Home The anti-self-dual deformation complex and a conjecture of Singer
Article Open Access

The anti-self-dual deformation complex and a conjecture of Singer

  • A. Rod Gover and Matthew J. Gursky EMAIL logo
Published/Copyright: May 16, 2024

Abstract

Let ( M 4 , g ) be a smooth, closed, oriented anti-self-dual (ASD) four-manifold. ( M 4 , g ) is said to be unobstructed if the cokernel of the linearisation of the self-dual Weyl tensor is trivial. This condition can also be characterised as the vanishing of the second cohomology group of the ASD deformation complex, and is central to understanding the local structure of the moduli space of ASD conformal structures. It also arises in construction of ASD manifolds by twistor and gluing methods. In this article, we give conformally invariant conditions which imply an ASD manifold of positive Yamabe type is unobstructed.

1 Introduction

Let M 4 be a smooth, closed, oriented four-manifold. Given a Riemannian metric 𝑔 on M 4 , the bundle of two-forms Λ 2 = Λ 2 ( M 4 ) splits into the subbundles of self-dual and anti-self-dual two-forms under the action of the Hodge ⋆-operator:

Λ 2 = Λ + 2 Λ 2 .

By a result of Singer–Thorpe [37], the curvature operator Rm : Λ 2 Λ 2 has a canonical block decomposition of the form

Rm = ( A + B B t A ) ,

where A ± : Λ ± 2 Λ ± 2 and B : Λ + 2 Λ 2 . If 𝑊 denotes the Weyl tensor, then W ± : Λ ± 2 Λ ± 2 and

A ± = W ± + 1 12 R I ,

where 𝐼 is the identity and 𝑅 is the scalar curvature.

Definition 1.1

We say that ( M 4 , g ) is anti-self-dual (ASD) if W g + 0 .

The notion of (anti-)self-duality is conformally invariant: if W g + = 0 for a metric 𝑔 and g ̃ = e f g , then W g ̃ + = 0 . This property will be crucial for the proof of our main result below.

There are topological obstructions to the existence of ASD metrics. By the Hirzebruch signature formula,

48 π 2 τ ( M 4 ) = ( | W g + | 2 | W g | 2 ) d v g ,

where τ ( M 4 ) is the signature of the intersection form on H dR 2 ( M 4 ) . In particular, we see that if ( M 4 , g ) is ASD, then τ ( M 4 ) 0 , with equality if and only if 𝑔 is LCF. If ( M 4 , g ) is ASD with positive scalar curvature, then the intersection form is actually definite (see [32, Proposition 1]). To see this, we first observe that the splitting of Λ 2 ( M 4 ) induces a splitting on the space of harmonic two-forms; hence H dR 2 ( M 4 ) = H + 2 ( M 4 ) H 2 ( M 4 ) . If ω H + 2 ( M 4 ) , then the Weitzenböck formula for the Hodge Laplacian Δ 2 is given by

(1.1) Δ 2 ω = Δ ω + 2 W + ( ω ) 1 3 R ω ,

where Δ = g i j i j is the rough Laplacian. If ( M 4 , g ) is ASD and the scalar curvature R > 0 , then (1.1) immediately implies ω = 0 .

Examples of ASD manifolds include locally conformally flat (LCF) manifolds since, in dimensions greater than three, LCF is equivalent to the vanishing of the Weyl tensor. In particular, S 4 endowed with the round metric g c is ASD. Non-simply connected examples include the product metric on S 3 × S 1 , and more generally the metrics constructed via gluing on the connected sums S 3 × S 1 # # S 3 × S 1 .

A non-LCF example is given by complex projective space with the Fubini–Study metric, and we take the opposite of the its natural orientation as a complex manifold, i.e., ( C P 2 , g FS ) . Any scalar-flat Kähler metric is also ASD since the Kähler condition implies that W + is determined by 𝑅 (see [12, Section 3]). There are many constructions of ASD manifolds in the literature; see for example [35, 36, 33, 13, 17, 38, 27, 28, 26] and the references in [40, Lecture 6].

Roughly speaking, the constructions of ASD manifolds are based on either “twistor” or “analytic” methods. The former approach relies on the so-called Penrose correspondence, which will play no role in our work but is of profound importance in the study of ASD manifolds. Briefly, the unit sphere bundle 𝒵 of Λ + 2 ( M 4 ) carries a canonical complex structure. As shown in [1], this complex structure is integrable if and only if the metric is ASD. Therefore, we can associate to any ASD manifold a complex manifold of (complex) dimension three, called the twistor space of ( M 4 , g ) . This important observation allows one to use methods of complex geometry to study the existence and deformation theory of ASD conformal structures.

Analytic methods involve the construction of an ASD metric on the connected sum of two manifolds admitting ASD metrics via perturbative methods. As in other geometric gluing constructions, if ( M 1 , g 1 ) and ( M 2 , g 2 ) are ASD manifolds, one first constructs a metric ℎ on the connected sum M 1 # M 2 which is “approximately” ASD, i.e., W + ( h ) is small in some appropriately defined norm. This reduces the problem to the study of the mapping properties of the linearised operator, in order to perturb ℎ to produce an actual ASD metric. To make this more precise, we now introduce the ASD deformation complex.

1.1 The ASD deformation complex

Let M ( M 4 ) be the space of smooth Riemannian metrics on M 4 , and R ( M 4 ) the bundle of algebraic curvature tensors. We can view W + as a mapping

W + : M ( M 4 ) R ( M 4 ) .

Let g M ( M 4 ) be an ASD metric. We can identify the formal tangent space of ℳ at 𝑔 with sections of the bundle of symmetric two-tensors, S 2 ( T M 4 ) . Let

D : Γ ( S 2 ( T M 4 ) ) Γ ( R ( M 4 ) )

denote the linearisation of W + at 𝑔, i.e., for h Γ ( S 2 ( T M 4 ) ) ,

D g h = d d s W + ( g + s h ) | s = 0 .

The choice of a conformal class of metrics [ g ] determines the bundle of algebraic Weyl tensors, W = W ( M 4 , [ g ] ) , and the subbundle W + W S 0 2 ( Λ + 2 ) , where S 0 2 ( Λ + 2 ) is the bundle of symmetric, trace-free endomorphisms of Λ + 2 . Note that

D g : Γ ( S 2 ( T M 4 ) ) Γ ( W + ) Γ ( S 0 2 ( Λ + 2 ) ) .

In fact, since 𝑔 is ASD, by conformal invariance, D g ( f g ) = 0 for any f C ( M 4 ) ; hence

D g : Γ ( S 0 2 ( T M 4 ) ) Γ ( W + ) .

We also let

D g : Γ ( W + ) Γ ( S 0 2 ( T M 4 ) )

denote the L 2 -formal adjoint of D g . Although the formula for D g is somewhat involved, the formula for D is much more compact (see [24, Proposition A.4]):

D U i j = 2 ( k U i k j + P k U i k j ) ,

where 𝑃 denotes the Schouten tensor (see Section 2).

Let K g : Γ ( T M 4 ) S 0 2 ( T M 4 ) denote the Killing operator,

K g ( ω ) i j = i ω j + j ω i 1 2 ( δ g ω ) g i j ,

where δ g ω = k ω k is the divergence of 𝜔. The kernel of 𝒦 consists of those one-forms whose dual vector fields are conformal Killing. Moreover, by diffeomorphism and conformal invariance, Im K ker D .

The ASD deformation complex is given by

(1.2) Γ ( T M 4 ) K Γ ( S 0 2 ( T M 4 ) ) D Γ ( S 0 2 ( Λ + 2 ) ) .

This complex is elliptic; see [29, Section 2]. The associated cohomology groups are given by

H ASD 0 ( M 4 , g ) = ker K g , H ASD 1 ( M 4 , g ) = { h Γ ( S 0 2 ( T M 4 ) ) : δ g h = 0 , D g h = 0 } , H ASD 2 ( M 4 , g ) = ker D g ,

where, for h Γ ( S 2 ( T M 4 ) ) , ( δ g h ) j = g i k k h i j is the divergence. The Atiyah–Singer index theorem can be used to calculate the index of (1.2):

(1.3) Ind ASD ( M 4 , g ) = 1 2 ( 15 χ ( M 4 ) + 29 τ ( M 4 ) ) .

Vanishing of the cohomology groups also provides information on the local structure of the moduli space of ASD conformal structures.

