Singular Harmonic Maps in Relativity
Singular Harmonic Maps in Relativity
BY LUC L. NGUYEN
by Luc L. Nguyen
Dissertation Director: YanYan Li
Harmonic maps with singular boundary behavior from a Euclidean domain into hyper-
bolic spaces arise naturally in the study of axially symmetric and stationary spacetimes
in general relativity. In particular, the study of multi-black-hole configurations and the
force between co-axially rotating black holes requires, as a first step, an analysis on the
boundary regularity of the “next order term” of those harmonic maps. We carry out
this analysis by considering those harmonic maps as solutions to some homogeneous
divergence systems of partial differential equations with singular coefficients. We then
apply our result to study the regularity of axially symmetric and stationary electrovac
spacetimes, which extends previous works by Weinstein [22], [23] and by Li and Tian
[10], [11], [12]. This dissertation is based on a preprint of the author [16].
ii
Acknowledgements
I would like to thank Professor YanYan Li who suggested me this problem as the subject
of a Ph.D. dissertation and has given much advice since. I wish to thank Professors
Gang Tian and Gilbert Weinstein for many insightful discussions and for their support
and encouragement. I am grateful to Professors Piotr Chruściel and Michael Kiessling
whose comments on the draft help improve the presentation of this dissertation. I am
indebted to Professors Abbas Bahri, Haı̈m Brezis and Zheng-Chao Han for their interest
in this work and their willingness to listen to me several times. I thank Professor Penny
D. Smith for letting me know of [20]. Finally, I dedicate this dissertation to my parents,
Bô´ Minh and Me. Sen, and my wife, Duyên, without whose supports this work could
not have been done.
This research was funded in part by a grant from the Vietnam Education Foundation
(VEF)1 and the Rutgers University and Louis Bevier Dissertation Fellowship.
1
The opinions, findings, and conclusions stated herein are those of the author and do not necessarily
reflect those of VEF.
iii
Table of Contents
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
4. Hölder regularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1,p
Appendix B. The space WΣ (Ω, w) . . . . . . . . . . . . . . . . . . . . . . . 43
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Vita . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
iv
1
Chapter 1
Introduction
To describe the reduction used by Weinstein, we first introduce the notion of singular
harmonic maps (see [25]). Let Γ be a subset of the z-axis in R3 obtained by removing
some bounded line segments. Let h be the Newtonian potential created by a charge
distribution of strictly positive density along Γ. Note that h is a harmonic function on
2
R3 \ Γ and h ‘behaves like’ some negative multiple of log ρ near an interior point of Γ
where ρ is the distance to the z-axis. Let H be either the real or the complex hyperbolic
plane. For HR , we use the standard half-plane model {(X, Y ) : X > 0} with the metric
ds2 = X −2 (dX 2 + dY 2 ).
(See Appendix A for a derivation of this line element from the standard disk model
of HC .) Then given a geodesic ζ in H, ζ ◦ h is a harmonic map from R3 \ Γ into H.
Moreover, as x → Γ, ζ ◦ h(x) approaches the ideal point ζ(+∞) ∈ ∂H. Recall that a
map Φ : Ω ⊂ R3 → H is harmonic if it satisfies in local coordinates the equations
The Ernst-Geroch reduction formulation states that every axially symmetric, sta-
tionary exterior solution of the Einstein vacuum equation can be conveniently put in
the form (see [22] or [8], for example)
ρ2 2 1 2µ 2
ds2 = − dt + X(dϕ − p dt)2 + e (dρ + dz 2 )
X X
where ρ is the distance to the z-axis, ϕ is the cylindrical angle around the z-axis, and X,
p and µ are functions on R3 \ Γ which are determined by the following four conditions.
3
(ii) X is the first component of some axially symmetric singular harmonic map (X, Y )
from R3 \Γ into the real hyperbolic plane HR which is controlled by the Newtonian
potential created by a uniform line charge distribution h of unit density and some
ideal points on ∂HR . In particular, (X, Y ) satisfies the harmonic map equations
|DX|2 − |DY |2
∆X = , (1.1)
X
2DX · DY
∆Y = . (1.2)
X
(iii) p satisfies
ρ ρ
dp = − Yz dρ + 2 Yρ dz. (1.3)
X2 X
(iv) µ satisfies
ρ ρ
dµ = (X 2 − Xz2 + Yρ2 − Yz2 ) dρ + (Xρ Xz + Yρ Yz ) dz. (1.4)
4X 2 ρ 2X 2
Notice that the harmonic map equations (1.1) and (1.2) give the integrability conditions
needed to integrate (1.3) and (1.4).
Using this reduction and a variational approach, Weinstein proved the existence
of a family of (possibly singular) spacetimes which can be interpreted as equilibrium
configurations of asymptotically flat co-axially rotating vacuum black holes ([22], [23]).
Moreover, he showed that this family is uniquely parametrized by 3n − 1 parameters (n
is the number of black holes) which are interpreted as the masses and angular momenta
of the black holes, and the distances between them. This result implies in particular
Robinson’s theorem on the uniqueness of the Kerr solutions [18], [19].
After a first look at the reduction, one might have the impression that given any
solution to the singular harmonic map equations (with a right rate of singularity, of
course), one could easily cook up a solution to the Einstein vacuum equations. This is
in fact not the case, which was probably first observed by Bach and Weyl. As one tries
to construct a spacetime out of a singular harmonic map, one might possibly introduce
a conical singularity along Γ. A necessary and sufficient condition for regularity is
Bach and Weyl showed in [1] that this limit is nonzero in the static setting, i.e. Y ≡
0. Their method does not seem welcoming of the general setting since it relies on the
explicit form of the solution. Even though there has been some progress in getting the
explicit form of solutions for multiple-body spacetimes, e.g. the famous double Kerr
solutions of Kramer and Neugebauer [9], the dependence of the validity or invalidity of
(1.5) on the parameters seems still unclear in general.
Concerning the equation (1.5) itself, some regularity structure across the symmetry
axis Γ of the harmonic map (X, Y ) is required in order to make sense of the limit on the
left hand side. Note that because of the singularity of h, (X, Y ) necessarily approaches
∂HR near Γ. Therefore, the equations (1.1)-(1.2) are not satisfied across Γ, and so
one cannot apply directly well-known results on the regularity of harmonic maps into
negatively curved targets to obtain the required regularity for X and Y near Γ. This
regularity issue was settled independently by Weinstein and by Li and Tian. Weinstein
showed that, about any interior point of Γ, any singular harmonic maps corresponding
to the spacetimes he constructed in [22] and [23] can be “decomposed” into an explicit
singular part and a C ∞ -regular part. Independently, using a different approach, Li
and Tian [11], [12] proved a weaker version: C k,α regularity for the regular part, which
suffices to justify the limit on the left hand side of (1.5). On the other hand, their result
remains valid even if the harmonic map (X, Y ) is not axisymmetric or the singular
control h is the potential of a uniform charge distribution of arbitrary positive density
along Γ. To describe their results more precisely, we define the following weighted
spaces.
Note that these two spaces are Hilbert spaces. Also, if w is bounded from below by a
strictly positive number, they are naturally embedded into the familiar Sobolev spaces
H 1 (Ω) and H01 (Ω), respectively.
Li and Tian [11], [12] and Weinstein [22], [23] proved the following.
Remark 1.1 In the above theorem, we can allow k + α = 4γ. See Corollary 3.1 and
Remark 3.2.
