On Kähler Metrisability of Two-dimensional Complex Projective Structures

2 Complex projective surfaces

2.1 Definitions

Let \(M\) be a complex \(2\)-manifold with integrable almost complex structure map \(J\) and \(\nabla\) an affine torsion-free connection on \(TM\). We call \(\nabla\) complex-linear if \(\nabla J=0\). A generalised geodesic for \(\nabla\) is a smoothly immersed curve \(\gamma \subset M\) with the property that the \(2\)-plane spanned by \(\dot{\gamma}\) and \(J \dot\gamma\) is parallel along \(\gamma\), i.e. \(\gamma\) satisfies the reparametrisation invariant condition \[\tag{2.1} \nabla_{\dot\gamma}\dot\gamma\wedge \dot\gamma\wedge J\dot\gamma=0.\] We call two complex linear torsion-free connections \(\nabla\) and \(\nabla^{\prime}\) on \(M\) complex projectively equivalent, if they have the same generalised geodesics. An equivalence class of complex projectively equivalent connections is called a complex projective structure and will be denoted by \([\nabla]\). A complex \(2\)-manifold equipped with a complex projective structure will be called a complex projective surface.

Remark 2.1

What we here call a complex projective structure was originally called a holomorphic projective structure by Tashiro [37] and others. Once it was realised that in general complex projective structures are not holomorphic in any reasonable way, the name h-projective structure was used – and is still so – see for instance [15, 21, 26]. Furthermore, what we here call generalised geodesics are called h-planar curves in the literature using the name h-projective. One might argue that the notion of a complex projective structure can be confused with well-established notions in algebraic geometry. For this reason complex projective is sometimes also abbreviated to c-projective (see for instance [2]).

Extending \(\nabla\) to the complexified tangent bundle \(T^{\mathbb{C}}M \to M\), it follows from the complex linearity of \(\nabla\) that for every local holomorphic coordinate system \(z=(z^i) : U \to \mathbb{C}^2\) on \(M\) there exist unique complex-valued functions \(\Gamma^i_{jk}\) on \(U\), so that \[\nabla_{\partial_{z^j}}\partial_{z^k}=\Gamma^i_{jk}\partial_{z^i}.\] We call the functions \(\Gamma^i_{jk}\) the complex Christoffel symbols of \(\nabla\). Tashiro showed [37] that two torsion-free complex linear connections \(\nabla\) and \(\nabla^{\prime}\) on \(M\) are complex projectively equivalent if and only if there exists a \((1,\! 0)\)-form \(\beta \in \Omega^{1,0}(M,\mathbb{R})\) so that \[\tag{2.2} \nabla^{\prime}_ZW-\nabla_ZW=\beta(Z)W+\beta(W)Z\] for all \((1,\! 0)\) vector fields \(Z,W \in \Gamma(T^{1,0}M)\). In analogy to the real case one can use (2.2) to show that \(\nabla\) and \(\nabla^{\prime}\) are complex projectively equivalent if and only if they induce the same parallel transport on the complex projectivised tangent bundle \(\mathbb{P}T^{1,0}M\).

Writing \(\Gamma^i_{jk}\) and \(\hat{\Gamma}^i_{jk}\) for the complex Christoffel symbols of \(\nabla\) and \(\nabla^{\prime}\) with respect to some holomorphic coordinates \(z=(z^i)\) and \(\beta=\beta_i \mathrm{d}z^i\), equation (2.2) translates to \[\tag{2.3} \hat{\Gamma}^i_{jk}=\Gamma^i_{jk}+\delta^i_j\beta_k+\delta^i_k\beta_j.\] Note that formally equation (2.3) is identical to the equation relating two real projectively equivalent connections on a real manifold. In particular, similarly to the real case (see [11, 38]), the functions \[\tag{2.4} \Pi^i_{jk}=\Gamma^i_{jk}-\frac{1}{3}\left(\Gamma^l_{lj}\delta^i_k+\Gamma^l_{lk}\delta^i_j\right)\] are complex projectively invariant in the sense that they only depend on the coordinates \(z\). Moreover locally \([\nabla]\) can be recovered from the functions \(\Pi^i_{jk}\) and two torsion-free complex linear connections are complex projectively equivalent if and only if they give rise to the same functions \(\Pi^i_{jk}\) in every holomorphic coordinate system.