Proposition 1.2

Proposition 1.2 (see [24, 40])

Suppose ( M 4 , g ) is ASD with

H ASD 0 ( M 4 , g ) = { 0 } , H ASD 2 ( M 4 , g ) = { 0 } .

Then the moduli space of anti-self-dual conformal structures near 𝑔 is a smooth, finite-dimensional manifold of dimension dim H ASD 1 ( M 4 , g ) .

This leads to the following definition.

Definition 1.3

Let ( M 4 , g ) be ASD. We say that ( M 4 , g ) is unobstructed if

H ASD 2 ( M 4 , g ) = { 0 } .

By the work of Floer [17] and Donaldson–Friedman [13], if ( M 1 , g 1 ) and ( M 2 , g 2 ) are unobstructed ASD manifolds, then the connected sum M 1 # M 2 admits an ASD metric. Thus, we are lead to the question: under what condition is an ASD manifold unobstructed? The following conjecture is often attributed to Singer[1].

Conjecture 1.4

Let ( M 4 , g ) be ASD. If the Yamabe invariant of ( M 4 , g ) is positive, then ( M 4 , g ) is unobstructed.

For ASD Einstein manifolds, the operators D D and D D were explicitly computed in [25] and [30]. It follows from these calculations (and can also be seen by more or less direct calculation) that ( S 4 , g c ) and ( C P 2 , g FS ) are unobstructed. We remark that, for non-Einstein ASD manifolds, the formulas for these operators are fairly intractable.

The vanishing of H ASD 2 ( M 4 , [ g ] ) can sometimes be verified when the twistor space is explicitly known. For example, the LCF metrics k # S 3 × S 1 (see [34, Theorem 8.2], [14]), and the ASD metrics on m # ( C P 2 ) constructed by LeBrun [33], are unobstructed. Our goal in this paper is to provide a criterion for the vanishing of H ASD 2 that only involves conformal invariants of the ASD manifold (and in particular does not depend on verifying any properties of the twistor space). Our main result is the following.

Theorem 1.5

Suppose ( M 4 , g ) is ASD with Yamabe invariant Y ( M 4 , [ g ] ) > 0 . If

(1.4) 2 χ ( M 4 ) + 3 τ ( M 4 ) 1 24 π 2 Y ( M 4 , [ g ] ) 2 ,

then ( M 4 , g ) is unobstructed.

The proof of Theorem 1.5 relies on two key ideas. The first is that any element U ker D can be associated to a self-dual harmonic two-form z = z ( U ) Λ + 2 ( A ) taking its values in the adjoint tractor bundle (see Section 2). Therefore, 𝑧 satisfies a twisted version of the usual Weitzenböck formula for self-dual (real-valued) two-forms. This twisted version provides us with two identities for 𝑈 (see Theorem 3.6).

The second key idea is to make a judicious choice of conformal representative in order to show that (1.4) implies the vanishing of 𝑈. As explained in Section 4, we choose a conformal metric whose scalar curvature satisfies a differential inequality involving the Schouten tensor (see Theorem 4.1), and is adapted to the curvature terms appearing in the Weitzenböck formula(s) for 𝑈. Curiously, to prove the existence of this metric, we consider a modification of the functional determinant of a conformally covariant elliptic operator first computed by Branson and Ørsted [4]. A robust theory for the existence of critical points of this functional was developed by Chang and Yang [10], and we are able to show that their ideas also give existence for our modified functional.

If Y ( M 4 , [ g ] ) > 0 , then by the estimate of Aubin,

1 24 π 2 Y ( M 4 , [ g ] ) 2 1 24 π 2 Y ( S 4 , [ g c ] ) 2 16 .

Therefore, (1.4) will fail (for topological reasons) if ( 2 χ + 2 τ ) is sufficiently negative. On the other hand, using Kobayashi’s estimate [31] of the Yamabe invariant of connected sums, it is easy to construct examples of ASD metrics on connected sums of C P 2 or S 3 × S 1 satisfying (1.4).

1.2 Organisation

In Section 2, we provide the necessary background on the tractor bundle and associated connection and metric. The main result (for our purposes) is Proposition 2.2, in which we associate to U ker D a twisted SD harmonic two-form

z = z ( U ) Λ + 2 ( A ) .

In Section 3, we compute two Weitzenböck formulas for 𝑈 that follow from this correspondence. In Section 4, we give the proof of Theorem 1.5, and relegate the PDE aspects of our work to the appendix.

2 Background and the interpretation via tractor calculus

2.1 Some conventions for Riemannian geometry

For index calculations, we will use Penrose’s abstract index notation, unless otherwise indicated. In this, we write E a and E a as alternative notations for, respectively, the cotangent and tangent bundles and the contraction ω ( v ) of a one-form 𝜔 with a tangent vector v a is written with a repeated index ω a v a (mimicking the Einstein summation convention). Tensor bundles are denoted then by adorning the symbol ℰ with appropriate indices and sometimes also indicating symmetries. For example, E ( a b ) is the notation for S 2 T M , the subbundle of symmetric tensors in T M T M .

Then our convention for the Riemann tensor R a b c d is such that

[ a , b ] v c = R a b c d v d , [ a , b ] ω c = R a b c d ω d ,

where a is the Levi-Civita connection of a metric g a b , 𝑣 any tangent vector field, and 𝜔 is any one-form. Using the metric to raise and lower indices, we have, for example,

R a b c d = g c e R a b e d .

This may be decomposed as

(2.1) R a b c d = W a b c d + 2 ( g c [ a P b ] d + g d [ b P a ] c ) ,

where the completely trace-free part W a b c d is the Weyl tensor and P a b is the Schouten tensor. It follows that

P a b : = 1 n 2 ( R a b R 2 ( n 1 ) g a b ) ,

where R b c = R a b a c is the Ricci tensor and its metric trace R = g a b R a b is the scalar curvature. We will use 𝐽 to denote the metric trace of Schouten, i.e., J : = g a b P a b . Lastly, we have

C a b c := 2 [ b P c ] a , B a b := c C a b c + P d c C d a c b ,

where C a b c and B a b are Cotton and Bach tensors, respectively. It should be noted that the Bianchi identities imply ( n 3 ) C a b c = d W d a b c .

2.2 The tractor bundle and connection

To treat and work with objects that are conformally invariant, it is natural work, at least partly, in the setting of conformal manifolds. Here, by a conformal manifold ( M , c ) , we mean a smooth manifold of dimension n 3 equipped with an equivalence class 𝒄 of Riemannian metrics, where g a b , g ̂ a b c means that g ̂ a b = Ω 2 g a b for some smooth positive function Ω. On a general conformal manifold ( M , c ) , there is no distinguished connection on T M . But there is an invariant and canonical connection on a closely related bundle, namely the conformal tractor connection on the standard tractor bundle; see [2, 7].

Here we review the basic conformal tractor calculus on Riemannian and conformal manifolds. See [2, 11, 19] for more details. Unless stated otherwise, calculations will be done with the use of generic g c .

On an 𝑛-manifold 𝑀, the top exterior power of the tangent bundle Λ n T M is a line bundle. Thus its square K : = ( Λ n T M ) 2 is canonically oriented, and so one can take oriented roots of it: given w R , we set

E [ w ] : = K w 2 n

and refer to this as the bundle of conformal densities. For any vector bundle 𝒱, we write V [ w ] to mean V [ w ] : = V E [ w ] . For example, E ( a b ) [ w ] denotes the symmetric second tensor power of the cotangent bundle tensored with E [ w ] , i.e., S 2 T M E [ w ] on 𝑀. On a fixed Riemannian manifold, 𝒦 is canonically trivialised by the square of the volume form, and so 𝒦 and its roots are not usually needed explicitly. However, if we wish to change the metric conformally, or work on a conformal structure, then these objects become important.

Since each metric in a conformal class determines a trivialisation of 𝒦, it follows easily that, on a conformal structure, there is a canonical section g a b Γ ( E ( a b ) ) [ 2 ] . This has the property that, for each positive section σ Γ ( E + [ 1 ] ) (called a scale), g a b : = σ 2 g a b is a metric in 𝒄. Moreover, the Levi-Civita connection of g a b preserves 𝜎 and therefore g a b . Thus it makes sense to use the conformal metric to raise and lower indices, even when we are choosing a particular metric g a b c and its Levi-Civita connection for calculations. It turns out that this simplifies many computations, and so, in this section, we will do that without further mention. (In the subsequent sections, we will work with a fixed metric and use that to trivialise density bundles – so indices will be raised and lowered using the metric.)