After settling the regularity of the reduced harmonic maps, Li and Tian [11], [12],
[10] and Weinstein [24] went forward to study if there is a conical singularity in an
axially symmetric stationary vacuum spacetime. As mentioned above, such conical
singularity exists if and only if equation (1.5) is violated. Using Theorem A, the two
works showed that the limit on the left hand side of equation (1.5) exists and depends
continuously on the masses, angular momenta and distances between two bodies. More
6
importantly, they proved that there are several continuous sets of parameters which give
rise to spacetimes violating (1.5). Unfortunately, their results do not reveal whether
there is any set of parameter that realizes (1.5) except those corresponding to the Kerr
spacetimes, i.e. one-body spacetimes.
Later, in an attempt to shed more insight into the problem, Weinstein started
analyzing a generalization of the problem for Einstein-Maxwell equations ([25], [26],
[27]). It should be noted that there exists regular axially symmetric stationary charged
spacetimes which have more than one black hole: the Papapetrou-Majumdar solutions
[17], [13], [7]. However, these have degenerate event horizons. It is not known if there
are any regular charged spacetimes having non-degenerate event horizons and more
than one body.
Similar to the case of vacuum, by the Ernst-Geroch formulation, one can write the
metric of any axially symmetric stationary charged spacetime in the form (see [27], [8])
where ρ is again the distance to the z-axis, ϕ is the cylindrical angle around the z-axis
and u, p and µ are determined by
(ii) u is the first component of some axially symmetric singular harmonic map (u, v, χ, ψ)
from R3 \ Γ into the complex hyperbolic plane HC , i.e. it satisfies
moreover, it is singular near Γ and the singular rate is controlled by the Newtonian
h 1
potential creaated by a uniform line charge distribution 2 of density 2 and some
ideal points on ∂HC ;
7
(iii) p satisfies
e4u
dµ = ρ[(u2ρ − u2z ) + ((vρ − ψ χρ + χ ψρ )2 − (vz − ψ χz + χ ψz )2 )
4
+ e2u (χ2ρ − χ2z + ψρ2 − ψz2 )]dρ
e4u
+ 2ρ[uρ uz + (vρ − ψ χρ + χ ψρ )(vz − ψ χz + χ ψz )
4
+ e2u (χρ χz + ψρ ψz )]dz. (1.11)
Again, the harmonic map equations (1.6)-(1.9) are the integrability conditions for (1.10)
and (1.11). Moreover, in the absence of charge, i.e. χ = ψ = 0, the system (1.6)-(1.9)
reduces to the system (1.1)-(1.2) via the transformation X = e−2u and Y = 2v.
The existence problem for the asymptotically flat case was settled by Weinstein him-
self. Physically speaking, this solution represents (possibly singular) asymptotically flat
co-axially rotating electro-vacuum, or charged, black holes in gravitational equilibrium.
Moreover, he also showed that this family of solutions is parametrized by 4n − 1 pa-
rameters, n being the number of black holes, which can be interpreted as the masses,
angular momenta and charges of the black holes, and the distances between them. Es-
pecially, when n = 1, he recovered the uniqueness of the Kerr-Newman solutions which
was proved independently by Mazur and Bunting ([14], [15], [2]).
Here we would like to address the regularity of the harmonic maps obtained by
performing the Ernst-Geroch reduction to the solutions constructed by Weinstein. In
this work, we prove:
Theorem B Let Γ be the z-axis in R3 with some bounded line segments removed and
let ρ be the distance function to the z-axis. Let h be the potential of a uniform charged
distribution of density γ > 0. If (u, v, χ, ψ) is a singular harmonic map from R3 \ Γ into
h
the complex hyperbolic plane HC controlled by 2 and some ideal points on ∂HC and if
h
u− 2 ∈ H 1 (R3 ), v ∈ H 1 (R3 , e2h ), and χ, ψ ∈ H 1 (R3 , eh ), then u − h2 , v χ, ψ ∈ C k,α
8
across the interior of Γ for any k + α < 2γ. In addition, near any compact subset of
the interior of any component of Γ, v, χ and ψ admit the asymptotic expansion
v = C1 − C2 χ + C1 ψ + O(ρ2k+2α ),
χ = C2 + O(ρk+α ),
ψ = C2 + O(ρk+α ),
Again, we emphasize that this theorem does not rule out the possibility of having
conical singularity along the symmetry axis. For the metric to be regular across the
axis, one needs
lim (µ + 2u + log ρ) = 0 along Γ. (1.12)
ρ→0
In view of Theorem B, one can verify easily that the limit on the left hand side exists.
It is then possible to carry out an analysis similar to that in [11], [12] and [24] to
prove nonexistence of regular spacetimes corresponding to certain class of parameters.
However, we will not pursue this direction in the present work.
9
Chapter 2
In R3 , let Γ be the z-axis with some line segments removed. Let h be the Newtonian
potential created by a uniform line charge distribution of density γ > 0 along Γ. We
would like to understand the regularity of a singular harmonic map (u, v, χ, ψ) from
R3 \ Γ into HC controlled by h. Recall that (u, v, χ, ψ) satisfies (1.6)-(1.9),
Since HC has negative sectional curvature, standard harmonic map theory implies that
(u, v, χ, ψ) is smooth away from the singularity set Γ. Thus we only need to study its
behavior near Γ. Also, as we are only interested in the interior of Γ, it suffices to restrict
our attention to a subset Ω of R3 such that Σ = Ω ∩ Γ has only one component whose
endpoints lie on ∂Ω. Also, we only need to pick one controlling ideal point p ∈ ∂HC .
Using an isometry of HC , we can assume without loss of generality that p is the +∞
point sitting on the u-axis of HC . We pick the geodesic going to p to be
t 7→ ζ(t) = (t, 0, 0, 0) ∈ HC .
We have shown:
This a priori estimate shows that the harmonic map equations (1.6)-(1.9) is a sin-
gular semilinear elliptic system. As h ‘behaves like’ −2γ log ρ near Σ, the singularity
types are negative powers of the distance function to a line.
(In the vacuum case, this divergence structure is more apparent (cf. [21], Chapter 7).)
The drawback of this way of writing the harmonic map equations is that the coefficients
cαβ might not satisfy the Legendre-Hadamard condition. Indeed, for ξ = (ξ 1 = a, ξ 2 =
1, ξ 3 = 0, ξ 4 = 0) ∈ R4 ,
cαβ ξ α ξ β = a2 − 2e4u v a + e4u .
It is readily seen that when e2u v > 1, the right hand side changes sign as a varies in R.
Nevertheless, when the a priori estimate (2.1) holds, the Legendre-Hadamard condition
can be recovered (for a different but equivalent set of coefficients). To this end we
rewrite (2.2)-(2.5) in the form
In view of (2.1), we can always pick l sufficiently large and positive λ, Λ such that
Let d = dΣ denote the distance function to Σ. Assume that there exist constants
k(1), k(2), . . . , k(m) ≥ 0 such that for any given R > 0, there exist λ = λ(R) > 0 and
Λ = Λ(R) such that
m
X m
X
λ d(x)−2k(α) |ξ α |2 ≤ aαβ (x, u) ξ α ξ β ≤ Λ d(x)−2k(α) |ξ α |2 (2.7)
α=1 α=1
Assume that u solves (2.6) in some appropriate sense. Determine how regular u is!
Remark 2.1 Analogous to the case of harmonic maps into the complex hyperbolic
plane, regularity for harmonic maps with prescribed singularity into any real, complex,
or quaternionic hyperbolic space in the sense described in [25] can always be recast as
a part of Problem 1. Consequently, the result in Theorem B extends parallelly to those
cases.
