-2\mathrm{i}\phi} & \mathrm{e}^{-2\mathrm{i}\phi}\end{pmatrix}=\begin{pmatrix} 0 & s\end{pmatrix}\] which holds true if and only if \(z=0\). Consequently, there exists a unique diffeomorphism \(\psi : P/\mathrm{CO}(2) \to Y\) making the following diagram commute: \[\begin{CD} P @>\tilde{\psi}>> P\times \mathbb{C}_{2}\setminus\{0\}\\ @VV\tau V @VVqV\\ P/\mathrm{CO}(2) @>\psi >> Y \end{CD}\] The complex structure on \(Y \subset \mathbb{P}(T^*_{\mathbb{C}}Z^{1,0})\) is such that its \((1,\! 0)\)-forms pull back to \(P\times \mathbb{C}_2\setminus\{0\}\) to become linear combinations of the complex-valued \(1\)-forms \(\zeta_1,\zeta_2,\mathrm{d}w_1, \mathrm{d}w_2\), where \(w=(w_1 \; w_2) : P\times \mathbb{C}_2 \to \mathbb{C}_2\) denotes the projection onto the linear factor. Clearly, these forms pull back under \(\tilde{\psi}\) to become linear combinations of the forms \(\zeta_1,\zeta_2,\zeta_3\), hence \(\psi\) is a biholomorphism.

Finally, note that the complex version of the Liouville \(1\)-form on \(T^*_{\mathbb{C}}Z^{1,0}\) – whose kernel defines the canonical contact structure on \(\mathbb{P}(T^{*}_{\mathbb{C}}Z^{1,0})\) – pulls back to \(P\times \mathbb{C}_2\) to become \(w_1\zeta_1+w_2\zeta_2\). Since \[\tilde{\psi}^*\left(w_1\zeta_1+w_2\zeta_2\right)=\zeta_2,\] the claim follows.

Remark 3.4

Whereas the definition of the forms \(\zeta_i\) is a natural consequence of the Lie algebra structure of \(\mathrm{CO}(2)\subset \mathbb{R}_2\rtimes \mathrm{GL}^+(2,\mathbb{R})\), the definition of the form \(\varphi\) in (3.10) is somewhat mysterious at this point. The choice will be clarified during the proof of Proposition 4.9 below.

We will henceforth identify \(Y\simeq P/\mathrm{CO}(2)\) and think of \(\tau\) as the projection map onto \(Y\). Denoting the integrable almost complex structure on \(Y\) by \(J\), the first part of the following proposition is therefore clear:

Proposition 3.5

There exists a unique indefinite Kähler structure on \((Y,J)\) whose Kähler-form \(\Omega_Y\) satisfies \[\begin{aligned} \tau^*\Omega_Y&=-\frac{\mathrm{i}}{4}\left(\zeta_1\wedge \overline{\zeta_3}+\zeta_3\wedge\overline{\zeta_1}+\zeta_2\wedge\overline{\zeta_2}\right). \end{aligned}\] Moreover, the indefinite Kähler metric \(h_{\mathfrak{p}}(\cdot,\cdot):=\Omega_Y(J\cdot,\cdot)\) is Einstein with non-zero scalar curvature.

Proof. The first part of the statement is an immediate consequence of the fact that \(\gamma\) takes values in \(\mathfrak{u}(2,1)\). The skeptical reader might also verify this using the structure equations (3.11). Furthermore, by definition, the associated Kähler metric satisfies \[\tau^*h=\frac{1}{2}\left(\zeta_1\circ \overline{\zeta_3}+\zeta_3\circ\overline{\zeta_1}+\zeta_2\circ\overline{\zeta_2}\right)\\ \] and hence the forms \(\frac{1}{\sqrt{2}}\zeta_i\) are a unitary coframe for \(\tau^*h_{\mathfrak{p}}\). In order to verify the Einstein condition it is therefore sufficient that the trace of the curvature form \[\Gamma=\mathrm{d}\gamma+\gamma\wedge\gamma\] is a non-zero constant (imaginary) multiple of \(\tau^*\Omega_Y\). We compute \[\begin{aligned} 0&=\mathrm{d}^2\zeta_3\wedge\zeta_1\wedge\overline{\zeta_1}=-\mathrm{d}\left(\gamma_{31}\wedge\zeta_1+\gamma_{32}\wedge\zeta_2+\gamma_{33}\wedge\zeta_3\right)\wedge\zeta_1\wedge\overline{\zeta_1}\\ &=\zeta_1\wedge\overline{\zeta_1}\wedge\left(\mathrm{d}L+\frac{1}{2}\overline{L}\zeta_2-L\varphi-2L\overline{\varphi}\right), \end{aligned}\] where we have used (3.11) and (3.12). It follows that there exist unique complex-valued functions \(L^{\prime}\) and \(L^{\prime\prime}\) on \(P\) such that \[\tag{3.13} \mathrm{d}L=L^{\prime}\zeta_1+L^{\prime\prime}\overline{\zeta_1}-\frac{1}{2}\overline{L}\zeta_2+L\varphi+2L\overline{\varphi}.\] Using the structure equations (3.11),(3.12) and (3.13) we compute \[\Gamma=\frac{1}{4}\begin{pmatrix} \Gamma_{11} & -\zeta_1\wedge\overline{\zeta_2} & \zeta_1\wedge\overline{\zeta_1}\\ -\zeta_2\wedge\overline{\zeta_3} & \Gamma_{22} & \overline{\zeta_1}\wedge\zeta_2\\ \zeta_3\wedge\overline{\zeta_3}+* & \overline{\zeta_2}\wedge\zeta_3& \Gamma_{33}\end{pmatrix},\] with \[\begin{aligned} \Gamma_{11}&=\frac{1}{4}\left(\zeta_3\wedge\overline{\zeta_1}-\zeta_2\wedge\overline{\zeta_2}-4\zeta_1\wedge\overline{\zeta_3}\right),\\ \Gamma_{22}&=\frac{1}{4}\left(-\zeta_1\wedge\overline{\zeta_3}-2\zeta_2\wedge\overline{\zeta_2}-\zeta_3\wedge\overline{\zeta_1}\right),\\ \Gamma_{33}&=\frac{1}{4}\left(\zeta_1\wedge\overline{\zeta_3}-\zeta_2\wedge\overline{\zeta_2}-4\zeta_3\wedge\overline{\zeta_1}\right). \end{aligned}\] and where \(*=4\left(L^{\prime}+\overline{L^{\prime}}\right)\zeta_1\wedge\overline{\zeta_1}\). In particular, we obtain \[\Gamma_{11}+\Gamma_{22}+\Gamma_{33}=4\mathrm{i}\tau^*\Omega_Y,\] thus verifying the Einstein property.