Considering Taylor series for sections of E [ 1 ] , one recovers the jet exact sequence at 2-jets,

(2.2) 0 E ( a b ) [ 1 ] ι J 2 E [ 1 ] J 1 E [ 1 ] 0 .

Note that J 2 E [ 1 ] and its sequence (2.2) are canonical objects on any smooth manifold. But, with a conformal structure 𝒄, we have the orthogonal decomposition of E a b [ 1 ] into trace-free and trace parts

E a b [ 1 ] = E ( a b ) 0 [ 1 ] g a b E [ 1 ] .

Thus we can canonically quotient J 2 E [ 1 ] by the image of E ( a b ) 0 [ 1 ] under 𝜄 (in (2.2)). The resulting quotient bundle is denoted T (or E A in abstract indices) and called the conformal cotractor bundle. Observing that the jet exact sequence at 1-jets (of E [ 1 ] ),

0 E b [ 1 ] ι J 1 E [ 1 ] E [ 1 ] 0 ,

we see at once that T has a composition series (or filtration structure)

(2.3)

What this notation means is that E [ 1 ] is a subbundle of T , and the quotient of T by E [ 1 ] (which is J 1 E [ 1 ] ) has E a [ 1 ] as a subbundle, whereas there is a canonical projection X : T E [ 1 ] . In abstract indices, we write X A for this map and call it the canonical tractor.

Given a choice of metric g c , the formula

(2.4) σ 1 n [ D A σ ] g : = ( σ a σ 1 n ( Δ + J ) σ )

(where Δ is the Laplacian a a ) gives a second-order differential operator on E [ 1 ] which is a linear map J 2 E [ 1 ] E [ 1 ] E a [ 1 ] E [ 1 ] that clearly factors through T and so determines an isomorphism

(2.5) T [ T ] g = E [ 1 ] E a [ 1 ] E [ 1 ] .

In subsequent discussions, we will use (2.5) to split the tractor bundles without further comment. Thus, given g c , an element V A of E A may be represented by a triple ( σ , μ a , ρ ) , or equivalently by

(2.6) V A = σ Y A + μ a Z A a + ρ X A .

The last display defines the algebraic splitting operators Y : E [ 1 ] T and Z : T M [ 1 ] T (determined by the choice g a b c ) which may be viewed as sections Y A Γ ( E A [ 1 ] ) and Z A a Γ ( E A a [ 1 ] ) . We call the sections X A , Y A and Z A a tractor projectors.

By construction, the tractor bundle is conformally invariant, i.e., determined by ( M , c ) and independent of any choice of g c . However, the splitting (2.6) is not. Considering the transformation of the operator (2.4) determining the splitting, we see that if g ̂ = Ω 2 f , the components of an invariant section of T should transform according to

(2.7) [ T ] g ̂ ( σ ̂ μ ̂ b ρ ̂ ) = ( 1 0 0 Υ b δ b c 0 1 2 Υ 2 Υ c 1 ) ( σ μ c ρ ) ( σ μ b ρ ) [ T ] g ,

where Υ a = Ω 1 a Ω , and conversely, this transformation of triples is the hallmark of an invariant tractor section. Equivalent to the last display is the rule for how the algebraic splitting operators transform

(2.8) X ̂ A = X A , Z ̂ A b = Z A b + Υ b X A , Y ̂ A = Y A Υ b Z A b 1 2 Υ b Υ b X A .

Given a metric g c , and the corresponding splittings, as above, the tractor connection is given by the formula

(2.9) a T ( σ μ b ρ ) : = ( a σ μ a a μ b + P a b σ + g a b ρ a ρ P a c μ c ) ,

where on the right-hand side the ∇s are the Levi-Civita connection of 𝑔. Using the transformation of components, as in (2.7), and also the conformal transformation of the Schouten tensor,

(2.10) P g ̂ a b = P a b a Υ b + Υ a Υ b 1 2 g a b Υ c Υ c , g ̂ = Ω 2 g ,

reveals that the triple on the right-hand side transforms as a one-form taking values in T , i.e., again by (2.7) except twisted by E a . Thus the right-hand side of (2.9) is the splitting into slots of a conformally invariant connection T on (a section of) the bundle T .

There is a nice conceptual origin for the connection (2.9). Using (2.10) and the transformation of the Levi-Civita connection, it is straightforward to verify that the equation

(2.11) ( a b ) 0 σ + P ( a b ) 0 σ = 0

on conformal densities σ Γ ( E [ 1 ] ) is conformally invariant. As this is an overdetermined PDE, solutions in general do not exist. Overdetermined linear PDEs are typically studied by prolongation, and it is quickly verified that the tractor parallel transport given by (2.9) is exactly the closed system that arises from prolonging (2.11) (see [2, 11]). From this observation and formula (2.9), it follows that non-trivial solutions of (2.11) are non-vanishing on an open dense set (for 𝑀 connected) on which the metric σ 2 g a b is Einstein – an observation that has a number of applications; see [11, 18] and references therein.

Table 1

Tractor inner product

Y A Z A c X A
Y A 0 0 1
Z A b 0 δ b c 0
X A 1 0 0

The tractor bundle is also equipped with a conformally invariant signature ( n + 1 , 1 ) metric h A B Γ ( E ( A B ) ) , defined as quadratic form by the mapping

[ V A ] g = ( σ μ a ρ ) μ a μ a + 2 σ ρ = : h ( V , V ) ,

and the polarisation identity. This is important not only by dint of its conformal invariance, but it is easily checked that this tractor metricℎ is preserved by a T , i.e., a T h A B = 0 . Thus it makes sense to use h A B (and its inverse) to raise and lower tractor indices, and we do this henceforth without further comment. In particular, X A = h A B X B is the canonical tractor (and hence our use of the same kernel symbol). For computations, Table 1 is useful. We see that ℎ may be decomposed into a sum of projections

h A B = Z A a Z B b g a b + X A Y B + Y A X B .

Finally, for this section, we note that, of course, the curvature of the tractor connection κ a b C D is determined by

2 [ a T b ] T V C = κ a b C D V D for all V C Γ ( E A ) ,

and can be written in terms of tractor projectors as

κ a b C D = W a b c d Z C c Z D d + C c a b Z C c X D C c a b X C Z D c .

The bundle A : = Λ 2 T is often termed the adjoint tractor bundle, as it is a vector bundle modelled on the Lie algebra of the conformal group SO ( n + 1 , 1 ) . So, as expected, the tractor curvature is a two-form taking values in this bundle.

2.3 The differential splitting operator

In dimensions n 4 , the tractor curvature is the image of a conformally invariant operator 𝑧 acting on the Weyl curvature. The operator is as follows.

Lemma 2.1

In any dimension n 4 , there is a conformally invariant differential map

(2.12) z : Γ ( W ) Γ ( Λ 2 Λ 2 T ) ,

given by

(2.13) U i j k z i j A B = U i j k Z k A Z B 1 n 3 ( δ U ) k i j ( X A Z k B Z k A X B ) = 1 2 U i j k ( Z k A Z B Z A Z k B ) 1 n 3 ( δ U ) k i j ( X A Z k B Z k A X B ) ,

where ( δ U ) k i j = m U m k i j .

The conformal invariance is easily verified directly using (2.8) and the conformal transformation of the Levi-Civita connection. It can also be deduced from the conformal invariance of the tractor curvature κ a b C D . In fact, the map (2.12) is a standard “BGG-splitting operator”, as in the theory [8, 5], and the conformal deformation sequence can be understood as arising from a twisting by Λ 2 T of the de Rham complex [6, 20].

We are, in particular, interested in the case of dimension n = 4 . Then it is evident, from formula (2.13), that if 𝑈 is SD (or ASD), then so is z ( U ) as a tractor-twisted two-form, as the Hodge-⋆ commutes with the Levi-Civita connection. We obtain more if also U ker D .

Proposition 2.2

Suppose ( M 4 , g ) is ASD, and 𝑈 is SD. Then U ker D if and only if z = z ( U ) Λ + 2 ( A ) is a harmonic, self-dual two-form.

Proof

First note that, by standard BGG theory [8, 5, 15], this is true in the conformally flat setting. Now consider the curved case.