1 (Ω, d−2k(α) ).
for any ξ α ∈ H0,Σ
14
Chapter 3
In this chapter, we consider a special case of Problem 1 where the unknown is a scalar
map and the coefficient is independent of the unknown. Recall that Ω is an open subset
of Rn and Σ is a (n−k)-dimensional submanifold of Ω with k ≥ 2. Let d be the distance
function to Σ and w be a weight that satisfies
for some γ > 0, 0 < λ ≤ Λ < ∞. We are interested in the regularity of weak solutions
of
div (w(x)Du = 0 in Ω \ Σ. (3.1)
Recall that by weak solution we mean a function u ∈ HΣ1 (Ω, d−γ ) such that
Z
1
w(x) Du Dξ dx = 0 for any ξ ∈ H0,Σ (Ω, d−γ ).
Ω
This problem has been studied extensively in the literature. When γ < k − 2, w
belongs to the Muckenhoupt class A2 and the result of Fabes, Kenig and Serapioni
[4] implies that u is Hölder continuous across Σ. Unfortunately, in our application to
the problem from general relativity, w might get too singular in a way that it is not
even integrable across Σ, and so their work does not apply directly. Moreover, for
other purposes, we are also interested in whether w vanishes along Σ and how fast it
decays there. A result in this direction was given by Li and Tian in [11] when Σ has
codimension 2. We generalize their work to the case where Σ is a general submanifold
of any codimension k ≥ 2 and sharpen the decay estimate to its optimal form.
λ d−γ ≤ w ≤ Λ d−γ
15
for some γ ≥ 0 and 0 < λ ≤ Λ < ∞. Moreover, assume that |D(dγ w)| = o(d−1 ) in a
neighborhood of Σ.
div (w Du) = 0
+ γ
sup d−(γ−k+2) |u| ≤ C kd− 2 ukL2 (Ω)
ω
for any ω compactly supported in Ω. The constant C depends only on Λ, λ and the
modulus of continuity of d|D(dγ w)|.
Remark 3.1 To see that the decay estimate is optimal, consider the case where Σ is
an n − k hyperplane and w = d−γ .
When γ > k−2, (3.1) has a special solution u = dγ−k+2 which belongs to HΣ1 (Ω, d−γ ).
This solution vanishes along Σ and exhibits Hölder continuity along Σ.
When γ < k−2, the above special solution does not belong to HΣ1 (Ω, d−γ ). Moreover,
constant functions are solutions which are in HΣ1 (Ω, d−γ ) and, of course, they need not
vanish along Σ. Nevertheless, as mentioned above, w is in the Muckenhoupt class A2
and so u is Hölder continuous across Σ.
Before proving Theorem 3.1, let us give an application which improves Theorem A.
Y = C + O(ρ4γ )
Rewrite (1.2) as
div (X −2 DY ) = 0
and apply Theorem 3.1, we get the required decay for Y and then its C k,α regularity.
To get the regularity for X, we rewrite (1.1) as
|DY |2
∆(log X + h) = −
X2
Remark 3.2 To see that the C k,α regularity in Corollary 3.1 is optimal, consider for
example the case where Γ is the whole z-axis, h = −2γ log ρ, γ > 0 and
ρ2γ
X= ,
ρ4γ + 1
ρ4γ
Y = 4γ .
ρ +1
This example also shows that the C k,α regularity in Theorem B is almost optimal as a
harmonic map into the real hyperbolic plane is a special harmonic map into the complex
hyperbolic plane.
We will prove Theorem 3.1 through a sequence of lemmas. Also, to avoid technical-
ity, we will assume that Σ is a (n − k)-hyperplane and w = d−γ . In fact, if w̄ = dγ w is
not constant, the proofs below require minor changes.
Also, we will frequently use the following inequality (see Appendix B) without ex-
plicitly mentioning,
Z Z
−γ−2
d 2
|f | dx ≤ C [|f |2 + d−γ |Df |2 ] dx
ω Ω
Since γ < 2(k − 2), w̃ belongs to the Muckenhoupt class A2 and the above equation
is satisfied across Ω. Again, the result of Fabes et al. in [4] applies showing that v is
locally bounded in Ω and for any ω compactly supported in Ω,
γ γ
sup |v| ≤ C kd 2 −k+2 vkL2 (Ω) ≤ C kd− 2 vkL2 (Ω) .
ω
Lemma 3.2 Assume that γ ≥ 2(k − 2) and u is as in Theorem 3.1. Then u satisfies
γ γ
sup d− 2 |u| ≤ C kd− 2 ukL2 (Ω)
ω
γ
Proof. Write v = d− 2 u. Then v ∈ H 1 (Ω) ∩ L2 (Ω, d−2 ) and satisfies in Ω \ Σ the
equation
∆v − c v = 0
where
γ γ 1
c= ( − k + 2) 2 .
2 2 d
Note that c is positive. We therefore can apply De Giorgi or Moser techniques to show
that v is locally bounded and
Lemma 3.3 Assume that u ∈ H 1 (Ω)∩C 0 (Ω̄) solves the following inhomogeneous equa-
tion in Ω \ Σ
div (w(x)Du) = f, (3.2)
Proof. Fix some ω compactly supported in Ω. Fix some 0 < h < 1. Assume for the
moment that d−c u is bounded. We will show that if c + h ≤ µ then
We compute
Dθ = (c + h) dc+h−1 Dd + c dc−1 |x − x0 |2 Dd + 2 dc (x − x0 ),
+ c(c − γ + k − 2) d−γ+c−2 |x − x0 |2
Lemma 3.4 Let v ∈ H 1 (Ω) ∩ L2 (Ω, d−2 ) solve the following equation in Ω \ Σ:
Dd 1
∆v − p · Dv − q 2 v = 0 (3.3)
d d
• q ≥ 0,
• (p − k + 2)2 + 4q − p2 > 0,
2q 4q 2 4q(p−k+2)
• and p − k + 2 ≤ k−2 or (k−2)2
− k−2 + p2 − 4q < 0.
Then v ∈ L∞
loc (Ω) and for any ω compactly supported in Ω,
Remark 3.3 The classical De Giorgi - Nash - Moser estimate does not apply here as
d−1 ∈
/ L1,n−1+ and d−2 ∈
/ L1,n−2+ for any > 0.
Proof. We can assume that Ω is the unit ball B1 . We will show that v is bounded in
B1/2 . The idea is to adapt De Giorgi’s proof exploiting the fact that q is negative.
where
(p − b)2
α= (k−2)b
+ .
4(q − 2 − 0 )
2q
Our hypotheses on p and q are exactly to allow us to find some b < k−2 so that α < 1.
We have therefore shown that
Z Z
2 2
η |Du| dx ≤ C |u − k|2 |Dη|2 dx. (3.5)
A(k,ρ) A(k,ρ)
Using this estimate, we can run through De Giorgi’s proof for local boundedness and
obtain
sup u ≤ C ku+ kL2 (B1 ) .
B1/2
Proof of Theorem 3.1. The case 0 ≤ γ < 2(k − 2) is taken care by Lemma 3.1. We
hence assume that γ ≥ 2(k − 2).
γ
By Lemma 3.2, d− 2 u is locally bounded. Thus, by elliptic theory, u is continuous
in Ω and vanishes along Σ. Lemma 3.3 hence shows that d−λ u is locally bounded for
any λ < γ − k + 2. By elliptic theory, we infer that d−λ+1 |Du| is locally bounded for
the same λ’s.
Dd 1
∆vm − pm · Dvm − qm 2 vm = 0
d d
21
sup d−λm |u| = sup |vm | ≤ Cm (ω) kvm kL2 (Ω) = Cm (ω) kd−λm ukL2 (Ω) .