Remark 3.6

In [15], it is shown how to canonically associate a split-signature anti-self-dual Einstein metric on the total space of a certain rank two affine bundle \(A\) fibering over a projective surface \((\Sigma,\mathfrak{p})\). The indefinite Kähler–Einstein manifold \((Y,J,\Omega_Y\)) constructed here may be interpreted as the twistor space of this anti-self-dual Einstein metric.

3.4 The canonical flat case

In this subsection we identify the spaces \[Y=P/\mathrm{CO}(2) \quad\text{and} \quad Z=P/\left(\mathbb{R}_2\rtimes \mathrm{CO}(2)\right)\] in the case where \((\Sigma,\mathfrak{p})\) is the canonical flat projective structure on the projective \(2\)-sphere. Recall that in this case \(P=\mathrm{SL}(3,\mathbb{R})\). The group \(\mathrm{SL}(3,\mathbb{R})\) also acts naturally on \(\mathbb{C}_3\) by complexification, that is, by the rule \[g \cdot (\xi+\mathrm{i}\chi)=\xi g^{-1}+\mathrm{i}\chi g^{-1}\] for all \(g \in \mathrm{SL}(3,\mathbb{R})\). Clearly, this action descends to define a left action on \(\mathbb{CP}_2\). However, this action is not transitive, but has two orbits. The first orbit is \(\mathbb{RP}_2\subset \mathbb{CP}_2\), where we think of \(\mathbb{RP}_2\) as those points \([\xi+\mathrm{i}\chi] \in \mathbb{CP}_2\) which satisfy \(\xi\wedge\chi=0\), that is, \(\xi\) and \(\chi\) are linearly dependent over \(\mathbb{R}\). Assume therefore \([\varepsilon]\) is an element in the complement \(\mathbb{CP}_2\setminus\mathbb{RP}_2\) of \(\mathbb{RP}^2\) in \(\mathbb{CP}^2\). Since \(\mathrm{SL}(3,\mathbb{R})\) acts transitively on unimodular triples of vectors in \(\mathbb{R}_3\), we can assume without losing generality that \(\varepsilon=(0\;-\mathrm{i}\; 1)\). For \(g\in\mathrm{SL}(3,\mathbb{R})\) we write \(g=(g_0\;g_1\;g_2)\) with \(g_i \in \mathbb{R}^3\). We will next determine the stabiliser subgroup of \([\varepsilon]\). A simple computation gives \[g\cdot \varepsilon=g_0\wedge\left(g_1+\mathrm{i}g_2\right).\] An elementary calculation shows that \([g\cdot \varepsilon]=[\varepsilon]\) implies that we must have \[\begin{pmatrix} c_1 \\ c_2\end{pmatrix}=\begin{pmatrix}g^2_1 & -g^1_1 \\ g^2_2 & -g^1_2\end{pmatrix}\begin{pmatrix}g^1_0 \\ g^2_0 \end{pmatrix}=\begin{pmatrix} 0 \\ 0 \end{pmatrix}.\] Since \[\det g=g^0_2\,c_1-g^0_1\,c_2+g^0_0\,\det\begin{pmatrix}g^2_1 & -g^1_1 \\ g^2_2 & -g^1_2\end{pmatrix}=1,\] it follows that \(g^1_0=g^2_0=0\). Therefore, the stabiliser subgroup of \([\varepsilon]\) is a subgroup of \(\mathbb{R}_2\rtimes \mathrm{GL}(2,\mathbb{R})\). Writing \(a=(a^i_j)\), we obtain \[\left(b\rtimes a\right) \cdot \varepsilon=\det a^{-1}\begin{pmatrix}0 & -a^2_1-\mathrm{i}a^2_2 & a^1_1+\mathrm{i}a^1_2 \end{pmatrix},\] from which it follows that \([(b\rtimes a)\cdot \varepsilon]=[\varepsilon]\) if and only if \(a^1_1=a^2_2\) and \(a^1_2+a^2_1=0\), that is, \(a \in \mathrm{CO}(2)\). Concluding, we have shown \[\mathrm{SL}(3,\mathbb{R})/\left(\mathbb{R}_2\rtimes\mathrm{CO}(2)\right)\simeq \mathbb{CP}_2\setminus \mathbb{RP}_2\] and the projection map is \[\mu:\mathrm{SL}(3,\mathbb{R}) \to \mathbb{CP}_2\setminus \mathbb{RP}_2, \quad \begin{pmatrix} g_0 & g_1 & g_2 \end{pmatrix} \mapsto [g_0\wedge (g_1+\mathrm{i}g_2)],\] where we use \(\mathbb{R}_3\simeq \Lambda^2(\mathbb{R}^3)\).