Certainly, δ z is conformally invariant as it is a twisting of the usual divergence of two-forms (in dimension four) with the conformally invariant tractor connection. Thus, starting from the top (meaning from the left in the filtration on the adjoint tractor bundle that is induced from (2.3)), the first non-zero slot of δ z must be conformally invariant and constructed from 𝑈, its covariant derivatives and possibly curvature contracted with 𝑈. From order considerations, the last of these can only happen in the bottom slot. The very top slot is rank 2 and involves no derivatives of 𝑈. Thus it is zero, as 𝑈 is trace-free. At the next level, we have one derivative of 𝑈, but it is well known that, in dimension 4, there is no such conformal invariant.

Thus the image of δ z lies in the bottom slot which contains a rank 2 tensor. Considering the knowledge of conformally flat case, it follows that this conformally invariant object must be a multiple of D U , plus possibly another conformally invariant rank 2 tensor constructed by contracting curvature into 𝑈. It is easily checked that it is not possible to construct a rank 2 conformal invariant by contracting curvature into 𝑈, as 𝑈 is SD, while the Weyl curvature is, by assumption, ASD. Thus the only possibility is that the bottom slot is a non-zero multiple of D U . ∎

We will give another proof of this result (by direct calculation) in the next section; see Corollary 3.3.

Corollary 2.3

Corollary 2.3 (see [3, Theorem 3.10], [23, Lemma 2.1])

Suppose ( M 4 , g ) is ASD and U ker D . Let 𝑧 be defined as in (2.13). Then

Δ z = 1 3 R z = 2 J z .

Proof

𝑧 satisfies a twisted version of the Weitzenböck formula as in [23, Lemma 2.1]. If ( M 4 , g ) is ASD, the only non-zero curvature term is given by 1 3 R . ∎

3 Weitzenböck formula(s)

In this section, we use Proposition 2.2 and Corollary 2.3 to prove a Weitzenböck formula for z = z ( U ) Λ + 2 ( A ) when U ker D . We begin with some more general calculations.

Proposition 3.1

Let U Γ ( W + ) and let z = z ( U ) be given by (2.13). Then the covariant derivative of 𝑧 is given by

(3.1) m z i j A B = ( δ U ) m i j ( X A Y B Y A X B ) + ( m ( δ U ) i j U i j k P m k ) ( X A Z B Z A X B ) + U i j k m ( Y A Z k B Z k A Y B ) + ( 1 2 m U i j k + ( δ U ) k i j δ m ) ( Z k A Z B Z k B Z A ) .

For the proof of this and the following proposition, we will use the following formulas (see [19, (6)]):

(3.2) k X A = Z k A , k Z A = P k X A Y A g k , k Y A = P k Z A .

Lemma 3.2

The following formulas hold:

m ( X A Y B Y A X B ) = ( Y A Z m B Z m A Y B ) + ( P m q X A Z B q P m q X B Z A q ) ,
m ( X A Z k B Z k A Z B ) = g k m ( X A Y B Y A X B ) ( Z k A Z m B Z m A Z k B ) ,
m ( Y A Z k B Z k A Y B ) = P m k ( X A Y B Y A X B ) + P m q ( Z q A Z k B Z q B Z k A ) ,
m ( Z k A Z B Z k B Z A ) = P m k ( X A Z B Z A X B ) + P m ( X A Z k B Z k A X B ) g m k ( Y A Z B Y B Z A ) + g m ( Y A Z k B Z k A Y B ) .

Proof

First, by the Leibniz rule,

m ( X A Y B Y A X B ) = ( m X A ) Y B + X A ( m Y B ) ( m Y A ) X B Y A ( m X B ) .

By (3.2), we find

m ( X A Y B Y A X B ) = Z m A Y B + P m q X A Z B q P m q Z A q X B Y A Z m B = ( Y A Z m B Z m A Y B ) + ( P m q X A Z B q P m q Z A q X B ) .

Similarly,

m ( X A Z k B Z k A X B ) = ( m X A ) Z k B + X A ( m Z k B ) ( m Z k A ) X B Z k A ( m X B ) = Z m A Z k B + X A ( P m k X B Y B g m k ) ( P m k X A Y A g m k ) X B Z k A Z m B = g k m ( X A Y B Y A X B ) ( Z k A Z m B Z m A Z k B ) ,
m ( Y A Z k B Z k A Y B ) = ( m Y A ) Z k B + Y A ( m Z k B ) ( m Z k A ) Y B Z k A ( m Y B ) = P m q Z q A Z k B + Y A ( P m k X B Y B g m k ) ( P m k X A Y A g m k ) Y B P m q Z q B Z k A = P m k ( X A Y B Y A X B ) + P m q ( Z q A Z k B Z q B Z k A ) ,
m ( Z k A Z B Z k B Z A ) = ( m Z k A ) Z B + Z k A ( m Z B ) ( m Z k B ) Z A Z k B ( m Z A ) = ( P m k X A Y A g m k ) Z B + Z k A ( P m X B Y B g m ) ( P m k X B Y B g m k ) Z A Z k B ( P m X A Y A g m ) = P m k ( X A Z B Z A X B ) + P m ( X A Z k B Z k A X B ) g m k ( Y A Z B Y B Z A ) + g m ( Y A Z k B Z k A Y B ) .

Proof of Proposition 3.1

By (2.13),

m z i j A B = m { 1 2 U i j k ( Z k A Z B Z A Z k B ) ( δ U ) k i j ( X A Z k B Z k A X B ) } = 1 2 m U i j k ( Z k A Z B Z A Z k B ) + 1 2 U i j k m ( Z k A Z B Z A Z k B ) m ( δ U ) k i j ( X A Z k B Z k A X B ) ( δ U ) k i j m ( X A Z k B Z k A X B ) = 1 2 m U i j k ( Z k A Z B Z A Z k B ) + 1 2 U i j k { P m k ( X A Z B Z A X B ) + P m ( X A Z k B Z k A X B ) g m k ( Y A Z B Y B Z A ) + g m ( Y A Z k B Z k A Y B ) } m ( δ U ) k i j ( X A Z k B Z k A X B ) ( δ U ) k i j { g k m ( X A Y B Y A X B ) ( Z k A Z m B Z m A Z k B ) } = ( δ U ) m i j ( X A Y B Y A X B ) + ( m ( δ U ) i j U i j k P m k ) ( X A Z B Z A X B ) + U i j k m ( Y A Z k B Z k A Y B ) + ( 1 2 m U i j k + ( δ U ) k i j δ m ) ( Z k A Z B Z k B Z A ) .

We can now give another proof of Proposition 2.2.

Corollary 3.3

Suppose ( M 4 , g ) is ASD, and 𝑈 is SD. Then U ker D if and only if z = z ( U ) Λ + 2 ( A ) is a harmonic, self-dual two-form.

Proof

Since 𝑧 is self-dual, it is harmonic if and only if δ z = 0 . By (3.1),

( δ z ) j A B = g i m m z i j A B = g i m ( δ U ) m i j ( X A Y B Y A X B ) + g i m ( m ( δ U ) i j U i j k P m k ) ( X A Z B Z A X B ) + g i m U i j k m ( Y A Z k B Z k A Y B ) + g i m ( 1 2 m U i j k + ( δ U ) k i j δ m ) ( Z k A Z B Z k B Z A ) .

Since 𝑈 is a curvature-type tensor for which all contractions vanish, it follows that

( δ z ) j A B = 1 2 ( D U ) j ( X A Z B Z A X B )

Proposition 3.4

With the same assumptions as Proposition 3.1, the rough Laplacian of 𝑧, i.e., Δ z i j A B = g k k z i j A B , is given by

Δ z i j A B = { ( W + ) p q i k U q p k j + ( W + ) p q j k U p q i k } ( X A Y B Y A X B ) + { Δ ( δ U ) i j 2 m U i j k P m k U i j k k J + J ( δ U ) i j } × ( X A Z B Z A X B ) + { 1 2 Δ U i j k k ( δ U ) i j + ( δ U ) k i j U i j k m P m + U i j m P m k } × ( Z k A Z B Z k B Z A ) .

Proof

By (3.1),

Δ z i j A B = m ( m z i j A B ) = m { ( δ U ) m i j ( X A Y B Y A X B ) + ( m ( δ U ) i j U i j k P m k ) ( X A Z B Z A X B ) + U i j k m ( Y A Z k B Z k A Y B ) + ( 1 2 m U i j k + ( δ U ) k i j δ m ) ( Z k A Z B Z k B Z A ) } = ( m ( δ U ) m i j ) ( X A Y B Y A X B ) + ( Δ ( δ U ) i j m U i j k P m k U i j k m P m k ) ( X A Z B Z A X B ) + ( m U i j k m ) ( Y A Z k B Z k A Y B ) + ( 1 2 Δ U i j k + ( δ U ) k i j ) ( Z k A Z B Z k B Z A ) + ( δ U ) m i j m ( X A Y B Y A X B ) + ( m ( δ U ) i j U i j k P m k ) m ( X A Z B Z A X B ) + U i j k m m ( Y A Z k B Z k A Y B ) + ( 1 2 m U i j k + ( δ U ) k i j δ m ) m ( Z k A Z B Z k B Z A ) .

