NON-SINGULAR SPACETIMES WITH A NEGATIVE COSMOLOGICAL CONSTANT: III. STATIONARY SOLUTIONS WITH MATTER FIELDS

arXiv:1701.03718v1 [gr-qc] 13 Jan 2017

´ PIOTR T. CHRUSCIEL, ERWANN DELAY, AND PAUL KLINGER Abstract. We construct infinite-dimensional families of non-singular stationary space times, solutions of Yang-Mills-Higgs-Einstein-MaxwellChern-Simons-dilaton-scalar field equations with a negative cosmological constant. The families include an infinite-dimensional family of solutions with the usual AdS conformal structure at conformal infinity.

Contents 1. Introduction 2. Stationary Einstein-Maxwell-Chern-Simons-dilaton-scalar field equations in n + 1 dimensions 3. Definitions, notations and conventions 4. The method 5. Static metrics 5.1. Purely electric configurations 5.2. Purely magnetic configurations 5.3. Yang-Mills-Higgs fields 5.4. Time-periodic scalar fields 6. Stationary metrics 6.1. Time-periodic scalar fields 7. Asymptotics, energy Appendix A. The Ω equation Appendix B. The Christoffels Appendix C. AdS, Hn and a quotient References

1 3 4 4 6 7 10 11 13 13 17 18 18 20 21 22

1. Introduction There is currently considerable interest in the literature in space-times with a negative cosmological constant. This is fueled on one hand by studies of the AdS-CFT conjecture and of the implications thereof. On the other hand, these solutions are interesting because of a rich dynamical morphology: existence of periodic or quasi-periodic solutions, and of instabilities. All this leads naturally to the question of existence of stationary solutions of the Einstein equations with Λ < 0, with or without sources, and of properties Date: January 16, 2017. Preprint UWThPh-2016-30. 1

2

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

thereof. Several families of such solutions have been recently constructed numerically [5, 9, 11, 18, 25]. In an accompanying paper [15] two of us (PTC and ED) have constructed an infinite dimensional family of non-singular static space times, solutions of the Einstein-Maxwell equations with a negative cosmological constant. These families include an infinite-dimensional family of solutions with the usual AdS conformal structure at conformal infinity. The object of this work is to generalise the construction there to obtain a similar large family of singularity-free stationary solutions of the Yang-Mills-Higgs-EinsteinMaxwell-dilaton-Chern-Simons-scalar field equations, including a class of boson-star solutions with stationary metric but periodic complex scalar field. Note that existence of such solutions of the Einstein-Maxwell-dilaton field equations with the Kaluza-Klein value of the coupling constant is a special case of the results in [4]. More precisely, we construct strictly stationary solutions of the Einsteinmatter field equations with a negative cosmological constant and with a smooth conformal boundary for a large class of matter models. Here we say that a space-time (M , g) is strictly stationary if there exists on M a Killing vector field which is timelike everywhere. Such a solution is defined to be non-degenerate if a certain operator associated with the linearisation of the field equations is an isomorphism, cf. Section 3 for a precise definition. An example of a non-degenerate solution is anti-de Sitter space-time. Our solutions are constructed using an implicit function theorem near a nondegenerate vacuum metric (M ,˚ g). We also construct solutions with a timeperiodic complex scalar field accompanied by time-independent metric and Maxwell fields. The solutions are uniquely determined by certain freely prescribable coefficients in the asymptotic expansion of the metric, of the Yang-Mills or Maxwell fields, of the dilaton field and of the scalar fields. Here uniqueness is guaranteed in a neighborhood of the metric ˚ g. In this way we obtain infinite dimensional families of solutions with, if desired, the same conformal structure at infinity as the initial static vacuum metric ˚ g. By switching-off some free data at the conformal boundary, or setting to zero one of the coupling constants, one can obtain non-trivial solutions of the Einstein-Yang-Mills equations, or of the Einstein-scalar field equations, or of static Einstein-Maxwell-Chern-Simons-dilaton solutions, etc. In particular, we establish rigorously existence of the Einstein-Yang-Mills solutions discovered numerically in [9] for “weak” Yang-Mills fields, and in fact we provide a much larger family of such solutions. The method is a rather straightforward repetition of the arguments in [14, 15], so that our presentation will be suitably sketchy: we will only provide details at places which require technical or calculational changes. Our hypothesis of strict stationarity excludes black hole solutions. The extension of our analysis to black holes will be discussed elsewhere [16]. Similar construction works near any non-degenerate stationary solutions of the equations under considerations, provided that the linearisations of the matter equations also lead to isomorphisms. This appears to require a case-by-case analysis of the solutions at hand.

STATIONARY SPACETIMES WITH NEGATIVE Λ

3

2. Stationary Einstein-Maxwell-Chern-Simons-dilaton-scalar field equations in n + 1 dimensions We consider the Einstein equations for a metric g = gµν dxµ dxν in space-time dimension n + 1, n ≥ 3, Trg Ric(g) (2.1) Ric(g) − g + Λg = 8πGT , 2 where T is the energy-momentum tensor of matter fields. A constant rescaling of g allows one to normalise a negative cosmological constant to n(n − 1) , (2.2) Λ=− 2 and we will often use this normalisation. The space-time manifold M will be taken of the form R × M , with the R coordinate running along the orbits of a Killing vector field which is timelike everywhere. In the Einstein-Maxwell-Chern-Simons-dilaton-scalar field case we have [10, 22] 1 1 Tαβ = (2.3) ∂α φ∂β φ + 2W (φ)Fαµ Fβ µ 8πG 2  W (φ) 2 1 1 |F | + V (φ) , − gαβ (∇φ)2 + 4 2 2 with action (2.4)   Z p 1 1 2 n+1 2 S= d x − det g R(g) − W (φ)|F | − (∇φ) − V (φ) + SCS , 16πG 2 where R(g) is the Ricci scalar of the metric g and where, in odd space-time dimensions, SCS is the Abelian Chern-Simons action:   n is odd;  0, Z λ (2.5) SCS = A∧F · · ∧ F}, n = 2k,  | ∧ ·{z  16πG k times

for a constant λ ∈ R. We will assume that W and V are smooth functions, and require (2.6)

W (0) = 1 ,

V (0) = 0

(note that this differs from the conventions of the accompanying paper [15], where φ ≡ 0 and where the normalisation W ≡ 1/2 has been used). We can view φ as taking values in a Euclidean RN +1 for some N ≥ 0, with the first component φ1 corresponding to the dilaton field, and with W depending only upon φ1 . Then the remaining components (φ2 , . . . , φN +1 ) of φ describe N minimally-coupled scalar fields, possibly interacting with each other through the potential V which might or might not depend upon φ1 . Taking N = 0, V ≡ 0, φ = 2u, W (u) = e−2au for a constant a ∈ R, and setting the Chern-Simons coupling constant λ to zero one obtains the usual

4

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

Einstein-Maxwell-dilaton equations with action [20] Z p   1 (2.7) S= dn+1 x − det g R − e−2au |F |2 − 2(∇u)2 . 16πG

Similarly, we can view F as taking values in a Euclidean RN1 for some N1 ≥ 0, in which case we obtain a collection of Abelian Yang-Mills fields B dxµ ∧ dxν , B = 1, . . . , N . The Chern-Simons action (2.5) can then be Fµν 1 replaced by   0, n is odd, Z 1 (2.8) SCS = λBB1 ...Bk AB ∧ F B1 ∧ · · · ∧ F Bk , n = 2k,  16πG for a set of constants λBB1 ...Bk , totally symmetric in the last k indices. Our analysis extends to general Yang-Mills-Higgs-dilaton-Chern-Simons fields in the obvious way, by replacing ∂φ by a gauge-covariant derivative. This is addressed in Section 5.3 below. 3. Definitions, notations and conventions Our definitions and conventions are identical to those in [14, Section 2]. Recall that the linearisation of the Ricci tensor in dimension (n + 1) equals [7, Equations (1.180a)-(1.180b), p. 64]  1 ∆L h − 2δ∗ δh − Dd(trh) , (3.1) 2 or in index notation 1 ∆L hij + Di D k hkj + Dj D k hki − Di Dj hk k ) , 2 where ∆L h is the Lichnerowicz Laplacian acting on the symmetric twotensor field h, defined as [7, § 1.143] ∆L hij = −∇k ∇k hij + Rik hk j + Rjk hk i − 2Rikjl hkl .