Remark 3.7

We have only shown that \(Z=\mathrm{SL}(3,\mathbb{R})/\left(\mathbb{R}_2\times \mathrm{CO}(2)\right)\) is diffeomorphic to \(\mathbb{CP}_2\setminus\mathbb{RP}_2\). Since \(Z\) carries an integrable almost complex structure \(J\), we may ask if \((Z,J)\) is biholomorphic to \(\mathbb{CP}_2\setminus\mathbb{RP}_2\) equipped with the standard complex structure. This is indeed the case, see [37]. As a consequence of this result one can prove that the conformal connections on the \(2\)-sphere whose (unparametrised) geodesics are the great circles are in one-to-one correspondence with the smooth quadrics in \(\mathbb{CP}_2\setminus\mathbb{RP}_2\), see [37].

Remark 3.8

In fact [31], if \(\mathfrak{p}\) is a projective structure on the \(2\)-sphere, all of whose geodesics are simple closed curves, then \(Z\) can be compactified and the compactification is biholomorphic to \(\mathbb{CP}_2\). This allowed Lebrun and Mason to prove that there is a nontrivial moduli space of such projective structures on the \(2\)-sphere.

We will show next that \(Y\) is a submanifold of \(F(\mathbb{C}_3)\). Clearly, the action of \(\mathrm{SL}(3,\mathbb{R})\) on the space \(F(\mathbb{C}_3)\) of complete complex flags is not transitive, there is however an open orbit. Let \(F(\mathbb{C}_3)^*\) denote the \(\mathrm{SL}(3,\mathbb{R})\) orbit of the flag \[(\ell,\Pi)=\left(\mathbb{C}\{\varepsilon_1\},\mathbb{C}\{\varepsilon_1,\varepsilon_2\}\right),\] where \[\varepsilon_1=\begin{pmatrix} 0&-\mathrm{i}&1\end{pmatrix}, \quad \varepsilon_2=\begin{pmatrix} 1&0&0\end{pmatrix}.\quad %\eps_3=\begin{pmatrix} 0&1&0\end{pmatrix}.\] We already know that the stabiliser subgroup \(\mathrm{G}_0\) of \((\ell,\Pi)\) must be a subgroup of \(\mathbb{R}_2\rtimes \mathrm{CO}(2)\). For \(b\rtimes a \in \mathbb{R}_2\rtimes \mathrm{CO}(2)\) we write \[b\rtimes a=\begin{pmatrix} \frac{1}{x^2+y^2} & b_1 & b_2 \\ 0 & x & y \\ 0 & -y & x\end{pmatrix},\] with \(x^2+y^2>0\). We compute \[\varepsilon_2 \cdot (b\rtimes a)=\begin{pmatrix} x^2+y^2 & -xb_1-yb_2 & -xb_2+yb_1\end{pmatrix}\] which is easily seen to lie in the complex linear span of \(\varepsilon_1,\varepsilon_2\) if and only if \(b_1=b_2=0\), hence \[\mathrm{SL}(3,\mathbb{R})/\mathrm{CO}(2)\simeq F(\mathbb{C}_3)^*\] and the projection map is \[\tau : \mathrm{SL}(3,\mathbb{R}) \to F(\mathbb{C}_3) \quad \begin{pmatrix} g_0 & g_1 & g_2 \end{pmatrix} \mapsto \left(\mathbb{C}\{\varepsilon_1\},\mathbb{C}\{\varepsilon_1,\varepsilon_2\}\right),\] with \[\varepsilon_1=g_0\wedge (g_1+\mathrm{i}g_2), \qquad \varepsilon_2=g_1\wedge g_2.\] Since \(F(\mathbb{C}_3)\) is real six-dimensional and since \(\dim \mathrm{SL}(3,\mathbb{R})-\dim \mathrm{CO}(2)=6\), it follows that \(F(\mathbb{C}_3)^*\) is open.