Then, applying the formulas from Lemma 3.2 and using the Bianchi identity to write

m P m k = k J ,

we get

Δ z i j A B = ( m ( δ U ) m i j ) ( X A Y B Y A X B )
+ ( Δ ( δ U ) i j m U i j k P m k U i j k k J ) ( X A Z B Z A X B )
+ ( m U i j k m ) ( Y A Z k B Z k A Y B )
+ ( 1 2 Δ U i j k + ( δ U ) k i j ) ( Z k A Z B Z k B Z A )
+ ( δ U ) m i j { ( Y A Z B m Z A m Y B ) + ( P q m X A Z A q P q m Z A q X B ) }
+ ( m ( δ U ) i j U i j k P m k )
× { δ m ( X A Y B Y A X B ) ( Z A Z B m Z A m Z B ) }
+ U i j k m { P k m ( X A Y B Y A X B ) + P q m ( Z q A Z k B Z q B Z k A ) }
+ ( 1 2 m U i j k + ( δ U ) k i j δ m )
× { P k m ( X A Z B Z A X B ) + P m ( X A Z k B Z k A X B )
δ k m ( Y A Z B Y B Z A ) + δ m ( Y A Z k B Z k A Y B ) }
= 2 ( m ( δ U ) m i j ) ( X A Y B Y A X B )
+ { Δ ( δ U ) i j 2 m U i j k P m k U i j k k J + J ( δ U ) i j }
× ( X A Z B Z A X B )
+ { 1 2 Δ U i j k + 2 ( δ U ) k i j + U i j m k P m + U i j m P m k }
× ( Z k A Z B Z k B Z A ) .
Since the very last term is skew-symmetric in k , , we can rewrite this as

Δ z i j A B = 2 ( m ( δ U ) m i j ) ( X A Y B Y A X B ) + { Δ ( δ U ) i j 2 m U i j k P m k U i j k k J + J ( δ U ) i j } × ( X A Z B Z A X B ) + { 1 2 Δ U i j k k ( δ U ) i j + ( δ U ) k i j U i j k m P m + U i j m P m k } × ( Z k A Z B Z k B Z A ) .

The proposition will follow, once we prove the following lemma.

Lemma 3.5

(3.3) m ( δ U ) m i j = 1 2 ( W + ) p q i k U q p k j 1 2 ( W + ) p q j k U q p k i .

In particular, if ( M 4 , g ) is ASD, then m ( δ U ) m i j = 0 .

Proof

By definition of the divergence and the fact that U m i j = U m i j ,

m ( δ U ) m i j = m U m i j = g m p g q p q U m i j = 1 2 g m p g q ( p q q p ) U m i j = 1 2 g m p g q ( R p q k U k m i j + R p q m k U k i j + R p q i k U m k j + R p q j k U m i k ) = 1 2 ( R m k U k m i j + R k U k i j ) + 1 2 g m p g q ( R p q i k U m k j + R p q j k U m i k ) = 1 2 g m p g q ( R p q i k U m k j + R p q j k U m i k ) .

Using the decomposition of the curvature tensor as in (2.1), it follows that

m ( δ U ) m i j = 1 2 g m p g q ( W p q i k + g i p P q k δ p k P i q g i q P p k + δ q k P i p ) U m k j
+ 1 2 g m p g q ( W p q j k + g j p P q k δ p k P j q g j q P p k + δ q k P j p ) U m i k
= 1 2 W p q i k U q p k j + 1 2 ( δ i m P k g k m P i δ i P k m + g k P i m ) U m k j
+ 1 2 W p q j k U q p i k + 1 2 ( δ j m P k g k m P j δ j P k m + g k P j m ) U m i k .
Using the fact that all contractions of 𝑈 vanish, the above simplifies to

m ( δ U ) m i j = 1 2 W p q i k U q p k j + 1 2 W p q j k U q p i k + 1 2 ( P k U i k j P k m U i m k j + P k U j i k P k m U j m i k ) = 1 2 W p q i k U q p k j + 1 2 W p q j k U q p i k = 1 2 W p q i k U q p k j 1 2 W p q j k U q p k i

(note that all terms involving the Schouten tensor can be seen to cancel after re-indexing). Finally, since U Γ ( W + ) , (3.3) follows. ∎

This ends the proof of Proposition 3.4. ∎

Theorem 3.6

If ( M 4 , g ) is ASD and U ker D , then

(3.4) Δ ( δ U ) i j = 2 m U i j k P m k U i j k k J + 3 J ( δ U ) i j ,
(3.5) 1 2 Δ U i j k = k ( δ U ) i j ( δ U ) k i j + U i j k m P m U i j m P k m + J U i j k .

Proof

If U ker D , then z Λ + 2 ( A ) is harmonic. Therefore, (3.4) and (3.5) follow from Proposition 3.4 and Corollary 2.3. ∎

Remark 3.7

Although we will not provide a proof, it is not difficult to show that (3.5) holds for any section U Γ ( W + ) , i.e., the condition U ker D is not necessary. However, this is not the case for (3.4).

4 Proof of Theorem 1.5

Let Q g denote the 𝑄-curvature of ( M 4 , g ) ,

Q g = 1 12 ( Δ g R + R g 2 3 | Ric g | 2 ) = 1 2 Δ g J + J 2 | P | 2 .

The total 𝑄-curvature is a conformal invariant, and we can rewrite (1.4) as a condition on the total 𝑄-curvature and the Yamabe invariant as follows. By the Chern–Gauss–Bonnet formula,

(4.1) Q g d v g = 4 π 2 χ ( M 4 ) 1 8 | W g | 2 d v g .

Since 𝑔 is ASD, (4.1) becomes

(4.2) Q g d v g = 4 π 2 χ ( M 4 ) 1 8 | W g | 2 d v g .

By the Hirzebruch signature formula,

48 π 2 τ ( M 4 ) = ( | W g + | 2 | W g | 2 ) d v g = | W g | 2 d v g .

Combining this with (4.2), we see that

Q g d v g = 2 π 2 ( 2 χ ( M 4 ) + 3 τ ( M 4 ) ) .

Therefore, assumption (1.4) is equivalent to

Q g d v g 1 12 Y ( M 4 , [ g ] ) 2 .

Proof of Theorem 1.5

Since our assumptions are conformally invariant, it suffices to prove the result when 𝑔 is replaced by any metric in the same conformal class. The metric we will use is a solution of a variational problem related to the regularised determinant of an elliptic operator. The precise formulation is contained in Theorem A.2 in the appendix, while the following consequence will suffice for our purposes.

Theorem 4.1

Let ( M 4 , g 0 ) be a closed Riemannian four-manifold with

  1. Y ( M 4 , [ g 0 ] ) > 0 ,

  2. M 4 Q g 0 d v g 0 1 12 Y ( M 4 , [ g 0 ] ) 2 .

Then there is a smooth, unit volume conformal metric g [ g 0 ] with J g = tr P g > 0 satisfying

(4.3) Δ J g | P ̊ g | 2 + 15 4 J g 2 ,

where P ̊ g = P 1 4 J g is the trace-free Schouten tensor.

The proof of this result is somewhat involved, and will also be given in the appendix. In the following, we will show how Theorem 1.5 follows from Theorem 4.1 and the Weitzenböck formulas of the preceding section. For the rest of the proof, the metric 𝑔 is assumed to satisfy the conclusions of Theorem 4.1, and we will usually suppress the subscript 𝑔.

Assume U ker D . We will record two integral identities that follow from Theorem 3.6 and inequality (4.3). First, by (3.5),

(4.4) 1 2 Δ | U | 2 = U , Δ U + | U | 2 = U i j k Δ U i j k + | U | 2 = U i j k { 2 k ( δ U ) i j 2 ( δ U ) k i j + 2 U i j k m P m 2 U i j m P k m + 2 J U i j k } + | U | 2 = 4 U i j k k ( δ U ) i j + 4 U i j k U i j k m P m + 2 J | U | 2 + | U | 2 .