An explicit form of ∆L for a metric of the form (4.1) below can be read-off from the formulae in [14, Appendix A]. We will say that a metric g is non-degenerate if ∆L + 2n has no L2 -kernel. Large classes of non-degenerate Einstein metrics are described in [1,2,4,23]. 4. The method We seek to construct Lorentzian metrics g in any space-dimension n ≥ 3, with Killing vector X = ∂/∂t. In adapted coordinates those metrics can be written as (4.1)

g = −V 2 (dt + θi dxi )2 + gij dxi dxj , | {z } | {z } =θ

(4.2)

=g

∂t V = ∂t θ = ∂t g = 0 .

Let us denote by ϕ = (ϕa ) all matter fields, where the index a runs over some index set {1, . . . , Nm }, 1 ≤ Nm < ∞. The ϕa ’s will be required to satisfy (4.3)

∂t ϕ = 0

STATIONARY SPACETIMES WITH NEGATIVE Λ

5

in the coordinate system of (4.1), except in Section 5.4, respectively Section 6.1, where time-periodic matter field configurations are considered with static, respectively stationary, metrics. In the case of the action (2.4) we thus have ϕ = (Aµ , φa ), but the overall argument applies to more general systems as long as the energy-momentum tensor is at least quadratic in the fields. Consider the Einstein equations (2.1), in space-time dimension n + 1, with a cosmological constant Λ. We impose (4.1)-(4.3), and assume that the energy-momentum tensor T does not depend upon more than one derivative of g. We further suppose that (4.4) whenever the matter field equations are satisfied we have ∇µ T µ ν = 0, regardless of whether or not the metric g satisfies (2.1). The following approach has become standard since [19] in this kind of problems, we review the method for completeness. We assume that the fields ϕ satisfy a system of equations of the form L(g, ∂g, ∂ 2 g)[ϕ] = 0 ,

(4.5)

where L is a partial-differential operator acting on ϕ with coefficients which might depend upon g and its derivatives up to order two. In order to obtain an elliptic system of equations for (V, θ, g) we replace (2.1) by (4.6) R(g)µν −

R(g) gµν + Λgµν + ∇µ Ων + ∇ν Ωµ − ∇α Ωα gµν = 8πGTµν , 2

where (4.7)

Ων

:= gνµ gαβ (Γ(˚ g)µαβ − Γ(g)µαβ )

(cf. Appendix B below). Assume that (4.5)-(4.6) can be solved for (V, θ, g, ϕ). The Bianchi identity and (4.4) imply that (4.8)

∇µ (∇µ Ων + ∇ν Ωµ − ∇α Ωα gµν ) = 0 .

We show in Appendix A that this equation implies Ω ≡ 0 whenever |Ω|g = o(ρ−1 ) (as will be the case for our solutions), and consequently we will obtain the desired solution of the original equations. For the Einstein-scalar field equations this is the end of the story, provided we can construct solutions of the system (4.5)-(4.6). This will be done using an implicit function theorem around the solution φ ≡ 0, g = ˚ g. Now, the associated linearised equations are rather complicated, in particular the question of invertibility of the linearisation of the modified Einstein equations is not a trivial issue. This is solved in [3, 4, 14, 15] by the following artefact, which we apply again here: It is well known that the Einstein tensor for a Riemannian metric g = V 2 dt2 + g on S 1 ×M coincides with the Einstein tensor of g = −V 2 dt2 +g. This implies that the isomorphism property of the linearised operator for the “harmonically reduced Riemannian Einstein equations”, at a static solution, carries over to the Lorentzian equations; compare [14, Section 3 and Appendix A]. Hence, our hypothesis of non-degeneracy of g together with the implicit

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

6

function theorem can be used to obtain solutions of the Lorentzian equations, provided that suitable isomorphism theorems can be established for the matter equations. This will be done for the equations at hand in Sections 5 and 6 below. In cases involving Maxwell fields there will arise an issue related to gauge freedom for Maxwell fields which will be addressed in a somewhat similar manner to the addition of the Ω-terms to the Einstein tensor: To render (4.5) well posed we will add to it a “gauge-fixing” term σF , which will need to be shown to be zero. For definiteness we consider the equations resulting from (2.4): (4.9) ∇µ (W (φ)(∇ν Aµ − ∇µ Aν )) ≡ −∇µ (W ∇µ Aν ) − W Rα ν Aα + W ′ ∇µ φ∇ν Aµ + W ∇ν ∇µ Aµ | {z } = −λǫνµ1 ...µ2k F

µ1 µ2

...F

µ2k−1 µ2k

,

≡σF

and note that the divergence of the rightmost term above vanishes. We will show that we can solve the equation obtained by setting σF to zero in (4.9): (4.10) Equivalently, (4.11)

−∇µ (W ∇µ Aν ) − W Rα ν Aα + W ′ ∇µ φ∇ν Aµ = −λǫνµ1 ...µ2k F µ1 µ2 . . . F µ2k−1 µ2k .

∇µ (W (φ)(∇ν Aµ − ∇µ Aν )) = −W ∇ν ∇µ Aµ − λǫνµ1 ...µ2k F µ1 µ2 . . . F µ2k−1 µ2k .

Since the divergence of the left-hand side vanishes, we obtain (4.12)

∇ν (W ∇ν ∇µ Aµ ) = 0 . | {z } =:σF

It follows e.g. from [15, Theorem 3.3] that if ∇µ Aµ →ρ→0 0, then ∇µ Aµ ≡ 0, so that the solution of (4.11) solves (4.9). This implies that the energymomentum tensor is divergence-free, and we conclude as in the case without Maxwell fields. For some purposes it is convenient to replace (4.6) with its equivalent form 2Λ − gµν Tµν gαβ . (4.13) Rαβ = Tαβ − ∇α Ωβ − ∇β Ωα + n−1 We note that the linearisation at ˚ g of the Ω-contribution above is ˚α∇ ˚ µ hβν + ∇ ˚β ∇ ˚ µ hαν ) − ∇ ˚α∇ ˚ β (˚ ˚ (4.14) gµν (∇ gµν hµν ) , which cancels exactly the non-∆L terms in (3.1). 5. Static metrics In this section we present the construction of static solutions of the equations at hand. Strictly speaking, the results in this section are a special case of those in Section 6, but it appears instructive to present them separately, taking into account that the analysis here is computationally less demanding than the general case.

STATIONARY SPACETIMES WITH NEGATIVE Λ

7

Assuming staticity, in adapted coordinates the metric g becomes (5.1)

g = −V 2 dt2 + g ,

∂t V = 0 = ∂t g .