A complex projective structure \([\nabla]\) is called holomorphic if the functions \(\Pi^i_{jk}\) are holomorphic in every holomorphic coordinate system. Gunning [17] obtained relations on characteristic classes of complex manifolds carrying holomorphic projective structures. The condition on a manifold to carry a holomorphic projective structure is particularly restrictive in the case of compact complex surfaces. See also the beautiful twistorial interpretation of holomorphic projective surfaces by Hitchin [19] and Remark 2.8.

2.2 Cartan geometry

A complex projective structure admits a description in terms of a normal Cartan geometry modelled on complex projective space \(\mathbb{CP}^n\), following the work of Ochiai [34]: see [20] and [40]. The reader unfamiliar with Cartan geometries may consult [9] for a modern introduction. We will restrict to the construction in the complex two-dimensional case.

Let \(\mathrm{PSL}(3,\mathbb{C})\) act on \(\mathbb{CP}^2\) from the left in the obvious way and let \(P\) denote the stabiliser subgroup of the element \([1,0,0]^t \in \mathbb{CP}^2\). We have:

Theorem 2.2

Suppose \((M,J,[\nabla])\) is a complex projective surface. Then there exists (up to isomorphism) a unique real Cartan geometry \((\pi : B \to M,\theta)\) of type \((\mathrm{PSL}(3,\mathbb{C}),P)\) such that for every local holomorphic coordinate system \(z=(z^i) : U \to \mathbb{C}^2\), there exists a unique section \(\sigma_z : U \to B\) satisfying \[\tag{2.5} (\sigma_z)^*\theta=\left(\begin{array}{ccc} 0 & \phi^0_1 & \phi^0_2\\ \phi^1_0 & \phi^1_1 & \phi^1_2 \\ \phi^2_0 & \phi^2_1 & \phi^2_2\end{array}\right)\] where \[\phi^i_0=\mathrm{d}z^i, \quad \text{and} \quad \phi^i_j=\Pi^i_{jk}\mathrm{d}z^k, \quad \text{and} \quad \phi^0_i=\Pi_{ik}\mathrm{d}z^k,\] with \[\Pi_{ij}=\Pi^k_{il}\Pi^l_{jk}-\frac{\partial \Pi^k_{ij}}{\partial z^k}\] and \(\Pi^i_{jk}\) denote the complex projective invariants with respect to \(z^i\) defined in (2.4).

Remark 2.3

Suppose \(\varphi : (M,J,[\nabla]) \to (M^{\prime},J^{\prime},[\nabla]^{\prime})\) is a biholomorphism between complex projective surfaces identifying the complex projective structures, then there exists a diffeomorphism \(\hat{\varphi} : B \to B^{\prime}\) which is a \(P\)-bundle map covering \(\varphi\) and which satisfies \(\hat{\varphi}^*\theta^{\prime}=\theta\). Conversely, every diffeomorphism \(\Phi : B \to B^{\prime}\) that is a \(P\)-bundle map and satisfies \(\Phi^*\theta^{\prime}=\theta\) is of the form \(\Phi=\hat{\varphi}\) for a unique biholomorphism \(\varphi : M \to M^{\prime}\) identifying the complex projective structures.

Example 2.4

Let \(B=\mathrm{PSL}(3,\mathbb{C})\) and let \(\theta\) denote its Maurer-Cartan form. Setting \(M=B/P\simeq \mathbb{CP}^2\) and \(\pi : \mathrm{PSL}(3,\mathbb{C}) \to \mathbb{CP}^2\) the natural quotient projection, one obtains a complex projective structure on \(\mathbb{CP}^2\) whose generalised geodesics are the smoothly immersed curves \(\gamma \subset \mathbb{CP}^1\) where \(\mathbb{CP}^1\subset \mathbb{CP}^2\) is any linearly embedded projective line. This is precisely the complex projective structure associated to the Levi-Civita connection of the Fubini-Study metric on \(\mathbb{CP}^2\). This example satisfies \(\mathrm{d}\theta+\theta\wedge\theta=0\) and is hence called flat.