ω ω
and
2qm
pm − k + 2 − → −γ + k − 2 < 0.
k−2
As a consequence, if ω ⊂⊂ ω 0 ⊂⊂ Ω,
Chapter 4
Hölder regularity
Throughout the chapter, the coefficients aαβ are assumed to satisfy the ellipticity con-
dition (2.7), i.e. for any R > 0, there exists λ = λ(R) > 0 and Λ = Λ(R) < ∞ such
that
m
X m
X
−2k(α)
λ d(x) α 2
|ξ | ≤ aαβ (x, u) ξ ξ ≤ Λ α β
d(x)−2k(α) |ξ α |2
α=1 α=1
In general, without any additional assumption on the coefficients aαβ or the un-
knowns uα , one does not expect uα to be regular, or even Hölder continuous, across Σ.
For example, when the coefficients aαβ satisfies the classical ellipticity, i.e. k(α) = 0,
it is well-known that u satisfies (2.6) in Ω in the weak sense and so is regular at any
point x ∈ Σ at which
Z
1
lim inf |Du|2 dx = 0.
δ→0 δ n−2 Bδ (x)
u, the points where regularity fails might constitute a non-vacuous subset of Σ (see [5],
Chapter 2).
23
We show that the above phenomenon also persists for the problem we are consider-
ing. For any Bδ (x) ⊂ Ω, we define
Z m
1 X
Eδ (x) = d(z)−2k(α) |Duα (z)|2 dz. (4.1)
δ n−2 Bδ (x) α=1
Also, define
If uα ∈ L∞ (Ω, d−k(α) ), then there exists ε0 > 0 such that if Eδ (x0 ) ≤ ε0 for some
0 < δ < dist (x0 , ∂Ω)/4, then
for some λ = λ(k(α), kuα kL∞ (Ω,d−k(α) ) ) > 0 and C = C(k(α), kuα kL∞ (Ω,d−k(α) ) ) ≥ 0. In
particular, u is Hölder continuous in Bδ/4 (x0 ) and for α ≥ τ , d−k(α)−λ |uα | is bounded
in Bδ/4 (x0 ).
Remark 4.1 The requirement that k(τ ) > k − 2 is probably merely technical. It would
be interesting to remove this extra hypothesis. However, in doing so, one must be careful
with the assumption that d−k(α) uα is bounded. For, in case of a single linear equation,
if k(α) < k − 2, there are examples that uα may not vanish along Σ as fast as dk(α)
(see Remark 3.1).
[11]. It seems that their proof does not apply in our context. However, as demonstrated
in Chapter 2, our harmonic map problem has an equivalent homogeneous divergence
form. This special structure makes it apparent to obtain a Caccioppoli inequality which
we will state shortly. When restricted to the vacuum case in [11], this Caccioppoli
inequality implies the monotonicity formula therein.
Observe that, as long as we stay away from Σ, (2.6) is an elliptic system for u with
bounded smooth coefficients, so classical elliptic theory tells us that a Caccioppoli in-
equality holds there. However, as we move towards Σ, the coefficients of these equations
behave badly signifying some possible deterioration in the structure of the Caccioppoli
inequality. This is indeed true as stated in the following result.
Proposition 4.1 (Caccioppoli inequality) Let uα ∈ HΣ1 (Ω, d−2k(α) )∩L∞ (Ω, d−k(α) )
be a weak solution of (2.6). Let x0 be a point in Ω and δ0 the distance from x0 to ∂Ω.
There exists C = C(Λ, λ) > 0 such that for δ ≤ δ0 /2 and b ∈ Rm whose last m − τ + 1
components vanish when B2δ (x0 ) ∩ Σ 6= ∅,
Z m Z m
X
−2k(α) C X
d α 2
|Du | dz ≤ 2 d−2k(α) |uα − bα |2 dz.
Bδ/2 (x0 ) δ Bδ (x0 )
α=1 α=1
Proof. Fix ε > 0 and δ < δ0 . Let η : Rn → R be a cut-off function satisfying η(z) = 0
if |z − x0 | ≥ δ and η(z) = 1 if |z − x0 | ≤ δ/2, and |Dη| ≤ 4δ −1 .
Observe that our constraints on b imply that uα − bα ∈ HΣ1 (Ω, d−2k(α) ). Thus, by
Corollary B.1 in Appendix B, we can take ξ = η 2 (u − b) as a test function for (2.6).
Using this special choice of ξ and recalling (2.7) yields
Z m
X Z
2 −2k(α) α 2
η d |Du | dz ≤ C η 2 aαβ Duα Duβ dz
Ω α=1 Ω
Z
= −C ηaαβ Duβ Dη (uα − bα ) dz
Ω
m h
Z X i
≤C η 2 d−2k(α) |Duα |2 + |Dη|2 d−2k(α) |uα − bα |2 dz.
Ω α=1
We next study the limiting system when doing a blow-up analysis and then prove
that smallness in energy density implies regularity. We first recall a well known result
(see [5], Chapter 4).
25
λ|p|2 ≤ aij α β 2
αβ(h) (x, v)pi pj ≤ Λ|p| , (x, v, p) ∈ Ω × Rm × Rmn
ticity bound. Let f(h) be a sequence in L2 (B1 (0); Rm ) converging weakly in L2 (B1 (0))
1 (B (0); Rm ) ∩ L2 (B (0)) such that
to f . Let u(h) be a sequence in Hloc 1 1
β
∂ ij
∂u(h)
α
a = f(h)
∂xi αβ(h) ∂xj
• for any ρ < 1, u(h) converges strongly to u in L2 (Bρ (0); Rm ) and Du(h) converges
weakly to Du in L2 (Bρ (0); Rm ),
• and more importantly, u satisfies in the weak sense in B1 (0) the system
∂ β
ij ∂u
a = f.
∂xi αβ ∂xj
Lemma 4.2 Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) ) be a weak solution to (2.6). Let
d(x(h) )
B(h) = Bδ(h) (x(h) ) be a sequence of balls in Ω such that the ratios δ(h) are all less
than some fixed κ. Let Σ(h) be the image of Σ under the map x 7→ (x − x(h) )/δ(h) and
d(h) the distance function to Σ(h) . Assume that Σ(h) stabilizes to some Σ∗ , i.e. d(h)
converges uniformly away from Σ∗ to d∗ , the distance function to Σ∗ .
1
uα(h) (x) = k(α)
[uα (x(h) + δ(h) x) − ūα(h) ], x ∈ B1 (0),
ε(h) δ(h)
where
the average of uα over B if k(α) = 0,
(h)
ūα(h) =
0 otherwise.
quence,
• uα(h) converges weakly in H 1 (B1/2 (0)) and strongly in L2 (B1/2 (0)) to uα∗ ,
26
Moreover, u∗ satisfies
−[k(α)+k(β)]
bαβ∗ Duβ = 0
div d∗
in the weak sense in B1/2 (0) \ Σ∗ , where bαβ∗ are constant and satisfy
m m
−2k(α) α 2 −[k(α)+k(β)] −2k(α)
X X
ν1 d∗ |p | ≤ d∗ bαβ∗ pα pβ ≤ ν2 d∗ |pα |2 ,
α=1 α=1
and the ratio ν2 /ν1 does not depend on the sequence of balls Bδ(h) (x(h) ). Additionally,
bαβ∗ can be written in the form
Proof. Let’s assume for the moment that the above convergences have been established.
−k(α) −k(α)
Passing to a subsequence, we can assume that xh → x∗ , d(h) uα(h) → d∗ uα∗ a.e.