Equations (2.1) lead to the following set of equations, where we denote by D the covariant derivative of g (recall that ∇ is the covariant derivative of g), and where Rij is the Ricci tensor of g:   Trg T 2Λ −1 Rij = V Di Dj V + gij + 8πG Tij − gij , (5.2) n−1 n−1   2Λ Trg T (5.3) V D i Di V = 8πGV 2 Tαβ N α N β + , T0i = 0 , −V2 n−1 n−1 where N α ∂α is the g-unit timelike normal to the level sets of t. Choosing Λ as in (2.2), taking into account the Maxwell equations and the scalar field equations, together with (5.4)

∂t Fµν = 0 = ∂t φ ,

one is led to the system (5.5)  V2  V (−∆g V + nV ) = −2W (φ)F0i F0 i + n−1 (V (φ) − W (φ)|F |2 ) ,   gij  1 2 α −1   Rij + ng − V√ Di Dj V = 2 ∂i φ∂j φ + 2Fiα Fj W (φ) + n−1 (V (φ) − W |F | ) , 1 µν ν √ ∂ (V det gW (φ)F ) + BCS = 0 , V det g µ √   1 √  ∂i (V det gg ij ∂j φ) − W ′ (φ)|F |2 − V ′ (φ) = 0 ,    V det g W (φ)F0j Fi j = 0 ,

where W ′ and V ′ are understood as differentials of W and V when φ is ν is RN +1 valued with N ≥ 1, and where the Chern-Simons source-term BCS given by ( 0, n is odd; ν (5.6) BCS = λ να1 β1 ···αk βk − k+2 ǫ Fα1 β1 · · · Fαk βk , n = 2k. 2 5.1. Purely electric configurations. One way of satisfying the last equation in (5.5) is to assume a purely electric Maxwell field: (5.7)

F = d(U dt) ,

∂t U = 0 .

(Purely magnetic configurations will be considered in Section 5.2 below, while configurations with both electric and magnetic fields can be obtained by applying duality rotations to the Maxwell field at the end of the construction.) Equation (5.7) leads to the following form of (5.5):  1 2 2 V (−∆g V + nV ) = − 2(n−2)   n−1 W (φ)|dU |g + n−1 V V (φ) ,   Ric(g) + ng − V −1 Hessg V = 1 dφ ⊗ dφ + 1 V (φ)g   2 n−1   −2 −dU ⊗ dU + 1 |dU |2 g , (5.8) + 2W (φ)V g n−1     divg (V −1 W (φ)DU ) = 0 ,   −1 V divg (V Dφ) + 2V −2 |dU |2g W ′ (φ) − V ′ (φ) = 0

(the Chern-Simons term drops out because the purely spatial components of F vanish). When φ ≡ 0 the U -equation coincides with that in [15], thus Theorem 3.3 there with s = −1 applies. By continuity it still applies for all fields φ which

8

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

are sufficiently small in Cǫk+1,α, with any ǫ > 0. This motivates us to seek again solutions with U of the form b + O(ρ) , (5.9) U =U

b is smooth-up-to-boundary on M . (Here two comments are in order: where U b |∂M defined on the First, the key information is contained in the function U b defined on M , which boundary, but it is useful to invoke a function U avoids the issue of considering extensions to M of functions defined on ∂M . Next, the uniqueness part of our analysis below implies that solutions with b |∂M = c for some constant c ∈ R lead to configurations with U ≡ c, hence U b modulo constants trivial Maxwell fields. In other words, in (5.9) only U matters as far as physically relevant fields are concerned. Nevertheless, different c’s lead indeed to different fields U .) We will seek solution such that (5.10)

φ → 0 as ρ → 0.

Note that with the choice (5.9) the coefficient 2V −2 |dU |2g appearing in the φ-equation will be O(ρ4 ). Assuming (5.10) we must have V ′ (0) = 0, so that in the scalar case the indicial exponents for the φ-equation (cf., e.g., [23]) will be solutions of the equation r n2 n + V ′′ (0) . (5.11) σ(σ − n) − V ′′ (0) = 0 ⇐⇒ σ± = ± 2 4 When φ is RN +1 valued, with N ≥ 1, we obtain a collection of indicial exponents, with V ′′ (0) in (5.11) replaced by the eigenvalues of V ′′ (0). Now, to ensure useful properties of the operator associated with the equation for φ we need all σ± to be real with σ+ 6= σ− , which leads to the condition n2 < V ′′ (0) , 4 understood as a matrix inequality for the Hessian of V at φ = 0 when N ≥ 1. After diagonalising V ′′ (0), each component in the diagonalising basis of the solutions we will construct will have the asymptotic behaviour b σ− + o(ρσ− ) . (5.13) φ = φρ (5.12)



b ∂M ≡ 0, and note that we have possibly with φ|  b ∂M 6≡ 0 =⇒ (5.14) φ →ρ→0 0 and φ| σ− > 0 ⇔ V ′′ (0) < 0 .

The properties of solutions of the φ-equation depend now upon whether or not W ′ (0) = 0. Let us first assume that W ′ (0) = 0. We will then have a non-trivial b ∂M 6≡ 0. Since solution φ 6≡ 0 tending to zero when ρ → 0 if and only if φ| ′′ the case V (0) ≥ 0 leads then to solutions which do not tend to zero at the conformal boundary, the hypothesis that V ′′ ≥ 0 and W ′ (0) = 0 leaves us with Maxwell matter fields only. Assuming that V ′′ (0) < 0 and W ′ (0) = 0, it remains to check that the source terms in the remaining equations are compatible with the isomorphism ranges of the relevant operators. For this it is convenient to rewrite

STATIONARY SPACETIMES WITH NEGATIVE Λ

9

the V -equation as 1 2(n − 2) −1 V (φ))V = − V W (φ)|dU |2g , n−1 n−1 1 V goes to zero at the boundary when φ does, we Since the coefficient n−1 obtain the same indicial exponents as when φ ≡ 0, and thus again an isomorphism for φ small enough. No new conditions arise from the remaining equations either. b ∂M 6= 0 in view of (5.12) we must have Summarising, for φ|

(5.15)

−∆g V + (n +

− n2 < 4V ′′ (0) < 0 ,

(5.16)

If φ has more than one component then the above inequalities apply with V ′′ (0) replaced by the relevant eigenvalue of the Hessian of V (0) (in this case it is convenient to work in a diagonalising basis). If W ′ (0) 6= 0 the situation is different, as then the scalar fields φ are b ∂M . First, if U b |∂M ≡ 0, driven both by the term 2V −2 |dU |2g W ′ (0) and by φ| b ∂M vanishes as well, then φ ≡ 0, and we are in vacuum. then U ≡ 0, and if φ| b ∂M ≡ 0 and U b |∂M 6≡ 0 then we have non-trivial On the other hand, if φ| solutions with the following asymptotic behaviour:  if σ+ < 4;  O(ρσ+ ) 4 O(ρ ln ρ) if σ+ = 4; (5.17) φ=  O(ρ4 ) if σ+ > 4 .

An analysis similar to that of the case W ′ (0) = 0, taking into account that σ+ > n/2 under (5.12), shows that non-trivial φ’s tending to zero at the b |∂M 6≡ 0 if, in the one-component scalar boundary will be obtained when U field case,  −n2 < 4V ′′ (0) < 0, or (5.18) b ∂M ≡ 0. V ′′ (0) ≥ 0 and φ|

It is clear that the solutions will be unstable if V ′′ (0) < 0 (compare [25]), but this is irrelevant from the point of view of the question existence of static solutions, which is our only interest in this work. One can now proceed exactly as in [15] to obtain the following: Consider the field equations associated with the action (2.4) for time-independent fields, with (5.19) Let

W (0) = 1 ,

V (0) = 0 = V ′ (0) ,

V ′′ (0) > −n2 /4 .