Let \((\pi : B \to M,\theta)\) be the Cartan geometry of a complex projective structure \((J,[\nabla])\) on a simply-connected surface \(M\) whose Cartan connection satisfies \(\mathrm{d}\theta+\theta\wedge\theta=0\). Then there exists a local diffeomorphism \(\Phi : B \to \mathrm{PSL}(3,\mathbb{C})\) pulling back the Maurer-Cartan form of \(\mathrm{PSL}(3,\mathbb{C})\) to \(\theta\) and consequently, a local biholomorphism \(\varphi : M \to \mathbb{CP}^2\) identifying the projective structure on \(M\) with the standard flat structure on \(\mathbb{CP}^2\).

2.3 Bianchi-identities

Theorem 2.2 implies that the curvature form \(\Theta=\mathrm{d}\theta+\theta\wedge\theta\) satisfies \[\tag{2.6} \Theta=\mathrm{d}\theta+\theta\wedge\theta=\left(\begin{array}{ccc} 0 & \Theta^0_1 & \Theta^0_2\\ 0 & \Theta^1_1 & \Theta^1_2\\ 0 & \Theta^2_1 & \Theta^2_2\end{array}\right)\] with \[\Theta^0_i=L_i\theta^1_0\wedge\theta^2_0+K_{il\bar\jmath}\theta^l_0\wedge\overline{\theta^{\jmath}_0}, \quad \Theta^i_k=W^i_{kl\bar\jmath}\theta^l_0\wedge\overline{\theta^\jmath_0}\] for unique complex-valued functions \(L_i,K_{il\bar\jmath}\), and \(W^i_{kl\bar\jmath}\) on \(B\) satisfying \(W^l_{li\bar \jmath}=0\). Note that by construction, with respect to local holomorphic coordinates \(z=(z^i)\), we obtain \[\tag{2.7} (\sigma_z)^*W^i_{kl\bar\jmath}=-\frac{\partial \Pi^i_{kl}}{\partial \bar z^j}.\]

Differentiation of the structure equations (2.6) gives \[0=\mathrm{d}^2\theta^i_0=W^i_{lk\bar\jmath}\theta^l_0\wedge\theta^k_0\wedge\overline{\theta^\jmath_0}, \quad \text{and} \quad 0=\mathrm{d}^2 \theta^0_0=K_{ik\bar\jmath}\theta^i_0\wedge\theta^k_0\wedge\overline{\theta^\jmath_0}\] which yields the algebraic Bianchi-identities \[W^i_{lk\bar\jmath}=W^i_{kl\bar\jmath}, \quad \text{and} \quad K_{ik\bar\jmath}=K_{ki\bar\jmath}.\]