Also, as
Z
−2k(α) α 2 C
δ(h) |ū(h) | ≤ n d−2k(α) |uα |2 dz ≤ C,
δ(h) B(h)
−k(α)
we can assume that δ(h) ūα(h) → ũα∗ . It follows that
Therefore, if we define
−[k(α)+k(β)]
aαβ(h) (x) = δ(h) aαβ x(h) + δ(h) x, u(x(h) + δ(h) x)
−[k(α)+k(β)]
= d(h) bαβ x(h) + δ(h) x, ũ(x(h) + δ(h) x) .
then
aαβ(h) (x) → d∗ (x)−[k(α)+k(β)] bαβ x∗ , ũα∗ d∗ (x)−k(α) a.e.
Here we have used the assumption that Eδ(h) (x(h) ) = ε2(h) . In particular, this implies
that the L2 norm of Duα(h) over B1 (0) is uniformly bounded for α < τ . The standard
Poincaré inequality then implies that the H 1 norm of uα(h) over B1 (0) is also uniformly
bounded for α < τ . For α ≥ τ , we use Corollary B.1 in Appendix B to show that the
L2 norm and so the HΣ1 norm of uα(h) over B3/4 (0) is uniformly bounded. One can then
modify the proof of Proposition B.2 in Appendix B to get the required convergences.
The details work out in exactly the same manner and hence are omitted.
Lemma 4.3 Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) ) be a weak solution to (2.6). Let
d(x(h) )
B(h) = Bδ(h) (x(h) ) be a sequence of balls in Ω such that the ratios δ(h) are all at least
some fixed κ.
1
uα(h) (x) = k(α)
[uα (x(h) + δ(h) x) − ūα(h) ], x ∈ B1 (0),
ε(h) d(x(h) )
0
where ūα(h) is the average of uα over B(h) = Bκδ(h) /2 (x(h) ).
in the weak sense in B1 (0) \ Σ∗ , where aαβ∗ are smooth and satisfy
m
X m
X
α 2 α β
ν1 |p | ≤ aαβ∗ p p ≤ ν2 |pα |2 ,
α=1 α=1
and the ratio ν2 /ν1 does not depend on the sequence of balls Bδ(h) (x(h) ) and does not
exceed C κ−2 max k(α) .
Proof. The proof of this lemma is similar to but easier than that of Lemma 4.2 and is
omitted.
28
Lemma 4.4 Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) ) be a weak solution to (2.6).
Assume in addition that (4.2) holds and all the k(α) are either zero or bigger than
k − 2. Let x0 be a point in Ω and δ0 = dist (x0 , ∂Ω). There exists ε0 and λ0 such that
for any x ∈ Bδ0 /2 (x0 ) and δ ∈ (0, δ0 /4) satisfying Eδ (x) ≤ ε20 there holds Eλ0 δ (x) ≤
Eδ (x)/2.
Proof. Arguing indirectly, we assume that the conclusion fails. Fix some λ0 and κ for
the moment. They will be determined by a set of constraints which will be formulated
in the sequel. We can find sequences ε(h) → 0, x(h) ∈ Bδ0 /2 (x0 ), and δ(h) ∈ (0, δ0 /4)
such that Eδ(h) (x(h) ) = ε2(h) but Eλ0 δ(h) (x(h) ) > ε2(h) /2. In addition, we can assume that
d(x(h) )
one of the following two cases occurs: All the ratios δ(h) are less than κ or all are
not.
d(x(h) )
Case 1: δ(h) < κ for all h.
and
|uα∗ (x)| ≤ C d(x)2k(α)−k+2 α≥τ
Hence if
1
C1 λ20 + C2 (κ + 2λ0 )2k(τ )−2k+4 ≤ , (4.4)
8
ε2(h) ε2
we infer that 2 < Eλ0 δ(h) (x(h) ) ≤ 4 for h large enough, a contradiction.
d(x(h) )
Case 2: δ(h) ≥ κ for all h.
with ν2 /ν1 ≤ C κ−2 max k(α) . Also, the L2 norm of u∗ in Bκ/2 (0) is universally bounded.
Therefore
m
X n+A
|uα∗ (x) − uα∗ (0)| ≤ C κ− 2 |x|, x ∈ Bκ/4 (0)
α=1
for some A > 0. Like case 1, the constant C is independent of the balls Bδ(h) (x(h) ).
30
1
C3 κ−n−A λ20 ≤ , (4.5)
8
To complete the proof, we need to furnish (4.4) and (4.5). This is always doable by
1 1
first picking λ0 small enough such that λ0 ≤ √1 1
and (8C3 λ20 ) n+A ≤ ( 16C ) B − 2λ0 ,
4 C1 2
and then picking κ in between the last two figures. Here B = 2k(τ ) − 2k + 4 > 0.
Proof of Theorem 4.1. Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) ) be a weak solution
of (2.6). Let ε0 be as in Lemma 4.4. Assume that Eδ (x0 ) ≤ ε0 for some 0 < δ <
dist (x0 , ∂Ω)/4.
It follows from Lemma 4.4 and standard iteration techniques (cf. [6]) that there
exists positive constants C and λ depending only on k(α) and the L∞ (Ω, d−k(α) )-norm
of uα such that
Z m
1 X
Eσ (x) = d−2k(α) |Duα |2 dz ≤ C σ 2λ , x ∈ Bδ/2 (x0 ), σ < δ/4.
σ n−2 Bσ (x) α=1
Moreover, if k(α) > 0, the above estimate implies a rate of vanishing of uα near Σ.
To see this, observe that if σ < d(x)/2, then
Z
|Duα |2 dz ≤ C d(x)2k(α) σ n−2+2λ , 1 ≤ α ≤ m,
Bσ (x)
31
which implies
osc uα ≤ C d(x)k(α) σ λ .
Bσ/2 (x)
(See Theorem 7.19 in [6], for example.) Since uα vanishes along Σ when k(α) > 0, this
implies that dk(α)−λ uα is bounded in Bδ/4 (x0 ). The proof is complete.
32
Chapter 5
Applications
We now turn to the proof of Theorem B and C. We first consider a special case of
Problem 1 in which the smallness assumption in Theorem 4.1 is fulfilled. The condition
we impose additionally on the system (2.6) is a common feature that harmonic map
problems into hyperbolic spaces share. This was first used by Li and Tian in [11] in the
case of real hyperbolic spaces. We doubt that this phenomenon has some connection
to the underlying geometry of the target manifolds, but we have very limited evidence.
Proposition 5.1 Suppose all the assumptions of Theorem 4.1 hold. Assume in addi-
tion that τ = 2. If
for some l(1) = 0, l(2), . . . , l(m) > 0, then the uα are Hölder continuous across Σ.
Moreover, for α > 1, |uα | ≤ C dk(α)+λ for some λ > 0.
First observe that, due to (5.1), the first equation in the system (2.6) can be rewrit-
ten as
X
∆u1 = l(α) aαβ Duα Duβ . (5.2)
2≤α,β≤m
Note that the right hand side, which we will abbreviate as F (x, u, Du), is always non-
negative by ellipticity.
Lemma 5.1 Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) ) be a weak solution to (2.6). Let
x0 be a point in Ω and δ0 = dist (x0 , ∂Ω). Then there exists K such that if δ < δ0 /4
and x ∈ Bδ0 /2 (x0 ),
Z m
X
|x − z|−n+2 d−2k(α) |Duα |2 dz ≤ K,
Bδ (x) α=1
33
and so
Z m
1 X
Eδ (x) = d−2k(α) |Duα |2 dz ≤ K.
δ n−2 Bδ (x) α=1
Proof. Note that u1 satisfies equation (5.2) in the sense of distributions throughout Ω.
and |ηε0 |, |ηε00 | ≤ C in [δ, 2δ], and |ηε0 | ≤ Cε−1 , |ηε00 | ≤ Cε−2 in [ε, 2ε].