˚2 dt2 + ˚ ˚ g = −V g be smoothly conformally compactifiable at ∂M and satisfy the vacuum Einstein equations with a negative cosmological constant. We then have the following: ˚2 dt2 +˚ Proposition 5.1. Suppose that (S 1 ×M, V g ) is nondegenerate. Under (5.19), assume that b and φb which are smooth functions on M suffi(1) V ′′ (0) < 0 with U ciently close to zero, or b is a smooth function on M sufficiently close to zero. (2) φb ≡ 0 and U

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

10

Then there exists a static solution of the equations with U as in (5.9) and φ as in (5.13). The solutions are uniquely determined, within the class of b ∂M , b |∂M and φ| static solutions belonging to some neighbourhood of ˚ g, by U with all fields having a polyhomogeneous expansion at ∂M . 

b ∂M ≡ 0 and W ′ (0) = 0 then φ ≡ 0, so that we obtain the solutions of If φ| Einstein-Maxwell equations already constructed in [15]. We emphasise that uniqueness is in the gauges implicitly defined above; for instance, two solutions with U -fields differing by a constant are considered distinct, even though they define of course the same Maxwell field Fµν . Furthermore, uniqueness is up-to diffeomorphisms which are the identity at the boundary in any case. It is conceivable that uniqueness holds for arbitrary diffeomorphisms, but this does not follow from our arguments. 5.2. Purely magnetic configurations. Another way to satisfy the last equation in (5.5) is to consider purely magnetic fields, i.e. F = d(Ai dxi ) ,

(5.20)

∂t Ai = 0 .

i BCS

This implies = 0 and leads to the following matter equations,   Dj (V W gjk gil (Al,k − Ak,l )) = 0 , (5.21) V −1 D (V gij ∂j φ) − 2W ′ (φ)(Aj,i − Ai,j )gik gjl Al,k − V ′ (φ) = 0 ,  0 i BCS = 0 ,

with the Einstein equations taking now the form (5.22)  V2 V (−∆g V + nV ) = n−1 (V − 2W W ′ (φ)(Aj,i − Ai,j )gik g jl Al,k ) ,     1 Rij + ngij − V −1 Di Dj V = (∂i φ)(∂j φ) + 2W (Ak,i − Ai,k )(Al,j − Aj, l)gkl 2   gij   (V − 2W W ′ (φ)(Aj,i − Ai,j )gik gjl Al,k ) . + n−1 To satisfy the last equation of (5.21) one might as well assume that the Chern-Simons coupling constant λ vanishes. The first line of (5.21) can be rewritten, after introducing σF as in (4.12), in the form (5.23) 0 = Dj (V W gjk gil (Dk Al ))

  − D k (V W )D i Ak − V W Ri ℓ Aℓ − V W D i σF − V −1 Ak Dk V .

If we develop (5.23) and drop σF , the operator acting on A becomes

V W B(A)i + V D k W (Dk Ai − Di Ak ) + W Ak Di Dk V − V W Rik Ak ,

where B is the operator of Lemma A.3 in [15]. The operator P := B + (V −1 DDV − Ric(g))

appears as part of the (n + 1)-dimensional Riemannian Hodge Laplacian Dg∗ Dg + Ric(g) [14]. Recall that, in coordinates, the characteristic indices for Dg∗ Dg + Ric(g) ∼ Dg∗ Dg − n belong to {−1, n − 1} in the normal direction and to {0, n − 2} in the tangential one (cf., e.g. [6, Section 2.3]). We deduce from Theorem C(c) and Corollary 7.4 of [23] that if there are no harmonic

STATIONARY SPACETIMES WITH NEGATIVE Λ

11

forms in L2 for (S 1 × M,˚ g), then P will be an isomorphism from Cδk+2,α to Cδk,α for n n (5.24) δ − < − 1 . 2 2 Let us denote by (ρ, xa ) local coordinates near ∂M . From what has been said one expects solutions to take the form   bρ ρ−1 + O(1) dρ + A ba + O(ρ) dxa , (5.25) A= A

bi . As discussed in detail at the end with smooth-up-to-boundary functions A of Section 4, we will have σF ≡ 0 if and only if ∇µ Aµ tends to zero as ρ tends to zero. Now √ ∂i (gij det gV Aj ) µ √ = ∂ρ (ρ1−n Aρ )ρ1+n + O(ρ) . (5.26) ∇ Aµ = det gV bρ |∂M ≡ 0, without any restrictions on This will be satisfied if and only if A ba |∂M dxa . We conclude that: A

Proposition 5.2. Under the hypotheses of Proposition 5.1, suppose more˚ 2 dt2 +˚ over that the Einstein metric (S 1 ×M, V g ) has no harmonic one-forms 2 which are in L . Then the conclusions of Proposition 5.1 hold with U rebρ |∂M ≡ 0, so that U b |∂M is replaced placed by Ai dxi of the form (5.25) with A a ba |∂M dx . by A 

As discussed in Appendix C, it follows from [12] that there are no L2 harmonic one-forms on S 1 × M equipped with the Riemannian counterpart of the AdS metric ˚ g, so the same is true for nearby metrics. Further remarks concerning asymptotics and total energy are to be found in Section 7 below.

5.3. Yang-Mills-Higgs fields. The analysis so far readily generalises to Yang-Mills-Higgs-Chern-Simons fields. Here one often assumes that the Lie algebra G of the structure group G admits a positive-definite scalar product, but this is not needed in our considerations. We denote by A = Aµ dxµ the Yang-Mills connection, with Aµ taking values in G, and by F = 12 Fµν dxµ ∧ dxν its curvature: Fµν = ∂µ Aν − ∂ν Aµ + [Aµ , Aν ] .

The scalar fields φ are allowed to be coupled to A in the usual way, with derivatives involving φ replaced by gauge-covariant derivatives ∂µ φ 7→ Dµ φ := ∂µ φ + T (Aµ )φ ,

where T is the linear map determined by the relevant representation; e.g., if φ are sections of a bundle associated to the adjoint representation, then T (Aµ )φ = [Aµ , φ].

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

12

We suppose that the G-valued current vector j ν appearing in the YangMills equations, Dµ F µν := ∇µ F µν + [Aµ , F µν ] = j ν , depends only upon A, ∂A, φ, ∂φ, g, ∂g, and is at least quadratic in all the fields,

(5.27) (5.28) (5.29)

and satisfies the obvious compatibility conditions arising from (5.30)

Dν Dµ F µν = D[ν Dµ] F µν = [Fµν , F µν ] = 0

=⇒

Dν j ν = 0 .

Equivalently (5.31)

∇ν (j ν − [Aµ , F µν ]) = 0 .

More precisely, we will need that (5.32)

Dν j ν = 0 whenever the field equations for φ are satisfied.

This will be automatically satisfied from currents arising from gaugeinvariant Lagrange functions, and from the current arising from ChernSimons terms [13], whether Abelian or non-Abelian, in the Lagrangean. Let us, first, assume that the Yang-Mills principal bundle, say P (G), is trivial, so that F is globally defined as a two-form. The case where A = U dt, ∂t U = 0, where U is G-valued, works exactly as in the Maxwell case, leading immediately to an obvious Yang-Mills equivalent of Proposition 5.1 whenever P (G) = G × M . A purely-magnetic Yang-Mills potential, A0 ≡ 0, ∂t Ai ≡ 0, does not require much more work. Under our conditions, (5.21)-(5.22) are only modified by terms which are at least quadratic in the fields and which are lower order in terms of derivatives. Such terms, when small enough in relevant norms, do not affect the argument: One can view F as a collection of several electric fields, introduce a vector-valued gauge-source function σF (one such function for each component of F ), and conclude as before. In other words, assuming (5.27)-(5.29) and (5.32), we have established the Yang-Mills equivalent of Proposition 5.2 for trivial G-bundles. This establishes, for small A, existence of the solutions constructed numerically in [9], and in fact provides a much larger family of such solutions. (An even larger family of such solutions results from Theorem 6.1 below with ϕ| b ∂M = 0 = W ′ (0).) Summarising, we have proved: Proposition 5.3. The conclusions of Propositions 5.1 and 5.2 hold when Maxwell and scalar fields are replaced by Yang-Mills and Higgs fields on a trivial gauge bundle. 

