You are on page 1of 43

GLOBAL WELL-POSEDNESS FOR A SYSTEM OF QUASILINEAR

WAVE EQUATIONS ON A PRODUCT SPACE

CÉCILE HUNEAU AND ANNALAURA STINGO


arXiv:2110.13982v1 [math.AP] 26 Oct 2021

Abstract. We consider a system of quasilinear wave equations on the product space R1+3 ×
S1 , which we want to see as a toy model for Einstein equations with additional compact
dimensions. We show global existence for small and regular initial data with polynomial
decay at infinity. The method combines energy estimates on hyperboloids inside the light
cone and weighted energy estimates outside the light cone.

1. Introduction
In the present article we address the problem of global existence of small solutions to a
certain class of quasilinear systems of wave equations on the product space R1+3 × S1 . The
system under consideration has the following form
(
x,y u + u∂y2 u = 1≤i,j≤2 N1 (wi , wj )
P
(1.1) (t, x, y) ∈ R1+3 × S1
x,y v + u∂y2 v = 1≤i,j≤2 N2 (wi , wj )
P

with initial conditions set at time t0 = 2


(1.2) (u, v)(2, x, y) = (φ0 , ψ0 )(x, y), (∂t u, ∂t v)(2, x, y) = (φ1 , ψ1 )(x, y).
In the above system x,y = −∂t2 + ∆x + ∂y2 denotes the D’Alembertian operator in the
(t, x, y) variables where t ∈ R is the time coordinate, x = (x1 , x2 , x3 ) ∈ R3 are the Cartesian
coordinates and y ∈ S1 is the periodic coordinate. The nonlinearities N1 (·, ·), N2 (·, ·) are
linear combination of the following quadratic null forms:
Q0 (φ, ψ) = ∂t φ ∂t ψ − ∇x φ · ∇x ψ
(1.3) Qij (φ, ψ) = ∂i φ ∂j ψ − ∂i φ ∂i ψ 1≤i<j≤3
Q0i (φ, ψ) = ∂t φ ∂i ψ − ∂i φ ∂t ψ 1≤i≤3

where ∂j denotes the derivative ∂xj in the jth direction for j = 1, 3, and w1 , w2 = {u, v}.
The main result we present in this paper asserts the global existence of solutions to (1.1)
when the initial data are small and localized real functions. Our result also extends to
semilinear interactions of the form ∂α wi · ∂y wj with ∂α being any of the derivatives in the
(t, x, y)-variables.

1.1. Motivation and a brief history. Our interest in studying nonlinear wave equations
on product spaces comes from the theory of supergravity (SUGRA) in physics and more
precisely from the Kaluza-Klein theory, which represents the classical approach to the uni-
fication of general relativity with electromagnetism and more generally with gauge fields
1
(see the orignal works of Kaluza [13] and Klein [18]). The Kaluza-Klein approach considers
general relativity in 3 + 1 + d dimensions with space-time factorizing as
M (3+1+d) = R1+3 × K
where K is a compact d-manifold referred to as internal space. In the simplest case d = 1
dimensional gravity is compactified on a circle (K = S1 ) to obtain at low energies a coupled
Einstein-Maxwell-Scalar system in 3+1 space-time dimensions. Kaluza-Klein space-times
R1+3 × S1 have been studied by Witten in his influential work [39], where he proves insta-
bility at the semiclassical level but provides heuristic arguments for classical stability. The
first result proving the classical stability of the Kaluza-Klein theory is obtained by Wyatt
in [40], where only perturbations depending on the non-compact coordinates are considered,
using tools developed by Lindblad and Rodninaski [23]. More general spacetimes with su-
persymmetric compactifications M = R1+n × K have recently been studied by Andersson,
Blue, Wyatt, and Yau in [2]. The space-times M are equipped with the metric
ĝ = ηR1+n + k
where ηR1+n is the Minkowski metric in R1+n and k is such that (K, k) is a compact Ricci-
flat Riemannian manifold having a cover that admits a spin structure and a nonzero parallel
spinor. A global stability result is proved in [2] under the assumption n ≥ 9 and for Cauchy
data that are Schwarzschild near infinity, but it is conjectured that these conditions can
be relaxed and that space-times with a supersymmetric compactification and n = 3 are
nonlinearly stable.
In both the aforementioned works [40] and [2], as well as in many other works concerning
the global stability problem for Einstein equations, the use of the so-called wave-coordinates
allows one to write the Einstein equations as a system of quasilinear wave equations on the
metric g = (g αβ )αβ