Let R be a positive number such that |u1 | ≤ R. Let Gx (z) = |x − z|−n+2 and set
Gεx (z) = ηε (|x − z|)Gx (z). Inserting Gεx (z)(u1 + R + 1) as a test function into equation
(5.2) and note that F is nonnegative, we get
Z Z
Gεx (z) F (z, u, Du) dz ≤ F (z, u, Du) Gεx (z) (u1 + R + 1) dz
B2δ (x) B2δ (x)
Z
=− Du1 D[Gεx (z) (u1 + R + 1)] dz
B2δ (x)
Z h i
=− Gεx (z)|Du1 |2 + (u1 + R + 1) Du1 DGεx dz.
B2δ (x)
Hence
Z m
X
Gx (z) d−2k(α) |Duα |2 dz
Bδ (x)\B2ε (x) α=1
Z h i
≤C Gx (z) F (z, u, Du) + |Du1 |2 dz
Bδ (x)\B2ε (x)
Z
≤ −C (u1 + R + 1) Du1 DGεx dz.
B (x)
Z 2δ
=C (u1 + R + 1)2 ∆[ηε (|x − z|) Gx (z)] dz
B (x)
Z 2δ h i
≤C |ηε00 (|x − z|)| |x − z|−n+2 + |ηε0 (|x − z|)| |x − z|−n+1 dz
B2δ (x)
≤ K.
Proposition 5.2 (Smallness of density) Let uα ∈ HΣ1 (Ω, d−2k(α) ) ∩ L∞ (Ω, d−k(α) )
be a weak solution to (2.6). Let x0 be a point in M and δ0 = dist (x0 , ∂M ). Let K be
34
the constant obtained in Lemma 5.1. For given ε > 0, δ < δ0 /4, and x ∈ Bδ0 /2 (x0 ),
there exists σ = σ(ε, δ, x) between e−2K(n−2)/ε δ and δ such that
Eσ (x) ≤ ε.
We will use various equations derivable from (1.6)-(1.9). The first set gives the formulas
for the Laplacians of u, χ and ψ.
h
∆(u − ) = ∆u = 2e4u |Dv − ψ Dχ + χ Dψ|2 + e2u (|Dχ|2 + |Dψ|2 ), (5.3)
2
∆χ = −2Du · Dχ + 2e2u Dψ · (Dv − ψ Dχ + χ Dψ), (5.4)
The proof consists of two parts. In the first part, we prove the general case where
γ is arbitrary. In the second part, we consider the physical case where γ = 1 and
(u, v, χ, ψ) is axially symmetric around the z-axis.
For the first part, we apply Theorem 4.1 and Corollary 5.1 to obtain Hölder conti-
nuity. Unlike proving Corollary 3.1, we cannot apply Theorem 3.1 due to the coupling
of the unknowns. Instead, we repeatedly apply Lemma 3.3, a primitive of Theorem 3.1,
to (5.6)-(5.8) to get C k,α regularity.
For the second part, we follow the reversed Ernst-Geroch reduction scheme to prove
smoothness under axial symmetry. This was used by Weinstein in his work on the
vacuum case. His idea is the following. The system (1.1)-(1.2) was obtained by applying
the Ernst-Geroch formulation for axisymmetric stationary vacuum solutions, which was
done in the order of the axial Killing vector field being applied first and the stationary
Killing vector field second. This choice of order made it more convenient to establish
existence and uniqueness but regularity. If one applies the reduction scheme in the
reversed order, i.e. to the stationary Killing vector field first and then to the axial Killing
vector field, one obtains a harmonic map problem from R3 into the real hyperbolic plane
in the usual sense. Similarly, if one applies the reversed Ernst-Geroch reduction scheme
to the Einstein-Maxwell equations, the system acquired is, though no longer a harmonic
map problem, still elliptic. To finish one needs to prove some initial regularity for the
36
acquired system. In the vacuum case, this can be done much simpler by a clever use
of the De Giorgi-Nash-Moser regularity result for scalar elliptic equation (see [22]). It
seems unlikely that this technique applies in the electrovac case. It is precisely at this
point that we need to use the C k,α regularity obtained in the general case to go forward.
Proof of Theorem B. Note that since the complex hyperbolic plane is a manifold of
negative sectional curvature, (u, v, χ, ψ) is C ∞ in R3 \ Σ. Thus we only need to consider
the regularity question around Γ. Without loss of generality, we assume that the origin
is an interior point of Γ, h = −2γ log ρ and that the controlling ideal point on ∂HC is
(+∞, 0, 0, 0) as in Chapter 2.
The above estimates show that |Dψ| |Dv − ψ Dχ + χ Dψ| = O(ρ−2+3γ+2λ ). Thus,
γ
if λ < 2, an application of Lemma 3.3 to (5.7) shows that |χ| = O(ργ+2λ ). Similarly,
|ψ| = O(ργ+2λ ). Applying basic gradient estimates for Poisson equations to (5.7) and
(5.8), we deduce that |Dχ| + |Dψ| = O(ρ−1+γ+2λ ). Using equations (5.4) and (5.5)
we infer that |∆χ| + |∆ψ| = O(ρ−2+γ+2λ ). Hence the right hand side of (5.6) is at
most O(ρ−2−2γ+4λ ). Lemma 3.3 applies again, but to (5.6), yielding |v| = O(ρ2γ+4λ ).
Gradient estimates for Poisson equations then show that |Dv| = O(ρ−1+2γ+4λ ).
Repeating the argument in the previous paragraph we see that |v| = O(ρ2γ+2λ ) and
|χ| + |ψ| = O(ργ+λ ) for any λ < γ. The regularity result in the general case follows.
It suffices to show that (u, v, χ, ψ) is C ∞ in Bδ (0) for some δ > 0 small enough.
div2 (ρ D2 u) − 2 e4u ρ|D2 v − ψ D2 χ + χ D2 ψ|2 − e2u ρ(|D2 χ|2 + |D2 ψ|2 ) = 0, (5.9)
Note that they are well-defined up to a constant. Also, by the C k,α regularity result
from the general case, p, χ̃ and ψ̃ are locally bounded.
Set
1 −4u −1 2
ω̃ = (ω̃ρ , ω̃z ) = (e ρ p + ρ)pz − 4ρ p uz , −(e−4u ρ−1 p2 + ρ)pρ + 4ρ p uρ + 2p
2
1 2
= 2p ωρ + ρ pz − 4ρ p uz , 2p2 ωz − ρ pρ + 4ρ p uρ + 2p .