Non-trivial bundles can be handled by introducing a suitably regular back˚ This leads to a globally defined G-valued ground Yang-Mills connection A. ˚ and a corresponding globally defined G-valued σF function one-form A − A, ˚ µ (Aµ − A ˚µ ) := ∇µ (Aµ − A ˚µ ) + [A ˚µ , Aµ − A ˚µ ] . σF := D

The existence argument goes through if one moreover assumes that (1) there are no covariantly-constant Higgs fields φ which are in L2 (S 1 × M ), and that (2) there are no G-valued harmonic forms which are in L2 (S 1 × M ).

STATIONARY SPACETIMES WITH NEGATIVE Λ

13

5.4. Time-periodic scalar fields. Let us allow complex-valued φ’s, and assume that (5.33)

V (φ) = GV (|φ|2 ) and W (φ) = GW (|φ|2 )

for some differentiable functions GV and GW , with the term (∇φ)2 in the action replaced by ∇α φ∇α φ, where φ is the complex conjugate of φ. Considering, as in [21], a time dependent field of the form (5.34)

φ(t, x) = eiωt ψ(x) ,

with ω , ψ(x) ∈ R ,

(compare [18]) leads to (5.35)  V2  V (−∆g V + nV ) = − 12 ω 2 ψ 2 − 2GW F0i F0 i + n−1 (GV − GW Fαβ F αβ ) ,      Rij + ngij − V −1 Di Dj V = 1 ∂i ψ∂j ψ + 2Fiα Fj α GW    2   gij + (GV − GW Fαβ F αβ ) , n − 1 √   ν = 0,  √1 ∂j (V det gGW F jν ) + BCS  V det g   √  1  ij ∂ ψ) + V −2 ω 2 ψ − G ′ |F |2 + G ′ ψ = 0 , √  ∂ (V det gg j W i V    V det g j GW F0j Fi = 0 .

We see that the indicial exponents for the system remain unchanged, so that the existence and uniqueness theory with ω = 0, presented above, applies without changes for all sufficiently small ω ∈ R:

Proposition 5.4. The conclusions of Propositions 5.1 and 5.2 hold for all b and sufficiently small ω ∈ R where φ takes the form (5.34) with φb = eiωt ψ, b where ψ is smooth up-to-boundary. 

Recall that in [5] similar Einstein-scalar field solutions have been constructed with, however, V ≡ 0. Our condition V ′′ < 0 is evaded there by letting −ω 2 be an eigenvalue of the operator ψ 7→ V Di (V D i ψ). It would be of interest to provide a proof of existence of such solutions using our methods; compare [8] for the spherically symmetric asymptotically flat case. The discussion of the finiteness of energy of the resulting field configuration, to be found in Section 7, is identical to the ω = 0 case. 6. Stationary metrics b

We return to the metric (4.1)-(4.2). Let eµb be the coframe e0 := dt + θ, bi e := dxi . The corresponding components Rµbνb of the Ricci tensor of g read (see, e.g., [17])  1 2  Rb0b0 = V ∆g V + 4 |λ|g , Rbb = Ric(g)ij − V −1 Hessg Vij + 2V1 2 λik λj k , (6.1)  ij Rb0bj = − 21 V −1 Di (V λi j ) , where

λij = −V 2 (∂i θj − ∂j θi ) .

14

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

For a general energy-momentum tensor T the equations are (6.2)    1 2 − 8πG T + V 2 Tr T  V (−∆ V + nV ) = |λ| ,  g 00 g g  4 n−1    1  R + ng − V −1 D D V = λik λk j + 8πG θi θj T00 − θi T0j − θj Ti0 ij ij i j 2V 2   gij   + T − Tr T , ij g   n−1   j D (V λij ) = 16πGV (T0i − θi T00 ) . When T is given by (2.3)) we have   1 n−3 n−1 n+1 2 2 Trg T = − W (φ)|F | + |Dφ| + V (φ) . 8πG 2 4 2 Assuming moreover ∂t φ = 0 we obtain (6.3)  V2  (V − W |F |2 ) , V (−∆g V + nV ) = 41 |λ|2g − 2W F0i F0 i + n−1    1 1    Rij + ngij − V −1 Di Dj V = 2 λik λk j + (∂i φ)(∂j φ) + 2W Fiα Fj α   2V 2 gij (V − W |F |2 ) +   n − 1     − 2W F0k (Fj k θi + Fi k θj − F0 k θi θj ) ,    j D (V λij ) = 4V W F0j (Fi j − F0 j θi ) .

The matter equations remain formally unchanged, as compared to (5.5), when written in the form ( √ ν = 0, √1 ∂ (V det gW F µν ) + BCS V det g µ √ (6.4) √1 ∂ (V det gg ij ∂j φ) − W ′ (φ)|F |2 − V ′ (φ) = 0 . V det g i ν as in (5.6); indeed, the theta-dependent terms are hidden in |F |2 with BCS and F µν . Letting F = d(U dt + Ai dxi ) and ∂t U = ∂t A = 0 we have (6.5)  i = 0,  Dj (V W gjk gil (Al,k − Ak,l + θk U,l − θl U,k )) + V BCS 0 = 0, Dj (V W gjk (−V −2 U,k + g lm θl (θm U,k − θk U,m + Ak,m − Am,k ))) + V BCS  −1 ij ′ 2 ′ V Di (V g ∂j φ) − W (φ)|F | − V (φ) = 0 ,

where

i h |F |2 = 2 (Aj,i − Ai,j )gik gjl (Al,k − 2U,k θl ) + |∇U |2g (|θ|2g − V −2 ) − (U,i θ i )2 .

Let ǫ > 0. For θ small in Cǫk,α-norm the last two operators in (6.5) acting on φ and U are close in norm to the operators considered in Section 5, and therefore isomorphisms as discussed there. Hence, in the implicit function b and φb as in Section 5. It remains to consider the argument we can choose U A-equation. p 1 σF ≡ ∇µ Aµ ≡ √ ∂µ ( det ggµν Aν ) (6.6) det g  −1 = V Di V (Ai − U θ i ) .

STATIONARY SPACETIMES WITH NEGATIVE Λ

15

Then the first line of (6.5) can be rewritten as (6.7)

i Dj (V W gjk gil (Dk Al − Dl Ak + θk U,l − θl U,k )) + V BCS

= Dj (V W gjk g il (Dk Al + θk U,l − θl U,k ))

i −D k (V W )D i Ak − V W Dj D i Aj + V BCS

= Dj (V W gjk g il (Dk Al + θk U,l − θl U,k )) − D k (V W )D i Ak   −V W Ri ℓ Aℓ − V W D i σF − V −1 Ak Dk V + V −1 Dk (U V θ k ) i +V BCS .