2.3.1 Complex projective Weyl curvature

The identities \(\mathrm{d}^2\theta^i_k=0\) yield \[\kappa^i_{kl\bar\jmath}\wedge\theta^l_0\wedge\overline{\theta^{\jmath}_0}=0\] with \[\kappa^i_{kl\bar\jmath}=\mathrm{d}W^i_{kl\bar\jmath}+W^i_{kl\bar\jmath}\left(\theta^0_0+\overline{\theta^0_0}\right)+K_{kl\bar\jmath}\theta^i_0-W^i_{ls\bar\jmath}\theta^s_k-W^i_{ks\bar\jmath}\theta^s_l+W^s_{kl\bar\jmath}\theta^i_s-W^i_{kl\bar s}\overline{\theta^s_l}\] which implies that there exist complex-valued functions \(W^i_{kl\bar\jmath\bar s}\) and \(W^i_{kl\bar\jmath s}\) on \(B\) satisfying \[W^i_{kl\bar\jmath\bar s}=W^i_{lk\bar\jmath\bar s}=W^i_{kl\bar s\bar\jmath}, \quad W^k_{kl\bar\jmath\bar s}=W^k_{kl\bar\jmath s}=0, \quad W^i_{kl\bar\jmath s}=W^i_{lk\bar\jmath s}\] such that \[\tag{2.8} \mathrm{d}W^i_{kl\bar\jmath}=\left(W^i_{kl\bar\jmath s}+\delta^i_kK_{sl\bar\jmath}+\delta^i_lK_{sk\bar\jmath}-3\delta^i_sK_{kl\bar\jmath}\right)\theta^s_0+W^i_{kl\bar\jmath\bar s}\overline{\theta^s_0}+\varphi^i_{kl\bar\jmath}\] where \[\tag{2.9} \varphi^i_{kl\bar\jmath}=-W^i_{kl\bar\jmath}\left(\theta^0_0+\overline{\theta^0_0}\right)+W^i_{ls\bar\jmath}\theta^s_k+W^i_{ks\bar\jmath}\theta^s_l-W^s_{kl\bar\jmath}\theta^i_s+W^i_{kl\bar s}\overline{\theta^s_j}.\]

Let \(\mathrm{End}_0(TM,J)\) denote the bundle whose fibre at \(p \in M\) consists of the \(J\)-linear endomorphisms of \(T_pM\) which are complex-traceless. It follows with the structure equations (2.6),(2.8),(2.9) and straightforward computations, that there exists a unique \((1,\!1)\)-form \(W\) on \(M\) with values in \(\mathrm{End}_0(TM,J)\) for which \[W\left(\frac{\partial }{\partial z^l},\frac{\partial }{\partial \overline{z^{\jmath}}}\right)\frac{\partial }{\partial z^k}=(\sigma_z)^*W^i_{kl\bar\jmath}\frac{\partial }{\partial z^i}=-\frac{\partial \Pi^i_{kl}}{\partial \bar z^j}\frac{\partial }{\partial z^i}\] in every local holomorphic coordinate system \(z=(z^i)\) on \(M\). Here, as usual, we extend tensor fields on \(M\) complex multilinearly to the complexified tangent bundle of \(M\). The bundle-valued \(2\)-form \(W\) is called the complex projective Weyl curvature of \([\nabla]\). We obtain:

Proposition 2.5

A complex projective structure \([\nabla]\) on a complex surface \((M,J)\) is holomorphic if and only if the complex projective Weyl tensor of \([\nabla]\) vanishes.

2.3.2 Complex projective Liouville curvature

From \(\mathrm{d}^2\theta^0_i\wedge\overline{\theta^1_0}\wedge\overline{\theta^2_0}=0\) one sees after a short computation that \[\tag{2.10} \mathrm{d}L_i=-4L_i\theta^0_0+L_j\theta^j_i+L_{ij}\theta^j_0+L_{i\bar\jmath}\overline{\theta^{\jmath}_0}\] for unique complex-valued functions \(L_{i\bar\jmath},L_{ij}\) on \(B\). Using this last equation it is easy to check that the \(\pi\)-semibasic quantity \[\tag{2.11} (L_1\theta^1_0+L_2\theta^2_0)\otimes \left(\theta^1_0\otimes \theta^2_0\right)\] is invariant under the \(P\) right action and thus the \(\pi\)-pullback of a tensor field \(\lambda\) on \(M\) which is called the complex projective Liouville curvature (see the note of R. Liouville [24] for the construction of \(\lambda\) in the real case).

Remark 2.6

In the case of real projective structures on surfaces, the projective Weyl curvature vanishes identically. Furthermore, note that contrary to the complex projective Liouville curvature, the complex projective Weyl tensor exists as well in higher dimensions, but also contains \((2,\! 0)\) parts (see [37] for details).