2
A straightforward calculation using (5.9), (5.13), (5.14) and (5.15) shows that
dṽ = (ω̃ρ + ψ̃ χ̃ρ − χ̃ ψ̃ρ )dρ + (ω̃z + ψ̃ χ̃z − χ̃ ψ̃z )dz. (5.17)
38
Finally, define ũ by
1 1
ũ = − log(e2u ρ2 − e−2u p2 ) = − log(e2ū − e−2ū ρ2 p2 ). (5.18)
2 2
By construction, (ũ, ṽ, χ̃, ψ̃) is smooth in Bδ (0) \ Γ. We claim that it satisfies the
following equations in Bδ (0) \ Γ:
div2 (ρ D2 ũ) − 2 e4ũ ρ |ω̃|2 + e2ũ ρ(|D2 χ̃|2 + |D2 ψ̃|2 ) = 0, (5.23)
Define
e−2u p e−2u p
p̃ = = . (5.27)
e2u ρ2 − e−2u p2 e−2ũ
Then
dp̃ = −2e4ũ ρ ω̃z dρ + 2e4ũ ρ ω̃ρ dz, (5.28)
We next verify (5.23). In the interior of {p̃ = 0}, p and ω̃ vanish, and ũ = −ū, so (5.23)
follows immediately from (5.9), (5.14) and (5.15). Thus, by continuity, it suffices to
consider the region where p̃ 6= 0. We note that
e−2ũ p̃ e−2ũ p̃
p= = .
e−2u e2ũ ρ2 − e−2ũ p̃2
1 −4ũ −1 2
ω= (e ρ p̃ + ρ)p̃z − 4ρ p̃ ũz , −(e−4ũ ρ−1 p̃2 + ρ)p̃ρ + 4ρ p̃ ũρ + 2p̃ ,
2
ωρ,z − ωz,ρ = 4e4ũ ρ p̃ |ω̃|2 − 2p̃ [(ρ ũz )z + (ρ ũρ )ρ ] + 2(e4ũ ρ2 + p̃2 )(ψ̃ρ χ̃z − χ̃ρ ψ̃z ).
= 2e2ũ ρ p̃(|D2 χ̃|2 + |D2 ψ̃|2 ) + 2(e4ũ ρ2 + p̃2 )(ψ̃ρ χ̃z − χ̃ρ ψ̃z ).
4e4ũ ρ p̃ |ω̃|2 − 2p̃ [(ρ ũz )z + (ρ ũρ )ρ ] = 2e2ũ ρ p̃(|D2 χ̃|2 + |D2 ψ̃|2 ).
Lemma 5.2 Assume that f is a regular function in B1 (0) \ Γ and that |f | = O(ρ−1+α )
for some α ∈ (0, 1). If N f be the Newtonian potential of f with respect to B1 (0), i.e.
Z
N f (x) = Φ(x − y)f (y) dy
B1 (0)
40
where Φ is the fundamental solution of the Laplace equation, then N f is C 1,α in B1 (0).
∆u = f
On the other hand, using the regularity result for the general case, we can show
that ũ, ṽ, χ̃ and ψ̃ are C 0,α in Bδ (0) for any α ∈ (0, 1). Moreover,
Applying Lemma 5.2 to (5.19), (5.21), (5.22) and then (5.20), we can show that ũ,
ṽ, χ̃ and ψ̃ are C 1,α . This allows us to bootstrap in between (5.19)-(5.22) to obtain
smoothness for ũ, ṽ, χ̃ and ψ̃ in Bδ (0).
The smoothness of χ and ψ follows from (5.29) and (5.30). By (5.28), p̃ and so
e−2u p = e−2ũ p̃ are smooth. That of e−2u and of ū follows from the identity
Next, since ρ2 p is smooth and p ∈ C 0,α , p is smooth too. The smoothness of v follows
from (5.13). The proof of the theorem is complete.
Proof of Theorem C. The theorem follows immediately from Theorem B and the
results in [27].
41
Appendix A
Hyperbolic spaces
In this appendix, we present some relevant facts about the complex hyperbolic plane
HC ∼
= U (1, 2)/(U (1) × U (2)).
The geodesics at 0 are precisely the R-lines through 0. All other geodesics can be
obtained from those by some motion in U (1, 2). This implies that
|1 − ζ · ζ̂|
cosh dist(ζ, ζ̂) = p q .
1 − |ζ|2 1 − |ζ̂|2
We now derive the (u, v, χ, ψ) parametrization of HC from the disk model. See [25]
for a geometric interpretation.
Define
|1 + ζ1 |
u = log p ,
1 − |ζ|2
1
v = Im w1 ,
2
χ = Re w2
ψ = Im w2 ,
42
where
1 − ζ1
w1 = ,
1 + ζ1
ζ2
w2 = .
1 + ζ1
Note that
1 − |ζ|2
e−2u = = Re w1 − |w2 |2 .
|1 + ζ1 |2
Hence, for a given (u, v, χ, ψ), we can recover ζ by
1 − w1
ζ1 = ,
1 + w1
2w2
ζ2 = ,
1 + w1
and
w1 = e−2u + χ2 + ψ 2 + 2i v,
w2 = χ + i ψ.
Appendix B
1,p
The space WΣ (Ω, w)
We will study the space WΣ1,p (Ω, w) defined in the introduction. Most of the results
here are probably known to readers. However we show them for completeness.
Lemma B.1 Let 1 < p < ∞. Assume (C1)-(C3). Let α and β be functions on (0, ∞)
which are locally bounded measurable functions on compact subsets of (0, ∞) and satisfy
∆ϕ(x)
α(t) ≥ sup 2
,
ϕ(x)=t |Dϕ(x)|
(i) If β w−1/(p−1) e−A/(p−1) is locally integrable near 0, then there exists δ > 0 such
that d(Ωc , Σ) < δ implies
Z Z
p 1,p
|f | w̃−1/(p−1) dx ≤ C |Df |p w dx f ∈ W0,Σ (Ω, w).
Ω Ω
(ii) If, in addition, α is positive and the level surfaces of ϕ are regular hypersurfaces,
then there exists δ > 0 such that d(Ωc , Σ) < δ implies
Z Z
p 1,p
|f | w̃µ dx ≤ C |Df |p w dx f ∈ W0,Σ (Ω, w),
Ω Ω
for any µ for which β w−1/(p−1) eµ A is locally integrable near 0. The constant C
depends only on p and µ.
(i) Let Bt = {d(x, Σ) ≥ t}. For t sufficiently small, f vanishes outside of Bt . Thus,
→
if F is a C 1 vector-valued function on Ω \ Σ then
Z Z
→ → →
p
0= div (f F ) dx = [f p div (F ) + p f p−1 Df · F ] dx.
Bt Bt
→
Hence, if div (F ) is positive,
→
| F |p
Z Z
→
f p div (F ) dx ≤ C p
|∇f | → dx. (B.1)
Bt Bt (div (F ))p−1
Define
Z t −p+1
−A(t) −1/(p−1) −A(s)/(p−1)
F (t) = −e β(s) w (s) e ds < 0,
0
so that
F 0 + α F = (p − 1) β w−1/(p−1) |F |p/(p−1) .
→
Hence, if F (x) = F (ϕ(x))Dϕ(x) then
→ ∆ϕ
div (F ) = |Dϕ|2 F 0 + F ≥ |Dϕ|2 (F 0 + α F )
|Dϕ|2
(ii) We localize the proof of (i). Since α is positive, A is strictly increasing. Let tj be
an increasing sequence of positive real numbers such that 0 ≤ A(tj+1 ) − A(tj ) ≤
1. Let Sj = {d(x, Σ) = tj } and Aj = {tj+1 ≤ d(x, Σ) ≤ tj }. Similar to (i), any
→ →
vector-valued function F which is differentiable in Aj such that div (F ) > 0 gives
rise to
→
| F |p
Z Z
1 p
→
p
−Dj + f div (F ) dx ≤ C |∇f | → dx. (B.2)
2 Aj Aj (div (F ))p−1
where
Z Z
→ →
p
Dj = f F ·νj dσ − f p F ·νj+1 dσ.