The resulting linear operator acting on A coincides with the one in (5.23), so that the discussion there applies. Inserting the asymptotic expansions for V and g into (6.6) gives (6.8)

σF = ∇µ Aµ = ρ1+n ∂ρ (ρ1−n (Aρ − θρ U )) + O(ρ)

bρ |∂M ≡ 0 guarantees σF ≡ 0, as discussed at and, as U = O(1), setting A the end of Section 4. The rest of the proof is an application of the implicit function theorem, we sketch the details. We work with ˚ + v,˚ (V, g) = (V g + h) ˚,˚ close to (V g ). Keeping in mind that λ = −V 2 dθ, W (φ) = 1 + O(φ) and V ′ (φ) = V ′′ (0)φ + O(φ2 ),

the system obtained after the addition of the Ω-terms as in (4.6) is of the form:  Q(v, h, θ) − q1 [v, h, θ, A, U, φ] = 0 ,    P(A) − dσF − q2 [v, h, θ, A, U, φ] = 0 , (6.9)  V Dj (V −1 D j U ) − q3 [v, h, θ, A, U, φ] = 0 ,   −1 V Di (V D i φ) − V ′′ (0)φ − q4 [v, h, θ, A, U, φ] = 0 ,

where the qi ’s are at least quadratic in their arguments and their first derivatives, and where Q(v, h, θ) corresponds to the operators (l, L, L) of [14, Corollaries 3.2 and 3.3], which are the three components of the operator 1 2 ∆L + n, with (l, L) and L being isomorphisms. For s ∈ R we define, as in [15], the operators Ts = V −s Di (V s D i ·). We consider the modified system (6.9) with the Maxwell gauge-term dσF added:  Q(v, h, θ) − q1 [v, h, θ, A, U, φ] = 0 ,    P(A) − q2 [v, h, θ, A, U, φ] = 0 , (6.10) T−1 (U ) − q3 [v, h, θ, A, U, φ] = 0 ,    (T1 − V ′′ (0))(φ) − q4 [v, h, θ, A, U, φ] = 0 .

A solution, close to zero, of the elliptic system (6.10), with prescribed behavior at infinity, can be be constructed in the following way: Let us define X = (v, h, θ, U, A, φ) ,

we want to solve an equation of the form

F(X ) = 0 ,

16

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

b at large distance, for a prescribed small with X ∼ X bU b b = (b b , A, b φ) X v, b h, θ,

b + X and define (some of the components vanishing if desired).We let X = X b X) := F(X b + X). F(X,

We have F(0, 0) = 0, with the linearisation

DX F(0, 0) = diag(Q, P, T−1 , (T1 − V ′′ (0)))

b small, being an isomorphism. By the implicit function theorem, for all X b such that X is a solution there exist a small X, depending smoothly on X, of (6.10). We have already explained why this solution solves the desired original equations. For completeness we provide examples of functional spaces where the preceding procedure involving F applies. Taking into account the weights needed so that each of the operators involved is an isomorphism, a natural space for X is, without indicating the tensor character of the relevant bundles, (6.11)

k+2,α k+2,α k+2,α E k+2 := C−1+s × Csk+2,α × C1+s × C1+s × Csk+2,α × Cσk+2,α , − +s

b ∂M is where s is greater than and close to zero. The space for X|

ρ−1 C k+2,α × ρ0 C k+2,α × ρ0 C k+2,α × ρ0 C k+2,α × ρ0 C k+2,α × ρσ− C k+2,α .

The tensor fields in C k+2,α(∂M ), can then be extended away from ∂M ¯ in any convenient way, keeping in mind the to smooth tensor fields on M conditions bρ |∂M = θbρ |∂M = b (6.12) A hρi |∂M = 0 .

b .) and The reader can check that with the spaces chosen above, both F(X, k+2 k DX F(0, 0) map E to E . We have thus proved: ˚2 dt2 + ˚ Theorem 6.1. Suppose that the Einstein metric (S 1 × M, V g ) is non-degenerate and has no harmonic one-forms which are in L2 . Assume that V (0) = 0 = V ′ (0) , V ′′ (0) > −n2 /4 . b ∂M which b |∂M , A ba |∂M dxa , and φ| (1) and V ′′ (0) < 0 with θba |∂M dxa , U are sufficiently small smooth fields on ∂M , or b ∂M , and A ba |∂M dxa are sufficiently small (2) φb ≡ 0 and θba |∂M dxa , U| smooth fields on ∂M . Then there exist a solution of the Einstein-Maxwell-dilaton-scalar fieldsChern-Simons equations, or of the Yang-Mills-Higgs-Chern-Simons-dilaton equations with a trivial principal bundle, so that near ∂M we have ˚ , U →ρ→0 U b , A →ρ→0 A ba dxa , θ →ρ→0 θb , (6.14) g →ρ→0 ˚ g , V →ρ→0 V (6.13)

W (0) = 1 ,

with all convergences in ˚ g -norm. The hypothesis of non-existence of harba |∂M dxa ≡ 0 ≡ U b |∂M , in which case monic L2 -one-forms is not needed if A the Maxwell field or the Yang-Mills field are identically zero. 

STATIONARY SPACETIMES WITH NEGATIVE Λ

17

Remark 6.2. Some comments on properties of the solutions are in order. The case (b v |∂M , b hab |∂M dxa dxb , θba |∂M dxa ) = 0 leads to a solution with the usual AdS conformal boundary when ˚ g is taken to be the AdS metric. b ∂M = 0 with |b b |∂M = A ba |∂M dxa = φ| The case θba |∂M dxa = U v |∂M | + a b b |hab |∂M dx dx | = 6 0 leads to the non-trivial vacuum configurations constructed in [14]. b 6≡ 0 and A b 6≡ 0 might lead to a solution Note that θa |∂M dxa ≡ 0 but U with θ 6≡ 0 because the off-diagonal terms of the Maxwell energy-momentum tensor will drive a non-zero θ. ba |∂M dxa ≡ 0, U b |∂M ≡ 0 and θba |∂M dxa ≡ 0 we obtain, from If we choose A uniqueness of solutions, static solutions with scalar fields. The vanishing of φ|∂M will lead to φ = 0 everywhere only if W ′ (0) = 0, since otherwise the equation for φ is non-homogeneous.  6.1. Time-periodic scalar fields. Consider time-periodic scalar fields of the form (6.15)

φ(t, x) = eiωt ψ(x) , ω ∈ R ,

as done in the static case in section 5.4, but where now ψ(x) is allowed to be complex. Assume for simplicity that all Maxwell fields are Abelian. Using the notation (5.33), and adapting the action as before gives for the Einstein equations (6.16)   V2   (GV − GW |F |2 ) , V (−∆g V + nV ) = 14 |λ|2g − 12 ω 2 |ψ|2 − 2GW F0i F0 i + n−1     1 1   ¯ Rij + ngij −V −1 Di Dj V = λik λk j + ℜ(∂i ψ∂j ψ)   2  2V 2    1  ¯ j ψ) − ωθj ℑ(ψ∂ ¯ i ψ)) + (θi θj ω 2 |ψ|2 − ωθi ℑ(ψ∂ 2  gij   + 2GW Fiα Fj α + (GV − GW |F |2 )   n − 1      − 2GW F0k (Fj k θi + Fi k θj − F0 k θi θj ) ,       V −1 D j (V λij ) = 4GW F0j (Fi j − F0 j θi ) + ωℑ(ψ∂ ¯ i ψ) − θi ω 2 |ψ|2 , where ℑ denotes taking the imaginary part. The matter equations are then (6.17)   i = 0,  Dj (V GW g jk gil (Al,k − Ak,l + θk U,l − θl U,k )) + V BCS       Dj (V GW g jk (−V −2 U,k + g lm θl (θm U,k − θk U,m + Ak,m − Am,k ))) 0 = 0, +V BCS     −1 ij ′ 2 ′  V D (V g ∂ ψ) − G |F | + G ψ  i j W V    −2 k 2 j + (V − θk θ )ω ψ + iω(θ ∂j ψ + V −1 Dj (V θ j ψ)) = 0 .