(1.4) g αβ ∂α ∂β gµν = Pµν (∂g, ∂g) + Gµν (g)(∂g, ∂g)


where the sum is taken over repeated indices, Pµν are quadratic forms and Gµν (g)(∂g, ∂g)
contain cubic terms. In the present paper we focus on a toy model for the Einstein equations
on R1+3 ×S1 that only keeps a selection of terms from equation (1.4), precisely the semilinear
terms with null structure and the quasilinear term g yy ∂y2 where y is the periodic coordinate
on S1 . Our goal here is in fact to study the global well-posedness for quasilinear wave
equations on the product space R1+3 × S1 without having to make use of the full structure
of the Einstein equations. The unknown u in system (1.1) plays the role of the coefficient
gyy and the unknown v encodes any other metric coefficient gµν .
We mention here that wave equations on product spaces also appear in other contexts,
for instance when studying the propagation of waves along infinite homogeneous waveguides
(see [8, 22, 29, 30]).
One key observation when studying the small data global well-posedness problem for wave
equations on product spaces R1+3 × Td is that they are closely related to infinite systems
coupling wave and Klein-Gordon equations on the flat space R1+3 . In fact, if W = W (t, x, y)
is solution to
x,y W = F (t, x, y) ∈ R1+3 × Td
2
for some source term F , its Fourier modes in the periodic direction {Wk (t, x)}k∈Z solve the
following equations
Z
2 2
(−∂t + ∆x )Wk − |k| Wk = e−iy·k F dy, (t, x) ∈ R1+3 .
Td
In particular, the zero-mode W0 is solution to a wave equation and any other non-zero mode
Wk solves a Klein-Gordon equation of mass |k|, all on the flat space R1+3 .
The global well-posedness for systems coupling a finite number of wave and Klein-Gordon
equations on the flat 3+1 space-time with small data have largely been studied. We cite
the initial results by Georgiev [9] and Katayama [14], followed by LeFloch and Ma [19, 20],
Wang [37,38] and Ionescu and Pausader [11,12] who study such systems as a model for the full
Einstein-Klein-Gordon equations. In [19, 37] global well-posedness is proved for compactly
supported initial data and quadratic quasilinear nonlinearities that satisfy some suitable
conditions, including the null condition of Klainerman [16] for self-interactions between the
wave components of the solution. An idea used in these works is that of employing hyperbolic
coordinates in the forward light cone; this was first introduced by Klainerman [15] for Klein-
Gordon equations and Tataru in the wave context [33], and later reintroduced by LeFloch
and Ma in [19] under the name of hyperboloidal foliation method. In [12] global regularity
and scattering is proved in the case of small smooth initial data that decay at a suitable rate
at infinity and nonlinearities that do not verify the null condition but present a particular
resonant structure. We also cite the work by Dong and Wyatt [6], who prove global well-
posedness for a quadratic semilinear interaction in which there are no derivatives on the
massless wave component. Other related results are [3,5,17,31,34–36] and see [7,10,24–28,32]
for results about wave-Klein-Gordon systems in lower dimensions.
Our goal in this paper is to prove the global stability for (1.1) in the case where the initial
data are not compactly supported but only have a mild polynomial decay at infinity. Our
approach makes use of the vector field method by Klainerman [4] and follows LeFloch and
Ma [20] in that a big portion of the estimates recovered for the solution in the interior of the
cone {t = r + 1} × S1 are estimates on hyperboloids. The main difference with [19] is that
the interior estimates need to be coupled with exterior estimates in the region outside the
cone. Those are weighted energy estimates that we have to propagate in time.