The differential Bianchi-identity (2.8) implies that if the functions \(W^i_{kl\bar\jmath}\) vanish identically, then the functions \(K_{ik\bar\jmath}\) must vanish identically as well. We have thus shown:

Proposition 2.7

A complex projective structure \([\nabla]\) on a complex surface \((M,J)\) is flat if and only the complex projective Liouville and Weyl curvature vanish.

Remark 2.8

In [22] Kobayashi and Ochiai classified compact complex surfaces carrying flat complex projective structures. More recently Dumitrescu [13] showed among other things that a holomorphic projective structure on a compact complex surface must be flat (see also the results by McKay about holomorphic Cartan geometries [27]).

2.3.3 Further identities

We also obtain \[0=\mathrm{d}^2\theta^0_i=\kappa_{ik\bar\jmath}\wedge\overline{\theta^{\jmath}_0}\wedge\theta^k_0\] with \[\begin{aligned} \kappa_{ik\bar\jmath}=&-\mathrm{d}K_{ik\bar\jmath}+\frac{1}{2}\varepsilon_{sk}L_{i\bar\jmath}\theta^s_0-K_{ik\bar\jmath}\left(2\theta^0_0+\overline{\theta^0_0}\right)+K_{sk\bar\jmath}\theta^s_i+K_{si\bar\jmath}\theta^s_k-\\ &-W^s_{ik\bar\jmath}\theta^0_s+K_{ik\bar s}\overline{\theta^s_j}. \end{aligned}\] It follows that there are complex-valued functions \(K_{ik\bar\jmath l}\) and \(K_{kl\bar\imath\bar\jmath}\) on \(B\) satisfying \[K_{ik\bar\jmath l}=K_{ki \bar\jmath l}, \quad \text{and} \quad K_{kl\bar \imath \bar\jmath}=K_{lk\bar \imath \bar\jmath}=K_{kl \bar\jmath\bar \imath}\] such that \[\tag{2.12} \mathrm{d}K_{ik\bar\jmath}=\left(K_{ik\bar\jmath s}+\frac{1}{4}\left(\varepsilon_{sk}L_{i\bar\jmath}+\varepsilon_{si}L_{k\bar\jmath}\right)\right)\theta^s_0+K_{ik\bar\jmath\bar s}\overline{\theta^s_0}+\varphi_{ik\bar\jmath}\] where \[\varphi_{ik\bar\jmath}=-K_{ik\bar\jmath}\left(2\theta^0_0+\overline{\theta^0_0}\right)+K_{sk\bar\jmath}\theta^s_{i}+K_{si\bar\jmath}\theta^s_k-W^s_{ik\bar\jmath}\theta^0_s+K_{ik\bar s}\overline{\theta^s_j}.\]