Recalling that

U i j k U i j k m = 1 4 | U | 2 δ m ,

equation (4.4) implies

1 2 Δ | U | 2 = 4 U i j k k ( δ U ) i j + 3 J | U | 2 + | U | 2 .

If we multiply both sides by 𝐽 and integrate over M 4 , then

(4.5) 1 2 M 4 J Δ | U | 2 d v = M 4 { 4 J U i j k k ( δ U ) i j + 3 J 2 | U | 2 + J | U | 2 } d v .

Integrating by parts on the left and using (4.3), we find

(4.6) 1 2 M 4 J Δ | U | 2 d v = 1 2 M 4 ( Δ J ) | U | 2 d v M 4 ( 1 2 | P ̊ | 2 + 15 8 J 2 ) | U | 2 d v .

Combining (4.5) and (4.6),

(4.7) 0 M 4 { 4 J U i j k k ( δ U ) i j + 1 2 | P ̊ | 2 | U | 2 + 9 8 J 2 | U | 2 + J | U | 2 } d v .

We now use (3.4) to get

1 2 Δ | δ U | 2 = | δ U | 2 + δ U , Δ ( δ U ) = | δ U | 2 + ( δ U ) i j Δ ( δ U ) i j = | δ U | 2 + ( δ U ) i j { 2 m U i j k P m k U i j k k J + 3 J ( δ U ) i j } = | δ U | 2 2 ( δ U ) i j m U i j k P m k ( δ U ) i j U i j k k J + 3 J | δ U | 2 .

Integrating this over M 4 gives

(4.8) 0 = M 4 { | δ U | 2 2 ( δ U ) i j m U i j k P m k ( δ U ) i j U i j k k J + 3 J | δ U | 2 } d v .

If we integrate by parts in the second term and use the contracted second Bianchi identity m P m k = k J , then

M 4 2 ( δ U ) i j m U i j k P m k d v g = M 4 { 2 m ( δ U ) i j U i j k P m k + 2 ( δ U ) i j U i j k m P m k } d v g = M 4 { 2 m ( δ U ) i j U i j k P m k + 2 ( δ U ) i j U i j k k J } d v g .

Substituting this result back into (4.8) gives

(4.9) 0 = M 4 { | δ U | 2 + 2 m ( δ U ) i j U i j k P m k + U i j k ( δ U ) i j k J + 3 J | δ U | 2 } d v .

Next, integrate by parts in the second-to-last term in (4.9) to get

M 4 U i j k ( δ U ) i j k J d v = M 4 { U i j k k ( δ U ) i j J k U i j k ( δ U ) i j J } d v = M 4 { U i j k k ( δ U ) i j J J | δ U | 2 } d v .

Substituting this back into (4.9) gives

0 = M 4 { | δ U | 2 + 2 m ( δ U ) i j U i j k P m k J U i j k k ( δ U ) i j + 2 J | δ U | 2 } d v .

Finally, we rewrite the curvature terms above in terms of P ̊ and 𝐽 as follows:

0 = M 4 { | δ U | 2 + 2 m ( δ U ) i j U i j k P ̊ m k 1 2 J U i j k k ( δ U ) i j + 2 J | δ U | 2 } d v .

Multiplying by two and rearranging terms, we find

(4.10) M 4 J U i j k k ( δ U ) i j d v = M 4 { 2 | δ U | 2 + 4 m ( δ U ) i j U i j k P ̊ m k + 4 J | δ U | 2 } d v .

We want to combine (4.10) with (4.7). To do so, we first use (4.10) to write

M 4 4 J U i j k k ( δ U ) i j d v = 3 M 4 J U i j k k ( δ U ) i j d v + M 4 J U i j k k ( δ U ) i j d v = M 4 3 J U i j k k ( δ U ) i j d v + M 4 { 2 | δ U | 2 + 4 m ( δ U ) i j U i j k P ̊ m k + 4 J | δ U | 2 } d v = M 4 { 2 | δ U | 2 + 4 m ( δ U ) i j U i j k P ̊ m k + 3 J U i j k k ( δ U ) i j + 4 J | δ U | 2 } d v .

Substituting this into (4.7), we have

(4.11) 0 M 4 { 2 | δ U | 2 + 4 m ( δ U ) i j U i j k P ̊ m k + 3 J U i j k k ( δ U ) i j + 4 J | δ U | 2 + 1 2 | P ̊ | 2 | U | 2 + 9 8 J 2 | U | 2 + J | U | 2 } d v = M 4 { 2 | δ U | 2 + 4 T m k m ( δ U ) i j U i j k + 4 J | δ U | 2 + 1 2 | P ̊ | 2 | U | 2 + 9 8 J 2 | U | 2 + J | U | 2 } d v ,

where

T k m = P ̊ m k + 3 4 J g k m .

If we define the tensor

V m i j = T m k U i j k ,

then the term involving 𝑇 in (4.11) can be estimated (via Cauchy–Schwarz) by

| 4 T m k m ( δ U ) i j U i j k | = 4 | m ( δ U ) i j V m i j | 4 | δ U | | V | .

By the arithmetic-geometric mean inequality,

(4.12) | 4 T k m m ( δ U ) i j U i j k | 2 | δ U | 2 + 2 | V | 2 .

By the definition of 𝑉,

| V | 2 = V m i j V m i j = T p m U i j p T m k U i j k = T p m T m k U i j p U i j k = T p m T m k ( 1 4 | U | 2 δ k p ) = 1 4 | T | 2 | U | 2 = 1 4 ( | P ̊ | 2 + 9 4 J 2 ) | U | 2 .

Consequently, (4.12) implies

| 4 T k m m ( δ U ) i j U i j k | 2 | δ U | 2 + 1 2 | P ̊ | 2 | U | 2 + 9 8 J 2 | U | 2 ;

hence

4 T k m m ( δ U ) i j U i j k 2 | δ U | 2 1 2 | P ̊ | 2 | U | 2 9 8 J 2 | U | 2 .

Substituting this into (4.11), we conclude

0 M 4 { 4 J | δ U | 2 + J | U | 2 } d v .

Since J > 0 , it follows that U = 0 . However, it is then immediate from (4.7) that U 0 . ∎

Funding source: Marsden Fund

Award Identifier / Grant number: 19-UOA-008

Award Identifier / Grant number: DMS-2105460

Funding source: Simons Foundation

Award Identifier / Grant number: 923208

Funding statement: The first author is supported by the Royal Society of New Zealand Marsden grant 19-UOA-008. The second author is supported by NSF grant DMS-2105460 and a Simons Foundation Fellowship in Mathematics, award 923208.

A Appendix: The proof of Theorem 4.1

In this appendix, we give the proof of Theorem 4.1. The material in this section is an extension of the existence work of Chang–Yang [10] for critical points of the regularised determinant of conformally covariant operators. We begin with a brief overview of their work, omitting the motivation from spectral geometry and limiting ourselves to the underlying variational problem.

Let ( M 4 , g ) be a closed, four-dimensional Riemannian manifold, and W 2 , 2 ( M ) the Sobolev space of functions whose weak derivatives up to order two are in L 2 . Consider the following functionals on W 2 , 2 ( M ) :

I ± [ w ] = 4 w | W g ± | 2 d v g ( | W g ± | 2 d v g ) log e 4 w d v g , II [ w ] = w P g w d v g + 4 Q g w d v g ( Q g d v g ) log e 4 w d v g , III [ w ] = 12 ( Δ w + | w | 2 ) 2 d v g 4 ( w Δ R g + R g | w | 2 ) d v g ,

where P g denotes the Paneitz operator, Q g the scalar curvature, and denotes the normalised integral (i.e., divided by the volume of 𝑔). Let γ 1 ± , γ 2 , γ 3 be constants and let F : W 2 , 2 ( M ) R be given by

F [ w ] = γ 1 + I + [ w ] + γ 1 I [ w ] + γ 2 II [ w ] + γ 3 III [ w ] .

We also define the associated conformal invariant

κ g = γ 1 + | W + | 2 d v g γ 1 | W | 2 d v g γ 2 Q g d v g .

As explained in [10], critical points of 𝐹 determine a conformal metric satisfying a fourth order curvature condition. More precisely, if we define the 𝑈-curvature of 𝑔 by

U = U ( g ) = γ 1 + | W g + | 2 + γ 1 | W g | 2 + γ 2 Q g γ 3 Δ g R g ,

then 𝑤 is a smooth critical point of 𝐹 if and only if the conformal metric g F = e 2 w g satisfies U ( g F ) μ for some constant 𝜇. A general existence result for critical points of 𝐹 was proved in [10].