1.2. The main result. In order to describe the initial data we consider for our problem we
introduce the energy space H 0 endowed with the norm
k(u[t], v[t])k2H 0 := kuk2Hxy 2 2 2
1 + kut kL2 + kvkH 1 + kvt kL2
xy

and the higher order energy spaces H n for n ≥ 1 endowed with the norm
(∂ α u[t], ∂ α v[t]) 2 0
X
k(u[t], v[t])k2H n :=

xy xy H
|α|≤n

where we use the following notation for the Cauchy data in (1.2) at time t:
(u[t], v[t]) := (u(t), ut (t), v(t), vt (t)).
The global well-posedness result that is the object of this paper is proved under some decay
assumptions on the initial data. A premilinary version of our main theorem states the
following
3
Theorem 1. Assume the initial data (u[2], v[2]) for (1.1) satisfy
5
X ≤k+ α k
hxi 2∂ (u[2], v[2]) 0 ≤ ǫ ≪ 1,
xy H
k=0
p
for some positive fixed α and hxi = 1 + |x|2 . Then the system (1.1) is globally well-posed
in the space H 5 .
We remark here that our choice to set the initial data at time t0 = 2 over the conventional
t0 = 0 is more convenient for our computations and comes at no expense as the system (1.1)
is invariant under time translations.
1.3. The wave-Klein-Gordon structure. The Cauchy problem (1.1)-(1.2) can be written
in a more compact form as a vector equation for the unknown W = (u, v)T
(1.5) x,y W + u∂y2 W = N(W, W )
with data
(1.6) W |t=2 = Φ0 , ∂t W |t=2 = Φ1
where Φ0 = (φ0 , ψ0 )T and Φ1 = (φ1 , ψ1 )T and
X N (w , w )
1 i j
N(W, W ) = .
N2 (wi , wj )
1≤i,j≤2

The projection of W onto the periodic direction y reveals the nature of the equation (1.5)
as a system coupling one (vector) wave equation with an infinite sequence of (vector) Klein-
Gordon equations of mass |k| with k ∈ Z∗ . If we denote by Wk = (uk , vk )T the projection of
W onto the k-th frequency
Z
Wk (t, x) = e−iky W (t, x, y) dy, k∈Z
S1

we see that the functions {Wk }k satisfy the following coupled system
(
(−∂t2 + ∆x )Wk − |k|2 (1 + u0 )Wk = S1 e−iky N(W, W ) dy − S1 e−iky (u − u0 )∂y2 W dy
R R

k ∈ Z.
The zero mode W0 is solution to a wave equation while any other non-zero mode Wk is
solution to a nonlinear Klein-Gordon equation of mass |k|. For our analysis this distinc-
tion will be fundamental and we will often work throughout the paper with the following
decomposition of W
X
(1.7) W = W0 + W, W(t, x, y) = eiky Wk (t, x),
k6=0

so that the equation (1.5) is equivalent to the following system


(
(−∂t2 + ∆x )W0 = S1 N(W, W ) dy + S1 ∂y u ∂y W dy
R R
(1.8)
(−∂t2 + ∆x )W + (1 + u) ∂y2 W = N(W, W ) − S1 N(W, W ) dy − S1 ∂y u ∂y W dy.
R R

We are now able to state a more precise version of the main theorem.
4
Theorem 2. Assume that for some positive fixed α the initial data for (1.5) satisfy
5
X ≤k+ α k
x 2∂
xy W [2]
≤ ǫ ≪ 1.
H0
k=0
5
Then the solution W
R to (1.5)-(1.6) exists globally in time in H and the two components of
the solution W0 = S1 W dy and W = W − W0 satisfy the following pointwise bounds
j 1
|∂tx W0 (t, x)| . ǫht + |x|i−1 ht − |x|i− 2 , j = 1, 3
∂y ∂tx W(t, x, ·) 2 1 . ǫht + |x|i− 23 , j = 0, 1, k = 0, 1,
j k
Ly (S )

∂ ∂ W(t, x, ·) 2 1 . ǫht + |x|i−1 ht − |x|i− 21 , j = 0, 1.