2.4 Complex and generalised geodesics

It is worth explaining how the generalised geodesics of \([\nabla]\) appear in the Cartan geometry \((\pi : B \to M,\theta)\). To this end let \(G\subset P\subset \mathrm{PSL}(3,\mathbb{C})\) denote the quotient group of the group of upper triangular matrices of unit determinant modulo its center. The quotient \(B/G\) is the total space of a fibre bundle over \(M\) whose fibre \(P/G\) is diffeomorphic to \(\mathbb{CP}^1\). In fact, \(B/G\) may be identified with the total space of the the complex projectivised tangent bundle \(\tau : \mathbb{P}(T^{1,0}M) \to M\) of \((M,J)\). Writing \(\theta=(\theta^i_j)_{i,j=0..2}\), Theorem 2.2 implies that the real codimension \(4\)-subbundle of \(TB\) defined by \(\theta^2_0=\theta^2_1=0\) descends to a real rank \(2\) subbundle \(E\subset T\mathbb{P}(T^{1,0}M)\). The integral manifolds of \(E\) can most conveniently be identified in local coordinates. Let \(z=(z^1,z^2) : U \to \mathbb{C}^2\) be a local holomorphic coordinate system on \(M\) and write \(\phi\) for the pullback of \(\theta\) with the unique section \(\sigma_z\) associated to \(z\) in Theorem 2.2. We obtain a local trivialisation of Cartan’s bundle \[\varphi : U \times P \to \pi^{-1}(U)\] so that for \((z,p) \in U\times P\) we have \[\tag{2.13} (\varphi^*\theta)_{(z,p)}=\left(\omega_P\right)_p+\mathrm{Ad}(p^{-1})\circ \phi_z\] where \(\omega_P\) denotes the Maurer-Cartan form of \(P\) and \(\mathrm{Ad}\) the adjoint representation of \(\mathrm{PSL}(3,\mathbb{C})\). Consider the Lie group \(\tilde{P}\subset \mathrm{SL}(3,\mathbb{C})\) whose elements are of the form \[\tag{2.14} \left(\begin{array}{cc} \det a^{-1} & b \\ 0 & a \end{array}\right)\] for \(a\in \mathrm{GL}(2,\mathbb{C})\) and \(b^{t} \in \mathbb{C}^2\). By construction, the elements of \(P\) are equivalence classes of elements in \(\tilde{P}\) where two elements are equivalent if they differ by scalar multiplication with a complex cube root of \(1\). The canonical projection \(\tilde{P} \to P\) will be denoted by \(\nu\). Note that a piece \(N\) of an integral manifold of \(E\) that is contained in \(\tau^{-1}(U)\) is covered by a map \[(z^1,z^2,p) : N \to U \times \tilde{P}\] where \(p : N \to \tilde{P}\) may be taken to be of the form \[p=\left(\begin{array}{ccr} \frac{1}{(a_1)^2+(a_2)^2} & 0 & 0 \\ 0 & a_1 & -a_2 \\ 0 & a_2 & a_1\end{array}\right)\] for smooth complex-valued functions \(a_i : N \to \mathbb{C}\) satisfying \((a_1)^2+(a_2)^2\neq 0\).

We first consider the case where \(N\) is one-dimensional. We fix a local coordinate \(t\) on \(N\). It follows with (2.13) and straightforward calculations that \[\left(\varphi \circ (z^1,z^2,\nu\circ p)\right)^*\theta^2_0=\frac{a_1\dot{z}^2-a_2\dot{z}^1}{\left((a_1)^2+(a_2)^2\right)^2}\mathrm{d}t\] where \(\dot{z}^i\) denote the derivative of \(z^i\) with respect to \(t\). Hence we may take \[a_1=\dot{z}^1 \quad \text{and} \quad a_2=\dot{z}^2.\] Writing \(\beta=\left(\varphi \circ (z^1,z^2,\nu\circ p)\right)^*\theta^2_1\) and using (2.13) again, we compute \[\begin{aligned} \beta=&\left[\dot{z}^1\ddot{z}^2-\dot{z}^2\ddot{z}^1+\left(\dot{z}^1\dot{z}^2(\Pi^2_{21}-\Pi^1_{11})+(\dot{z}^1)^2\Pi^2_{11}-(\dot{z}^2)^2\Pi^1_{12}\right)\dot{z}^1+\right.\\ &\left.+\left(\dot{z}^1\dot{z}^2(\Pi^2_{22}-\Pi^1_{12})+(\dot{z}^1)^2\Pi^2_{12}-(\dot{z}^2)^2\Pi^1_{22}\right)\dot{z}^2\right]\frac{\mathrm{d}t}{(\dot{z}^1)^2+(\dot{z}^2)^2}. \end{aligned}\] Note that since \(\Pi^i_{ik}=0\) for \(k=1,2\), it follows that \(\beta\equiv 0\) is equivalent to \((z^1,z^2)\) satisfying the following ODE system \[\dot{z}^i\left(\ddot{z}^j+\Pi^j_{kl}\dot{z}^k\dot{z}^l\right)=\dot{z}^j\left(\ddot{z}^i+\Pi^i_{kl}\dot{z}^k\dot{z}^l\right), \quad i,j=1,2.\] This last system is easily seen to be equivalent to the system (2.1). Consequently, the one-dimensional integral manifolds of \(E\) are the generalised geodesics of \([\nabla]\).