Sj Sj+1
→
where Kj is a constant to be specified later. Set Fj (x) = Fj (ϕ(x)) Dϕ(x). Again,
On the other hand, recalling the definition of Fj and the sequence {tj }, we have
(Z )−p
t
|Fj |p/(p−1) (t) = e−p A(t)/(p−1) β(s)w−1/(p−1) (s) e−A(s)/(p−1) ds + Kj
tj
(Z )−p
t
= β(s)w−1/(p−1) (s) e(A(t)−A(s))/(p−1) ds + eA(t)/(p−1) Kj
tj
(Z )−p
t
−1/(p−1) −µ(A(t)−A(s)) A(t)/(p−1)
≥C β(s)w (s) e ds + e Kj
tj
(Z )−p
t
= Cep µ A(t) β(s)w−1/(p−1) (s) eµ A(s) ds + e(µ+1/(p−1))A(t) Kj
tj
(Z )−p
t
p µ A(t) −1/(p−1) µ A(s) (µ+1/(p−1))A(tj )
≥ Ce β(s)w (s) e ds + e Kj .
tj
46
Hence, by setting
Z tj
Kj = e−(µ+1/(p−1))A(tj ) β(s) w−1/(p−1) (s) eµ A(s) ds
0
we arrive at
w−1/(p−1) |Fj |p/(p−1) ≥ C w̃µ (t).
(ii) For ϕ(x) = d(x, Σ), we can take α(t) = k/t and β(t) = 1. Here k is the codimen-
sion of Σ. In that case,
Z t −p
pkµ −1/(p−1) −1/(p−1) kµ
w̃µ (t) = t w (t) w (s) s ds .
0
(iii) The proof is actually valid if we only assume that f ∈ WΣ1,p (Ω, w) and f = 0 at
points on ∂Ω where the normal vector is not perpendicular to Dϕ.
Proposition B.1 Let 1 < p < ∞. Assume (C1)-(C4). Let δ and µ be as in Lemma
B.1. Define
w̃µ (x) if d(x, Σ) < δ,
w̃(x) =
1 otherwise.
1,p
(i) For any f ∈ W0,Σ (Ω, w), f ∈ Lp (Ω, w̃) and
Z Z
|f |p w̃ dx ≤ C [|f |p + |Df |p w] dx.
Ω Ω
47
(ii) For any f ∈ WΣ1,p (Ω, w), f ∈ Lploc (Ω, w̃) and
Z Z
|f |p w̃ dx ≤ C [|f |p + |Df |p w] dx
ω Ω
Proof. Part (i) follows immediately from Lemma B.1 and the classical Poincare in-
equality.
For part (ii), it suffices to consider ω for which there is some r such that each
connected component of ω ∩{d(x, Σ) ≤ r} is bounded by the hypersurface {d(x, Σ) = r}
and two arbitrary hypersurfaces tangential to Dϕ. Also, we can assume that f ∈
Cc∞ (Ω̄ \ Σ). We will show that
Z Z
p
|f | w̃ dx ≤ C [|f |p + |Df |p w] dx (B.3)
ω ω
Let η be a standard cut-off function which vanishes in {d(x, Σ) > r} and is identically
1 in {d(x, Σ) < r/2}. Split f = f1 + f2 = f η + f (1 − η). Then f1 vanishes on
{x ∈ ∂ω|d(x, Σ) = r}. The remainder of ∂ω is tangential to Dϕ. Thus by the remark
following Lemma B.1,
Z Z
p
|f1 | w̃ dx ≤ C |D(f η)|p w dx
ω
Zω Z
p
≤C |f | w dx + C |Df |p w dx
{r/2<d(x,Σ)<r} ω
Z Z
p
≤ C(r) |f | dx + C |Df |p w dx.
ω ω
For f2 , we compute
Z Z Z
p p
|f2 | w̃ dx = |f2 | w̃ dx ≤ C(r) |f |p dx.
ω {r/2<d(x,Σ)<r} ω
Remark B.2 Using the inequality (B.3) instead of Proposition B.1, we can show that
WΣ1,p (Ω, w) embeds compactly into Lp (Ω, w̄) whenever there is some r such that each
component of Ω ∩ {d(x, Σ) ≤ r} is a set bounded by {d(x, Σ) = r} and two hypersurfaces
tangential to Dϕ.
Proposition B.3 Let 1 < p < ∞. Assume (C1)-(C5) and that Ω intersects Σ. Define
w̃ as in Proposition B.1. Let w̄ be a positive measurable weight in Ω such that w̄ =
o(w̃) in a neighborhood of Σ and w̄ is bounded on compact subsets of Ω̄ \ Σ. Assume in
49
addition that w̄ is bounded from below. If w̃ is nowhere locally integrable along Σ, then
for any ω compactly supported in Ω,
Z Z
p
|f | w̄ dx ≤ C |Df |p w dx.
ω Ω
for any ω compactly supported in Ω such that, for some r, ω ∩ {d(x, Σ) ≤ r} is a set
bounded by {d(x, Σ) = r} and two hypersurfaces tangential to Dϕ.
Since w̄ is bounded from, fh is uniformly bounded in Lp (ω), and so in WΣ1,p (ω, w) (as
h → ∞). Hence, by Remark B.2, we can assume that there exists f ∈ WΣ1,p (ω, w) such
that fh converges weakly in WΣ1,p (ω, w) and strongly in Lp (ω, w̄) to f . This results in
Z Z
p
|Df | w dx ≤ lim inf |Dfh |p w dx = 0,
ω h→∞ ω
Finally, we restate the above results for the case w = d(x, Σ)−γ .
1,p
Corollary B.1 (i) For f ∈ W0,Σ (Ω, d−γ ),
Z Z
−γ−p
|f | d p
, dx ≤ C |Df |p d−γ dx.
Ω Ω
where 0 ≤ γ 0 < γ + p. If Ω and ω are fixed concentric balls and Σ is straight, the
constant C depends only on p, γ, γ 0 and the distance from the center to Σ. The
bigger the distance, the bigger the constant.
0
(iii) The embedding WΣ1,p (Ω, d−γ ) ,→ Lploc (Ω, d−γ ) is compact for any 0 ≤ γ 0 < γ + p.
51
References
[2] G. Bunting, Proof of the uniqueness conjecture for black holes, PhD thesis, Univ.
of New England, Armadale N.S.W., 1983.
[8] M. Heusler, Black Hole Uniqueness Theorems, Cambridge Univ. Press, Cam-
bridge, 1996.
[9] D. Kramer and G. Neugebauer, The superposition of two Kerr solutions, Phys.
Lett. A, 75 (1980), pp. 259–261.
[13] S. Majumdar, A class of exact solutions of Einstein’s field equations, Phys. Rev.,
72 (1947), pp. 930–938.
[16] L. Nguyen, On the regularity of singular harmonic maps and axially symmetric
stationary electrovac space-times, preprint.
[17] A. Papapetrou, A static solution of the gravitational field for arbitrary charge
distribution, Proc. R. Irish Acad., A51 (1945), pp. 191–204.
[18] D. Robinson, Uniqueness of the Kerr black hole, Phys. Rev. Lett., 34 (1975),
pp. 905–906.
[19] , A simple proof of the generalization of Israel’s theorem, Gen. Rel. Grav., 8
(1977), pp. 695–698.
[20] P. Smith and E. Stredulinsky, Nonlinear elliptic systems with certain un-
bounded coefficients, Comm. Pure Appl. Math., 37 (1984), pp. 495–510.
[21] R. Wald, General Relativity, The Univ. of Chicago Press, Chicago and London,
1984.
[24] , On the force between rotating co-axial black holes, Trans. Amer. Math. Soc.,
343 (1994), pp. 899–906.
[25] , On the Dirichlet problem for harmonic maps with prescribed singularities,
Duke Math. J., 77 (1995), pp. 135–165.
[27] , N-black hole stationary and axially symmetric solutions of the Ein-
stein/Maxwell equations, Comm. PDE., 21 (1996), pp. 1389–1430.
53
Vita
Luc L. Nguyen