j 2
y tx L (S )
y

1.4. Vector Fields. In order to describe the global bounds and decay properties of the
solution W = (u, v)T to (1.5)-(1.6) we need to introduce the family of Killing vector fields
associated to our problem. Those are the vector fields that exactly commute with x,y :
(1.9) ∂t , ∂1 , ∂2 , ∂3 , ∂y
(1.10) Ωij = xj ∂i − xi ∂j 1≤i<j≤3
(1.11) Ω0i = t∂i + xi ∂t i = 1, 3.
The expressions in (1.9) correspond to the translations in the coordinate directions; (1.10)
correspond to the Euclidean rotations in the x coordinates; (1.11) are the hyperbolic rota-
tions, also called boosts. We also introduce the conformal scaling vector field
(1.12) S = t∂t + x · ∇x
which is not Killing for (1.5) but will appear later in the analysis of the problem. We refer
to (1.10) and (1.11) as Klainerman vector fields and generally denote them by Z:
Z := {Ωij , Ω0i }.
We denote the full set of admissible vector fields for x,y as
(1.13) Z := {∂t , ∂1 , ∂2 , ∂3 , ∂y , Ωij , Ω0i }
and for any multi-index γ = (α, β) we denote
Z γ = ∂αZ β .
For any two non-negative integers k, n with k ≤ n we say that the multi-index γ = (α, β)
is of type (n, k) if |γ| = |α| + |β| ≤ n and |β| ≤ k, in other words if there are at most k
Klainerman vector fields among the |γ| admissible vector fields in the product Z γ .
1.5. The null structure. The nonlinearities we consider in this work are linear combination
of the classical quadratic null forms (1.3). An important feature of the null forms is that they
are combination of three types of products and can be expressed schematically as follows
t−r
(1.14) N(φ, ψ) = ∂φ · ∂ψ + ∂φ · ∂ψ + ∂φ · ∂ψ
t
where ∂ j = t−1 Ω0j are the rescaled hyperbolic rotations, or also as
(1.15) N(φ, ψ) = T φ · ∂ψ + ∂φ · T ψ
5
x
where Tj = ∂j + rj ∂t for j = 1, 3 are the vector fields tangent to the cones {t − r = const}.
The T vector fields are related to the boosts in general via the following relation
1 (t − r) xj
Tj = Ω0j + ∂t .
t t r
We will often make use of the two representations (1.14) and (1.15) to recover suitable
pointwise and energy estimates for the solution as they allow to recover additional decay for
the interactions W0 × W0 .

1.6. The interior and exterior region. The proof of our main theorem is based on the
combination of a classical local existence result with a bootstrap argument. We will perform
such argument separately in the two regions in which we decompose our space-time:
interior region D in := {(t, x) : t ≥ 2 and |x| < t − 1} × S1

exterior region D ex := {(t, x) : t ≥ 2 and |x| ≥ t − 1} × S1 .


In order to describe our bootstrap assumptions we first introduce some notations. Given any
hyperboloid Hs in R1+3 × S1 we denote by Hsin (resp. Hsex ) the branch of Hs contained in
the interior region D in (resp. in the exterior region D ex ):
Hs = (t, x) : s2 = t2 − |x|2 × S1


Hsin := (t, x, y) ∈ Hs : |x| < (s2 − 1)/2 ,




Hsex := (t, x, y) ∈ Hs : |x| ≥ (s2 − 1)/2 .




in
Moreover we denote by H[2,s] the hyperbolic interior region above H2in and below Hsin , and
ex
by H[2,s] the portion of the exterior region below Hsex , for any s ≥ 2 (see figure 1)
in
:= (t, x, y) ∈ D in : 2 ≤ t2 − |x|2 ≤ s2

H[2,s]
ex
:= (t, x, y) ∈ D ex : t2 − |x|2 ≤ s2 .

H[2,s]
In the interior region the bootstrap assumptions will be energy bounds on the truncated
hyperboloids Hsin for s ≥ 2 and pointwise bounds on the Z derivative of the zero mode
of the solution. The local wellposedness theory for this problem assures the existence and
smallness of the solution W = (u, v)T to (1.5)-(1.6) up to the interior hyperboloid H2in ,
hence our goal will be to propagate the bootstrap assumptions in the hyperbolic interior
region above H2in
in
:= (t, x, y) ∈ D in : 2 ≤ t2 − |x|2 .

H[2,∞)
In the exterior region the bootstrap assumptions will instead be weighted energy bounds
on the constant time slices Σex
t which foliate D
ex
for t ≥ 2,
Σex 3 1
t := {x ∈ R : |x| ≥ t − 1} × S .