Note that in the case of two-dimensional integral manifolds the above computations carry over where \(t\) is now a complex parameter, i.e. the two-dimensional integral manifolds are immersed complex curves \(Y \subset M\) for which \(\nabla_{\dot{Y}}\dot{Y}\) is proportional to \(\dot{Y}\) for some (and hence any) \(\nabla \in [\nabla]\). This last condition is equivalent to \(Y\) being a totally geodesic immersed complex curve with respect to \(([\nabla],J)\) (c.f. [32]) . A totally geodesic immersed complex curve \(Y\subset M\) which is maximally extended is called a complex geodesic. Since the complex geodesics are the (maximally extended) two-dimensional integral manifolds of \(E\), they exist only provided that \(E\) is integrable. We will next determine the integrability conditions for \(E\). Recall that \(E\subset T\mathbb{P}(T^{1,0}M)\) is defined by the equations \(\theta^2_1=\theta^2_0=0\) on \(B\). It follows with the structure equations (2.6) that \[\mathrm{d}\theta^2_0=0 \quad \text{mod}\quad \theta^2_0,\theta^2_1\] and \[\mathrm{d}\theta^2_1=W^2_{11\bar\jmath}\theta^1_0\wedge \overline{\theta^{\jmath}_0}\quad \text{mod}\quad \theta^2_0,\theta^2_1.\] Consequently, \(E\) is integrable if and only if \(W^2_{11\bar 1}=W^2_{11\bar 2}=0\). As a consequence of (2.8) and \(W^2_{11\bar 1}=0\) we obtain \[0=\varphi^2_{11\bar 1}=-W^2_{11\bar 1}\left(\theta^0_0+\overline{\theta^0_0}\right)+W^2_{1s\bar 1}\theta^s_1+W^2_{1s\bar 1}\theta^s_1-W^s_{11\bar 1}\theta^2_s+W^2_{11\bar s}\overline{\theta^s_j},\] which is equivalent to \(2W^2_{12\bar 1}=W^1_{11\bar 1}\). Using the symmetries of the complex projective Weyl tensor we compute \[W^1_{11\bar 1}=-W^2_{21\bar 1}=2W^2_{12\bar 1}=2W^2_{21\bar 1},\] thus showing that \(W^1_{11\bar 1}=W^2_{12\bar 1}=0\). From this we obtain \[0=\varphi^1_{11\bar 1}=2W^1_{12\bar 1}\theta^2_1-W^2_{11\bar 1}\theta^1_2+W^1_{11\bar 2}\overline{\theta^2_1}.\] thus implying \(W^1_{12\bar 1}=W^2_{11\bar 1}=W^1_{11\bar 2}=0\). Continuing in this vein allows to conclude that all components of the complex projective Weyl tensor must vanish. We may summarise:

Proposition 2.9

Let \((M,J,[\nabla])\) be a complex projective surface. Then the following statements are equivalent:

  • \([\nabla]\) is holomorphic;

  • The complex projective Weyl tensor of \([\nabla]\) vanishes;

  • The rank \(2\) bundle \(E \to \mathbb{P}(T^{1,0}M)\) is Frobenius integrable;

  • Every complex line \(L\subset T^{1,0}M\) is tangent to a unique complex geodesic.

Remark 2.10

The standard flat complex projective structure on \(\mathbb{CP}^2\) is holomorphic and the complex geodesics are simply the linearly embedded projective lines \(\mathbb{CP}^1\subset \mathbb{CP}^2\).

Remark 2.11

Note that the integrability conditions for \(E\) are a special case of a more general result obtained by Čap in [8]. There it is shown that \(E\) is part of an elliptic CR structure of CR dimension and codimension \(2\), which the complex projective structure induces on \(\mathbb{P}(T^{1,0}M)\). Furthermore, it is also shown that the integrability of \(E\) is equivalent to the holomorphicity of the complex projective surface.