Theorem A.1

Theorem A.1 (see [10, Theorem 1.1], also [21, Corollary 1.1])

Assume

  1. γ 2 < 0 and γ 3 < 0 .

  2. κ g < ( γ 2 ) 8 π 2 .

Then sup w W 2 , 2 ( M ) F [ w ] is attained by some w W 2 , 2 .

Regularity of extremals was proved by the second author in joint work with Chang–Yang [9]; later Uhlenbeck–Viaclovsky proved a more general regularity result for arbitrary critical points of 𝐹 (see [39]).

To prove Theorem 4.1, we need to introduce another functional

IV [ w ] = ( R g + 6 | w | 2 ) e 2 w d v g ( e 4 w d v g ) 1 / 2 .

This is just the Yamabe functional, written in a slightly non-standard form: if g ̃ = e 2 w g , then

(A.1) IV [ w ] = R g ̃ d v g ̃ Vol ( g ̃ ) 1 / 2 .

Given a constant γ 4 , we define Φ : W 2 , 2 ( M ) R by

Φ [ w ] = γ 1 + I + [ w ] + γ 1 I [ w ] + γ 2 II [ w ] + γ 3 III [ w ] + γ 4 IV [ w ] .

A trivial modification of the existence result of Chang–Yang and the regularity results of Chang–Gursky–Yang and Uhlenbeck–Viaclovsky gives the following.

Theorem A.2

Assume

  1. γ 2 < 0 , γ 3 < 0 , γ 4 0 .

  2. κ g < ( γ 2 ) 8 π 2 .

Then sup w W 2 , 2 ( M ) Φ [ w ] is attained by some w W 2 , 2 ( M ) . Moreover, w C , and the conformal metric g ̃ = e 2 w g satisfies

γ 1 + | W g ̃ + | 2 + γ 1 | W g ̃ | 2 + γ 2 Q g ̃ γ 3 Δ g ̃ R g ̃ + 1 2 γ 4 R g ̃ = μ

for some constant 𝜇.

Proof

Note that the functional Φ is scale-invariant: Φ [ w + c ] = Φ [ w ] for any constant 𝑐. Therefore, we may normalise a maximising sequence { w k } for Φ so that

e 4 w k d v g = 1 .

Assuming R g C , it follows from the Schwartz inequality that

IV [ w ] = ( R g + 6 | w k | 2 ) e 2 w k d v g R g e 2 w k d v g C ( e 4 w k d v g ) 1 / 2 C .

Consequently, if γ 4 0 , then γ 4 IV [ w ] is bounded above. By (A.1), it is also bounded below (by the Yamabe invariant). Therefore, the addition of this term has no effect on the estimates in the existence proof of Chang–Yang. ∎

We are now ready to prove Theorem 4.1.

Proof of Theorem 4.1

We first remark that if ( M 4 , g 0 ) is conformally equivalent to the round sphere (suitably normalised), then g = g c satisfies the conclusions of the theorem. Therefore, we may assume ( M 4 , g 0 ) is not conformally the round sphere.

Taking

γ 1 ± = 0 , γ 2 = 6 , γ 3 = 1 2 , γ 4 = 2 Y ( M 4 , [ g 0 ] ) ,

then

κ g = 6 Q g 0 d v g 0 .

To use Theorem A.2, we need to verify assumption (ii), i.e.,

(A.2) Q g 0 d v g 0 < 8 π 2 .

By [22, Theorem B], (A.2) holds as long as ( M 4 , g 0 ) is not conformally equivalent to the round sphere. Therefore, by Theorem A.2, there is a smooth conformal metric g = e 2 w g 0 (which we can normalise to have unit volume) satisfying

(A.3) 6 Q g + 1 2 Δ g R g Y ( M 4 , [ g 0 ] ) R g = μ .

For the rest of the proof, we will omit the subscript 𝑔. If E = Ric 1 4 R g denotes the trace-free Ricci tensor of 𝑔, then we may use the definition of the 𝑄-curvature to rewrite (A.3) as

(A.4) Δ R = 1 8 R 2 3 2 | E | 2 + Y ( M 4 , [ g 0 ] ) R + μ .

By the arithmetic-geometric mean inequality,

Y ( M 4 , [ g 0 ] ) R 1 2 R 2 + 1 2 Y ( M 4 , [ g 0 ] ) 2 .

Therefore,

(A.5) Δ R 5 8 R 2 3 2 | E | 2 + 1 2 Y ( M 4 , [ g 0 ] ) 2 + μ .

Claim A.3

(A.6) μ + 1 2 Y ( M 4 , [ g 0 ] ) 2 0 .

For now, let us assume the claim and see how the theorem follows. From (A.6) and (A.5), it follows that

Δ R 5 8 R 2 3 2 | E | 2 .

Using the fact that J = 1 6 R and P ̊ = 1 2 E , this inequality can also be written

Δ J | P ̊ | 2 + 15 4 J 2 ;

hence (4.3) holds.

To see that R > 0 , we use (A.4) and the fact that μ < 0 to write

(A.7) Δ R 1 8 R 2 + Y ( M 4 , [ g 0 ] ) R 1 6 R 2 + Y ( M 4 , [ g 0 ] ) R .

Let ϕ > 0 denote the eigenfunction associated to the first eigenvalue λ 1 ( L ) of the conformal Laplacian

(A.8) L ϕ : = ( 6 Δ + R ) ϕ = λ 1 ( L ) .

Since Y ( M 4 , [ g ] ) > 0 , it follows that λ 1 ( L ) > 0 . An easy calculation using (A.7) and (A.8) gives

Δ R ϕ 2 ( R ϕ ) , ϕ ϕ + ( Y ( M 4 , [ g 0 ] ) + 1 6 λ 1 ( L ) ) R ϕ .

It follows from the strong maximum principle that R / ϕ > 0 on 𝑀; hence R > 0 . This completes the proof of the theorem, once we prove Claim A.3. ∎

Proof of Claim A.3

If we integrate (A.3) over 𝑀 and use the fact that 𝑔 has unit volume, we obtain

(A.9) μ = 6 Q g d v g Y ( M 4 , [ g 0 ] ) R g d v g .

By definition of the Yamabe invariant (again using the fact that 𝑔 has unit volume) and the fact that Y ( M 4 , [ g 0 ] ) > 0 ,

Y ( M 4 , [ g 0 ] ) R g d v g Y ( M 4 , [ g 0 ] ) 2 .

Therefore, by (A.9),

μ 6 Q g d v g Y ( M 4 , [ g 0 ] ) 2 .

Since the total 𝑄-curvature is a conformal invariant, using assumption (ii) of the theorem, we see that

μ 6 Q g d v g Y ( M 4 , [ g 0 ] ) 2 = 6 Q g 0 d v g 0 Y ( M 4 , [ g 0 ] ) 2 1 2 Y ( M 4 , [ g 0 ] ) 2 Y ( M 4 , [ g 0 ] ) 2 = 1 2 Y ( M 4 , [ g 0 ] ) 2 ,

which proves (A.6). ∎

Acknowledgements

The authors would like to express their sincere thanks to Claude LeBrun for initially suggesting this problem, and for being an invaluable resource during our work. This project began during a visit of the second author to the Department of Mathematics at the University of Auckland, in September 2022. The author would like to thank the department for its support and hospitality.