We warn the reader that throughout the paper we will work with functions defined on
the product space R1+3 × S1 as well as with functions not depending on the y variable and
defined on the flat space R1+3 . With the purpose of keeping notations as light as possible, a
region in R1+3 × S1 and its projection onto R1+3 will have the same name.
6
t t=r+1 t=r

Hsin

in
H[2,s]

in
Figure 1. Vertical section of the region H[2,s] projected onto R1+3

1.7. The energy functionals. In the interior region we aim to propagate a-priori energy
bounds on the solution on the truncated hyperboloids Hsin . The interior energy functional
on Hsin associated to the linear counterpart of equation (1.5) is
ZZ  s 2
in
E (s, W ) := |∂t W |2 + |∂W |2 + |∂y W |2dxdy
H in t
(1.16) Z Z s  2
s X
= |∂x W |2 + |S W |2 + t−2 |Ωij W |2 + |∂y W |2 dxdy
Hsin t 1≤i<j≤3

where |∂W |2 stands for 3i=1 |∂ i W |2 . The above functional can be decomposed as follows
P

E in (t, W ) = E in (t, W0 ) + E in (t, W)


where E in (t, W) is defined by replacing W with W in (1.16) and
Z  s 2
in
E (s, W0 ) := |∂t W0 |2 + |∂W0 |2 dx
H in t
Z s  2
s X
= |∂x W0 |2 + |S W0 |2 + t−2 |Ωij W0 |2 dx.
Hsin t 1≤i<j≤3

We also introduce and work with the conformal energy functional on truncated hyperboloids
associated to the linear flat wave equation on R1+3 . This is the functional defined as follows
3
1 s2 X
Z
c,in 2
(1.17) E (s, W0 ) := 2
|KW0 + 2tW0 | + 2 |Ω0i W0 |2 dx,
Hsin t t i=0

where K = (t2 + r 2 )∂t + 2rt∂r is the Morawetz multiplier.


In the exterior region we will describe the evolution of equation (1.5) by means of the
following weighted energy functional
ZZ
ex,α
(2 + r − t)α+1 |∂t W |2 + |∇x W |2 + |∂y W |2 dxdy
 
(1.18) E (t, W ) =
Σex
t
7
and of stronger norm XTex,α
(1.19)
Z T ZZ
kW k2X ex,α ex,α
(2 + r − t)α |T W |2 + |∂y W |2 dxdydt

:= sup E (t, W ) + (1 + α)
T
t∈[2,T ] 2 Σex
t

where |T W |2 stands for 3i=1 |Ti W |2 . In the above energy functionals we have T > 2 and
P
α > 0. The bootstrap assumptions in this region will be energy bounds on the XTex,α norm
of the solution, which not only controls the weighted energy of the solution but also the
weighted L2 space-time norm of the good derivatives: the tangential derivatives T to the
cones {t − r = const} and the derivative along the periodic direction ∂y .
As a result of the global energy bounds that will be proved to hold in the exterior region
(see Section 4) we also obtain a control on the energy of W on the exterior hyperboloids
ZZ  s 2
ex,h
E (s, W ) = |∂t W |2 + |∂W |2 + |∂y W |2dxdy
H ex t
(1.20) Z Z s  2
s X
= |∂x W |2 + |S W |2 + t−2 |Ωij W |2 + |∂y W |2 dxdy
Hsex t 1≤i<j≤3

as well as on the exterior conformal energy of W0 on constant time slices Σex


t
Z 3
X
(1.21) E c,ex (t, W0 ) := |S W0 + 2W0 |2 + |Ω0i W0 |2 dx.
Σex
t i=0

1.8. The quasilinear energies. Equation (1.5) is quasilinear and in order to propagate
both the interior and exterior a-priori energy bounds for the solution we need to consider a
cubic modification of the energies introduced in the previous subsection. Such quasilinear
energies are defined as follows
ZZ
in in
Equasi (s, W ) := E (s, W ) + u|∂y W |2 dxdy
Hsin
ZZ
ex,α ex,α
Equasi (t, W ) := E (t, W ) + u|∂y W |2 dxdy
Σex
t
ZZ
ex,h ex,h
Equasi (s, W ) := E (s, W ) + u|∂y W |2 dxdy.
Hsex

We also introduce the quasilinear modification of the stronger norm XTex,α


Z T ZZ
ex,α
kW k2X ex,α (2 + r − t)α |T W |2 + (1 + u)|∂y W |2 dxdydt.