It should be clear that all our previous arguments apply to this system of equations, leading to Proposition 6.3. Assuming (5.33), the conclusions of Theorem 6.1 concerning the Einstein-Maxwell-Chern-Simons-dilaton-scalar fields hold for all

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

18

b ∂M when φ takes the form (6.15) with sufficiently small ω ∈ R and ψ| iωt b b b φ = e ψ, where ψ is smooth up-to-boundary.  7. Asymptotics, energy

Whatever follows applies to the static, or stationary, or time-periodic solutions constructed above. All our solutions have a polyhomogeneous expansion at ∂M , that is, expansions in terms of integer powers of ln ρ and of suitable powers of ρ, as determined by all indicial exponents. It is straightforward but tedious to obtain a detailed description of the asymptotic behavior by inserting polyhomogeneous expansions in the equations and comparing coefficients. We emphasise that non-integer indicial exponents σ± for the scalar fields (cf. (5.11)) will introduce non-integer powers of ρ in asymptotic expansions for small ρ of all fields involved unless some miraculous cancellations occur. Logarithms of ρ are expected in the expansion for generic solutions regardless of whether or not σ± are in Z. On constant-t slices the matter energy-density Tb0b0 ≡ V −2 T00 reads  1  1 −2 (V ∂0 φ∂0 φ + |dφ|2g ) + V + W (4V −2 F0µ F0 µ + |F |2 ) . Tb0b0 = 16πG 2 As in [15], the total energy content of non-trivial Maxwell or Yang-Mills b ∂M 6≡ 0 fields will be finite only in space-dimensions n = 3 and n = 4. If φ| 2σ the φ-contribution to ρ behaves as ρ − , which will lead to a finite total energy of the scalar field if and only if (7.1) − n2 < 4V ′′ (0) < −n2 + 1 . b ∂M ≡ 0 then either φ ≡ 0 when there is no Maxwell field and so If φ| the space-time is vacuum, or there is a Maxwell field in which case the φ-contribution to the energy is  O(ρ2σ+ ) if σ+ < 4; (7.2) 2 8 O(ρ ln ρ) if σ+ = 4, which gives a finite integral in either case.

Acknowledgements The research of PTC was supported in part by the Austrian Research Fund (FWF), Project P 24170-N16. PK is supported by a uni:docs grant of the University of Vienna. Appendix A. The Ω equation Consider (4.7): (A.1)

(g P Ω)ν := ∇µ (∇µ Ων + ∇ν Ωµ − ∇α Ωα gµν ) = 0 .

STATIONARY SPACETIMES WITH NEGATIVE Λ

19

We have (A.2)

−( g P Ω)ν = ∇µ (∇µ Ων + ∇ν Ωµ )−∇ν ∇α Ωα = ∇µ (∇µ Ων − ∇ν Ωµ + 2∇ν Ωµ )−∇ν ∇α Ωα = ∇µ (∇µ Ων − ∇ν Ωµ ) + ∇ν ∇µ Ωµ + 2Rν µ Ωµ p  1 = p ∂µ | det g|gµα gνβ (∂α Ωβ − ∂β Ωα ) | det g| h p i 1 ∂µ | det g|gµβ Ωβ + 2Rν µ Ωµ . +gνα ∂α p | det g|

In the static case, with a time-independent Ων and an Einstein (n + 1)dimensional metric g normalised so that Rµν = −ngµν , this reads

(A.3)

(A.4)

∇µ (∇µ Ω0 + ∇0 Ωµ ) = g00 ∇µ (∇µ Ω0 + ∇0 Ωµ ) p  1 ∂i | det g|gij g00 ∂j Ω0 − 2nΩ0 = g00 p | det g| p  V ∂i V −1 det gg ij ∂j Ω0 − 2nΩ0 = √ det g = V Di (V −1 D i Ω0 ) − 2nΩ0 , ∇µ (∇µ Ωk + ∇k Ωµ ) = p  1 gkℓ ∂i | det g|gij gℓm (∂j Ωm − ∂m Ωj ) = p | det g| h p i 1 ∂i | det g|gij Ωj − 2nΩk . +∂k p | det g|

One notices that g P coincides with its Riemannian analogue g P when mapping covectors to covectors: g P =gP . In particular if ∇µ (∇µ Ων + ∇ν Ωµ −gµν ∇α Ωα ) = 0

(A.5)

in the Lorentzian metric, then the same remains true in the Riemannian one. Consider a one-form Ω which is in L2 and satisfies g

P Ω ν ≡ ∇µ ∇µ Ω ν + R µ ν Ω µ = 0 .

Multiplying (A.5) by −gνµ Ωµ , and integrating by parts over S 1 × M using g everywhere one finds Z Z (A.6) 0 = |∇Ω|2g + n|Ω|2g , ∇µ Ων ∇µ Ων + nΩαΩα ≡ S 1 ×M

S 1 ×M

where the vanishing of the boundary term follows from elliptic regularity and density arguments. Thus Ωµ ≡ 0; equivalently, the L2 kernel of g P is trivial. As g P is an elliptic, formally self-adjoint operator, geometric in the sense of [23], it is an isomorphism from Cδk+2,α to Cδk,α by the results in this last reference when n n δ − < + 1 . 2 2

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

20

Equivalently, in local coordinates, the indicial exponents belong to {−2, n}. 0,α In particular any C−1+ǫ one-form, with ǫ > 0, in the kernel of g P is trivial. In our applications, we will have Ω in the kernel with Ω ∈ Csk+1,α where s is positive close to zero when (b v, b h) = 0 and s = 0 if (b v, b h) 6= 0, which guarantees the vanishing of Ω. The usual perturbation arguments show that the above conclusions will remain true for all sufficiently small θ. The alert reader will have noted that the above discussion goes through for all negative-definite Ricci tensors such that R0i = 0. Appendix B. The Christoffels For further reference, and to get insight into (4.7), we compute the Christoffels symbols of a stationary metric. The computations will be made for a Lorentzian metric but the change V → iV , where i2 = −1 gives the results for a Riemannian metric. We recall the form of the metric considered here: g = −V 2 (dt + θi dxi )2 + gij dxi dxj , | {z } | {z }

(B.1)



(B.2)

=g

∂t V = ∂t θ = ∂t g = 0 .

Its inverse is (B.3) g# ≡ gµν ∂µ ∂ν = −(V −2 − g ij θi θj )∂02 − 2g ij θi ∂0 ∂j + gij ∂i ∂j = −V −2 ∂02 + g ij (∂i − θi ∂0 )(∂j − θj ∂0 ) .

The Christoffel symbols are then g 0 Γ00

= −θ i V Di V,

g k Γ00

= V D k V,

1 = −(|θ|2 − V −2 )V Dj V + θ k [∂j (V 2 θk ) − ∂k (V 2 θj )], 2 1 g 0 2 −2 2 Γij = − (|θ| − V )[∂j (V θi ) + ∂i (V 2 θj )] − θ l Γlij (g − V 2 θ ⊗ θ) , 2 1 g k Γi0 = θ k V Di V − gkj [∂i (V 2 θj ) − ∂j (V 2 θi )] , 2 1 g k Γij = θ k [∂i (V 2 θj ) + ∂j (V 2 θi )] + g kl Γlij (g − V 2 θ ⊗ θ) , 2 where, as usual, gij is inverse to gij and where 1 Γijk = (gik,j + gji,k − gjk,i ) . 2 g 0 Γ0j

STATIONARY SPACETIMES WITH NEGATIVE Λ

21

Appendix C. AdS, Hn and a quotient Let us consider the Poincar´e ball model of the hyperbolic space Hn . The hyperbolic space is then the unit ball of Rn endowed with the hyperbolic metric ˚ g = ρ−2 δ , where δ is the Euclidean metric and 1 ρ(x) = (1 − |x|2 ) . 2 We define ˚ = ρ−1 − 1 . V A model for (n + 1)-dimensional hyperbolic space is (see below for details of this model) R × Hn , equipped with the warped product metric ˚ 2 dt2 + ˚ ˚ g=V g, we thus denote as usual Hn+1 = R

˚2 × V

Hn .

Anti-de Sitter space is the same manifold with the Lorentzian metric ˚2 dt2 + ˚ ˚ g = −V g.