References

[1] M. F. Atiyah, N. J. Hitchin and I. M. Singer, Self-duality in four-dimensional Riemannian geometry, Proc. Roy. Soc. Lond. Ser. A 362 (1978), no. 1711, 425–461. 10.1098/rspa.1978.0143Search in Google Scholar

[2] T. N. Bailey, M. G. Eastwood and A. R. Gover, Thomas’s structure bundle for conformal, projective and related structures, Rocky Mountain J. Math. 24 (1994), no. 4, 1191–1217. 10.1216/rmjm/1181072333Search in Google Scholar

[3] J.-P. Bourguignon and H. B. Lawson, Jr., Stability and isolation phenomena for Yang–Mills fields, Comm. Math. Phys. 79 (1981), no. 2, 189–230. 10.1007/BF01942061Search in Google Scholar

[4] T. P. Branson and B. Ørsted, Explicit functional determinants in four dimensions, Proc. Amer. Math. Soc. 113 (1991), no. 3, 669–682. 10.2307/2048601Search in Google Scholar

[5] D. M. J. Calderbank and T. Diemer, Differential invariants and curved Bernstein–Gelfand–Gelfand sequences, J. reine angew. Math. 537 (2001), 67–103. 10.1515/crll.2001.059Search in Google Scholar

[6] A. Čap, Infinitesimal automorphisms and deformations of parabolic geometries, J. Eur. Math. Soc. (JEMS) 10 (2008), no. 2, 415–437. 10.4171/jems/116Search in Google Scholar

[7] A. Čap and A. R. Gover, Tractor calculi for parabolic geometries, Trans. Amer. Math. Soc. 354 (2002), no. 4, 1511–1548. 10.1090/S0002-9947-01-02909-9Search in Google Scholar

[8] A. Čap, J. Slovák and V. Souček, Bernstein–Gelfand–Gelfand sequences, Ann. of Math. (2) 154 (2001), no. 1, 97–113. 10.2307/3062111Search in Google Scholar

[9] S.-Y. A. Chang, M. J. Gursky and P. C. Yang, Regularity of a fourth order nonlinear PDE with critical exponent, Amer. J. Math. 121 (1999), no. 2, 215–257. 10.1353/ajm.1999.0011Search in Google Scholar

[10] S.-Y. A. Chang and P. C. Yang, Extremal metrics of zeta function determinants on 4-manifolds, Ann. of Math. (2) 142 (1995), no. 1, 171–212. 10.2307/2118613Search in Google Scholar

[11] S. N. Curry and A. R. Gover, An introduction to conformal geometry and tractor calculus, with a view to applications in general relativity, Asymptotic analysis in general relativity, London Math. Soc. Lecture Note Ser. 443, Cambridge University, Cambridge (2018), 86–170. 10.1017/9781108186612.003Search in Google Scholar

[12] A. Derdziński, Self-dual Kähler manifolds and Einstein manifolds of dimension four, Compos. Math. 49 (1983), no. 3, 405–433. Search in Google Scholar

[13] S. Donaldson and R. Friedman, Connected sums of self-dual manifolds and deformations of singular spaces, Nonlinearity 2 (1989), no. 2, 197–239. 10.1088/0951-7715/2/2/002Search in Google Scholar

[14] M. G. Eastwood and M. A. Singer, On the geometry of twistor spaces, unpublished manuscript, 1991. Search in Google Scholar

[15] M. Eastwood and J. Slovák, Semiholonomic Verma modules, J. Algebra 197 (1997), no. 2, 424–448. 10.1006/jabr.1997.7136Search in Google Scholar

[16] T. Eguchi, P. B. Gilkey and A. J. Hanson, Gravitation, gauge theories and differential geometry, Phys. Rep. 66 (1980), no. 6, 213–393. 10.1016/0370-1573(80)90130-1Search in Google Scholar

[17] A. Floer, Self-dual conformal structures on l C P 2 , J. Differential Geom. 33 (1991), no. 2, 551–573. 10.4310/jdg/1214446330Search in Google Scholar

[18] A. R. Gover, Almost Einstein and Poincaré–Einstein manifolds in Riemannian signature, J. Geom. Phys. 60 (2010), no. 2, 182–204. 10.1016/j.geomphys.2009.09.016Search in Google Scholar

[19] A. R. Gover and L. J. Peterson, Conformally invariant powers of the Laplacian, 𝑄-curvature, and tractor calculus, Comm. Math. Phys. 235 (2003), no. 2, 339–378. 10.1007/s00220-002-0790-4Search in Google Scholar

[20] A. R. Gover and L. J. Peterson, The ambient obstruction tensor and the conformal deformation complex, Pacific J. Math. 226 (2006), no. 2, 309–351. 10.2140/pjm.2006.226.309Search in Google Scholar

[21] M. J. Gursky, The Weyl functional, de Rham cohomology, and Kähler–Einstein metrics, Ann. of Math. (2) 148 (1998), no. 1, 315–337. 10.2307/120996Search in Google Scholar

[22] M. J. Gursky, The principal eigenvalue of a conformally invariant differential operator, with an application to semilinear elliptic PDE, Comm. Math. Phys. 207 (1999), no. 1, 131–143. 10.1007/s002200050721Search in Google Scholar

[23] M. J. Gursky, C. L. Kelleher and J. Streets, A conformally invariant gap theorem in Yang–Mills theory, Comm. Math. Phys. 361 (2018), no. 3, 1155–1167. 10.1007/s00220-017-3070-zSearch in Google Scholar

[24] M. Itoh, Moduli of half conformally flat structures, Math. Ann. 296 (1993), no. 4, 687–708. 10.1007/BF01445130Search in Google Scholar

[25] M. Itoh, The Weitzenböck formula for the Bach operator, Nagoya Math. J. 137 (1995), 149–181. 10.1017/S0027763000005109Search in Google Scholar

[26] D. D. Joyce, Explicit construction of self-dual 4-manifolds, Duke Math. J. 77 (1995), no. 3, 519–552. 10.1215/S0012-7094-95-07716-3Search in Google Scholar

[27] M. Kalafat, Scalar curvature and connected sums of self-dual 4-manifolds, J. Eur. Math. Soc. (JEMS) 13 (2011), no. 4, 883–898. 10.4171/jems/269Search in Google Scholar

[28] M. Kalafat, Y. Ozan and H. Argüz, Self-dual metrics on non-simply connected 4-manifolds, J. Geom. Phys. 64 (2013), 79–82. 10.1016/j.geomphys.2012.08.005Search in Google Scholar

[29] A. D. King and D. Kotschick, The deformation theory of anti-self-dual conformal structures, Math. Ann. 294 (1992), no. 4, 591–609. 10.1007/BF01934343Search in Google Scholar

[30] O. Kobayashi, On a conformally invariant functional of the space of Riemannian metrics, J. Math. Soc. Japan 37 (1985), no. 3, 373–389. 10.2969/jmsj/03730373Search in Google Scholar

[31] O. Kobayashi, Scalar curvature of a metric with unit volume, Math. Ann. 279 (1987), no. 2, 253–265. 10.1007/BF01461722Search in Google Scholar

[32] C. LeBrun, On the topology of self-dual 4-manifolds, Proc. Amer. Math. Soc. 98 (1986), no. 4, 637–640. 10.1090/S0002-9939-1986-0861766-2Search in Google Scholar

[33] C. LeBrun, Explicit self-dual metrics on C P 2 # # C P 2 , J. Differential Geom. 34 (1991), no. 1, 223–253. 10.4310/jdg/1214446999Search in Google Scholar

[34] C. LeBrun, Twistors, Kähler manifolds, and bimeromorphic geometry. I, J. Amer. Math. Soc. 5 (1992), no. 2, 289–316. 10.1090/S0894-0347-1992-1137098-5Search in Google Scholar

[35] Y. S. Poon, Compact self-dual manifolds with positive scalar curvature, J. Differential Geom. 24 (1986), no. 1, 97–132. 10.4310/jdg/1214440260Search in Google Scholar

[36] Y. S. Poon, On the algebraic structure of twistor spaces, J. Differential Geom. 36 (1992), no. 2, 451–491. 10.4310/jdg/1214448749Search in Google Scholar

[37] I. M. Singer and J. A. Thorpe, The curvature of 4-dimensional Einstein spaces, Global analysis (Papers in honor of K. Kodaira), Universtiy of Tokyo, Tokyo (1969), 355–365. 10.1515/9781400871230-021Search in Google Scholar

[38] C. H. Taubes, The existence of anti-self-dual conformal structures, J. Differential Geom. 36 (1992), no. 1, 163–253. 10.4310/jdg/1214448445Search in Google Scholar

[39] K. K. Uhlenbeck and J. A. Viaclovsky, Regularity of weak solutions to critical exponent variational equations, Math. Res. Lett. 7 (2000), no. 5–6, 651–656. 10.4310/MRL.2000.v7.n5.a11Search in Google Scholar

[40] J. A. Viaclovsky, Critical metrics for Riemannian curvature functionals, Geometric analysis, IAS/Park City Math. Ser. 22, American Mathematical Society, Providence (2016), 197–274. 10.1090/pcms/022/05Search in Google Scholar

Received: 2023-10-02
Revised: 2024-04-17
Published Online: 2024-05-16
Published in Print: 2024-06-01

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 15.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/crelle-2024-0028/html
Scroll to top button