:= sup Equasi (t, W ) + (1 + α)
quasi,T
t∈[2,T ] 2 Σex
t

We immediately observe that under smallness assumptions on u, e.g. |u| ≤ 1/10, our starting
energies are equivalent to their corresponding quasilinear counterparts, i.e. for any E =
{E in , E ex,α , E ex,h }
9 11
E(t, W ) ≤ Equasi (t, W ) ≤ E(t, W ).
10 10
The same holds true for the stronger norms XTex,α and Xquasi,T
ex,α
.
8
1.9. Higher order norms. We use the vector fields introduced before to define the higher
order counterparts of the energy functionals and of the stronger exterior norm XTex,α
X
En (s, W ) := E(s, Z γ W ), E = {E in, E ex,α , E ex,h }
|γ|≤n

X
kW kXTn,α := kZ γ W kXTex,α .
|γ|≤n

The higher order energies of W0 and W are defined analogously. We observe that the above
higher order energies control the high Sobolev regularity of the solution in the interior and
exterior region respectively and also keep track of the Z vector fields applied to the solution
in addition to usual derivatives. In the interior region it will be important to keep track of
the precise number of Klainerman vector fields acting on the W component of the solution
and to that purpose we also introduce the following energy
X
in
En,k (s, W) := E(s, Z γ W),
In,k

where In,k denotes the set of indices of type (n, k). We finally introduce the higher order
counterparts of the conformal energy functionals (1.17) and (1.21) in order to control the
conformal energies of pure products of Klainerman vector fields acting on W0
X X
Enc,in (s, W0 ) := E c,in (s, Z β W0 ), Enc,ex (t, W0 ) := E c,ex (s, Z β W0 ).
|β|≤n |β|≤n

1.10. Notations. Throughout the rest of the paper we will use the following notations:
• r = |x|
• ∇ = (∂1 , ∂2 , ∂3 ) denotes spatial derivatives
• ∂ = {∂t , ∂j , ∂y } denotes space-time derivatives
• ∂tx = {∂t , ∂j } denotes flat space-time derivatives
x
• ∂ j = t−1P
Ω0j and Tj = ∂j + rj ∂t for j = 1, 3
• Z ≤n = |γ|≤n Z γ

Acknowledgments. This material is based upon work supported by the National Science
Foundation under Grant No. DMS-1929284 while the second author was in residence at
the Institute for Computational and Experimental Research in Mathematics in Providence,
RI, during the Hamiltonian Methods in Dispersive and Wave Evolution Equations program.
The second author was also supported by an AMS Simons travel grant. The first author was
supported by the ANR-19-CE40-0004.

2. Overview of the proof


The proof of our main theorem is based on the combination of a classical local well-
posedness result for equation (1.5) with a bootstrap argument. We will perform this argu-
ment separately in the interior region D in and the exterior region D ex in which we divide
the space-time R1+3 × S1 .
9
The bootstrap assumptions in the exterior region D ex are uniform-in-time energy bounds
on the higher order stronger norm XT5,α
0
of the solution W , for any fixed α > 0:
(2.1) kW k2X 5,α ≤ 2C02 ǫ2 .
T0

Such bound in particular implies the existence of some function l ∈ L1 ([2, T0 ]) such that
ZZ
(2 + r − t)α |T Z ≤5 W |2 + |∂y Z ≤5 W |2 dxdy ≤ 2C02 ǫ2 l(t).

(2.2)
Σex
t

The result we want to prove in the exterior region is the following


Proposition 2.1. There exists a constant C0 > 0 sufficiently large and a constant ǫ0 > 0
sufficiently small such that, for every 0 < ǫ < ǫ0 if W = (u, v)T is a solution to (1.5)-(1.6)
in an interval [2, T0 ] and satisfies the energy bounds (2.1) then actually
kW k2X 5,α ≤ C02 ǫ2 .
T0

In the above proposition the time T0 is arbitrary, therefore the solution W exists globally
in D ex and satisfies the energy bound (2.1) for all times T0 > 2. In particular, one has that
(2.3) kZ ≤5 W kX∞
ex,α := lim kZ
≤5
W kXTex,α ≤ 2C02 ǫ2 .
T0 →∞ 0

The bootstrap assumptions in the interior region are higher order energy bounds on hy-
perboloids for W0 and W and pointwise bounds on the Z derivative of W0
(2.4) E5in (s, W0 ) ≤ 2A2 ǫ2
in
(2.5) E5,k (s, W) ≤ 2A2 ǫ2 s2δk , k = 0, 5