Let Γ = Z ⊂ R be a discrete subgroup of isometries of the R factor of Hn+1 . Then we can write Γ\Hn+1 = S1 V˚2 × Hn . Recall that the limit set Λ(Γ) (see eg. [24], §12.1 page 573) is a subset of the sphere at infinity of Hn+1 consisting of the union of the limits of all the orbits. We will show that Λ(Γ) consists of two points at infinity. It then follows from [12, Theorem C] that our quotient has no L2 harmonic one-forms. To justify our claim about Λ(Γ), consider the half space model of Hn+1 with (~x, y) ∈ Rn × (0, ∞) endowed with the metric ds2 = y −2 (|d~x| 2 + dy 2 ). Let H be the half sphere |~x|2 + y 2 = 1 with y > 0. It is well known that the totally geodesic hypersurface H with the induced metricph is a model of the hyperbolic space Hn . We reparametrize Hn+1 with (~u, 1 − |~u|2 ) ∈ H and t ∈ R by p (~x, y) = et (~u, 1 − |~u|2 ). In the coordinate system (t, ~u) ∈ R × B n (0, 1), the metric becomes ds2 = (1 − |~u|2 )−1 dt2 + h.

Comparing the Ricci curvature of the Einstein metrics ds2 and h (see appendix of [15]), we see that the radial function V = (1 − |~u|2 )−1/2 on H satisfies Hessh V = V h. As we also have min V = 1, the space Hn+1 = (R × H, ds2 ) is then (R × Hn ,˚ g). Now for ~u fixed with t → ±∞ the point (~x, y) tends to 0 or infinity. These correspond to two antipodal points of the sphere at infinity in the Poincar´e ball model.

22

´ P.T. CHRUSCIEL, E. DELAY, AND P. KLINGER

References 1. M.T. Anderson, Einstein metrics with prescribed conformal infinity on 4-manifolds, Geom. Funct. Anal. 18 (2001), 305–366, arXiv:math.DG/0105243. MR 2421542 2. , Boundary regularity, uniqueness and non-uniqueness for AH Einstein metrics on 4-manifolds, Adv. in Math. 179 (2003), 205–249, arXiv:math.DG/0104171. MR 2010802 3. M.T. Anderson, P.T. Chru´sciel, and E. Delay, Non-trivial, static, geodesically complete vacuum space-times with a negative cosmological constant, Jour. High Energy Phys. 10 (2002), 063, 22 pp., arXiv:gr-qc/0211006. MR 1951922 , Non-trivial, static, geodesically complete space-times with a negative cosmo4. logical constant. II. n ≥ 5, AdS/CFT correspondence: Einstein metrics and their conformal boundaries, IRMA Lect. Math. Theor. Phys., vol. 8, Eur. Math. Soc., Z¨ urich, 2005, arXiv:gr-qc/0401081, pp. 165–204. MR MR2160871 5. D. Astefanesei and E. Radu, Boson stars with negative cosmological constant, Nucl. Phys. B 665 (2003), 594–622, arXiv:gr-qc/0309131. MR 2000918 6. E. Aubry and C. Guillarmou, Conformal harmonic forms, Branson-Gover operators and Dirichlet problem at infinity, Jour. Eur. Math. Soc. (JEMS) 13 (2011), 911–957. MR 2800480 7. A.L. Besse, Einstein manifolds, Ergebnisse d. Math. 3. Folge, vol. 10, Springer, Berlin, 1987. 8. P. Bizo´ n and A. Wasserman, On existence of mini-boson stars, Commun. Math. Phys. 215 (2000), 357–373, arXiv:gr-qc/0002034. MR 1799851 9. J. Bjoraker and Y. Hosotani, Monopoles, dyons and black holes in the four-dimensional Einstein-Yang-Mills theory, Phys. Rev. D62 (2000), 043513, arXiv:hep-th/0002098. 10. J.L. Bl´ azquez-Salcedo, J. Kunz, F. Navarro-L´erida, and E. Radu, Charged rotating black holes in Einstein–Maxwell–Chern-Simons theory with negative cosmological constant, (2016), arXiv:1610.05282 [gr-qc]. 11. J.L Bl´ azquez-Salcedo, J Kunz, F Navarro-L´erida, and E Radu, Static EinsteinMaxwell Magnetic Solitons and Black Holes in an Odd Dimensional AdS Spacetime, Entropy 18 (2016), 438. 12. G. Carron and E. Pedon, On the differential form spectrum of hyperbolic manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 3 (2004), 705–747. MR 2124586 13. S.S. Chern and J. Simons, Characteristic forms and geometric invariants, Ann. of Math. (2) 99 (1974), 48–69. MR 0353327 14. P.T. Chru´sciel and E. Delay, Non-singular, vacuum, stationary space-times with a negative cosmological constant, Ann. Henri Poincar´e 8 (2007), 219–239. MR MR2314449 15. P.T. Chru´sciel and E. Delay, Non-singular spacetimes with a negative cosmological constant: II. Static solutions of the Einstein-Maxwell equations, (2016), arXiv:1612.00281 [math.DG]. 16. P.T. Chru´sciel, E. Delay, and P. Klinger, Non-singular spacetimes with a negative cosmological constant: IV. Stationary black hole solutions, in preparation. 17. R. Coquereaux and A. Jadczyk, Riemannian geometry, fiber bundles, Kaluza-Klein theories and all that, World Sci. Lect. Notes Phys., vol. 16, World Scientific Publishing Co., Singapore, 1988. MR MR940468 (89e:53108) 18. O.J.C. Dias, G.T. Horowitz, and J.E. Santos, Black holes with only one Killing field, Jour. High Energy Phys. (2011), 115, 43, arXiv:1105.4167 [hep-th]. MR 2875937 19. Y. Four`es-Bruhat, Th´eor`eme d’existence pour certains syst`emes d’´equations aux d´eriv´ees partielles non lin´eaires, Acta Math. 88 (1952), 141–225. 20. G.W. Gibbons, G.T. Horowitz, and P.K. Townsend, Higher-dimensional resolution of dilatonic black-hole singularities, Class. Quantum Grav. 12 (1995), 297–317, arXiv:hep-th/9410073. MR 1326239 21. D.J. Kaup, Klein-Gordon Geon, Phys. Rev. 172 (1968), 1331–1342. 22. B.S. Kim, Holographic Renormalization of Einstein-Maxwell-Dilaton Theories, Jour. High Energy Phys. 11 (2016), 044, arXiv:1608.06252 [hep-th].

STATIONARY SPACETIMES WITH NEGATIVE Λ

23

23. J.M. Lee, Fredholm operators and Einstein metrics on conformally compact manifolds, Mem. Amer. Math. Soc. 183 (2006), vi+83, arXiv:math.DG/0105046. MR MR2252687 24. J.G. Ratcliffe, Foundations of hyperbolic manifolds, second ed., Graduate Texts in Mathematics, vol. 149, Springer, New York, 2006. MR 2249478 25. T. Torii, K. Maeda, and M. Narita, Scalar hair on the black hole in asymptotically anti-de Sitter spacetime, Phys. Rev. D (3) 64 (2001), 044007, 9. MR 1853978 Piotr T. Chru´sciel, Faculty of Physics, University of Vienna, Boltzmanngasse 5, A1090 Wien, Austria E-mail address: [email protected] URL: http://homepage.univie.ac.at/piotr.chrusciel/ ´, Laboratoire Erwann Delay, Avignon Universite d’Avignon (EA 2151) F-84916 Avignon E-mail address: [email protected] URL: http://www.math.univ-avignon.fr

de Math´ ematiques

Paul Klinger, Faculty of Physics, University of Vienna, Boltzmanngasse 5, A1090 Wien, Austria E-mail address: [email protected]