(2.6) |ZW0 (t, x)| ≤ 2Bǫt−1 sσ


in
for all s ∈ [2, s0 ], (t, x) ∈ H[2,s 0]
, and s0 > 2 fixed. Here δ0 = 0, the parameters σ, δk s are
fixed small universal constants satisfying 0 < σ ≪ δk ≪ δk+1 for k = 1, 4, and A and B are
large universal constants which we will improve as a part of the conclusion of the proof. The
result we want to prove in this region requires the global exterior energy bounds (2.3) and
can be stated as follows
Proposition 2.2. There exist two constants A, B > 0 sufficiently large, 0 < ǫ0 , σ, δk ≪ 1
sufficiently small with δ0 = 0 and σ ≪ δk ≪ δk+1 for k = 1, 4 such that, for every 0 < ǫ < ǫ0
if W = (u, v)T is a solution to (1.5)-(1.6) in the region H[2,s
in
0]
∪ D ex and satisfies the global
exterior energy bounds (2.3) as well as the interior bounds (2.4)- (2.6) for all s ∈ [2, s0 ],
then it actually satisfies the enhanced interior bounds
E5in (s, W0 ) ≤ A2 ǫ2
in
E5,k (s, W) ≤ A2 ǫ2 s2δk , k = 0, 5
|ZW0 (t, x)| ≤ Bǫt−1 sσ
in
for all s ∈ [2, s0 ] and all (t, x) ∈ H[2,s 0]
.
10
In the above proposition the hyperbolic time s0 is arbitrary which implies the global
existence of the solution in the interior region D in . The reason why we distinguish between
the energies of W0 and of W and do not propagate uniform-in-time energy bounds for the
latter is related to some slow decaying semilinear terms that only appear in the equations
satisfied by the differentiated function Z γ W. One of such terms is for instance Z β u0 · ∂y2 W,
which appears when γ is a multi-index of type (k, k), i.e. when Z γ = Z β is a pure product
of Klainerman vector fields. The L2xy norm of such product can only be controlled in terms
of the (square root of the) conformal energy Ekc,in (s, W0 ), but even assuming the following
sharp bounds
3
Ekc,in (s, W0 ) . ǫ2 s and sup |∂y2 W| . ǫ2 s− 2
Hsin

we are not able to recover a better L2 bound than the following one, which is at the limit
of integrability and prevents us from obtaining uniform in time bounds for the higher order
energies of W
1
≤ Ekc,in (s, W0 ) 2 k∂y2 WkL∞ (Hsin ) . ǫ2 s−1 .
β
Z u0 · ∂ 2 W 2

y L (H in )
xy s

These problematic terms do not appear in the equation for Z γ W0 , for which we can instead
easily prove uniform in time energy bounds using the null structure.
In both the interior and exterior region the booststrap argument is performed in two
classical steps. The first step consists in recovering pointwise bounds for the solution. A
first set of estimates is obtained from the a-priori energy bounds using Klainerman-Sobolev
inequalities on hyperboloids in the interior region and weighted Sobolev embeddings on
constant time slices in the exterior region. The pointwise bounds obtained for W in the
interior region with this approach are not optimal due to the slow growth in time assumed
in (2.5), therefore more suitable bounds need to be recovered by directly analyzing the
equation satisfied by W. Since Klainerman-Sobolev inequalities only yield pointwise bounds
for derivatives of W0 , we will also need to recover L∞ − L∞ estimates for W0 using the
equation it satisfies.
The second step of the bootstrap argument consists in writing a higher order (interior and
exterior respectively) energy inequality for the solution and in using the pointwise bounds
previously obtained to perturbatively estimate the cubic terms appearing in the right hand
side of such inequality. As appears from the example briefly mentioned above, we also need
to recover higher order conformal energy bounds for W0 from the a-priori energy bounds.
This is especially important in order to propagate the interior energy bounds (2.4) and (2.5).
The paper is structured as follows. In Section 3 we derive the energy inequalities for the
linearized equation, both in the exterior and the interior region. We also derive the conformal
energy inequalities. In Section 4 we prove the global existence of the solution in the exterior
region. In Section 5 we recover the pointwise estimates for W and W0 . Finally in Section 6
we improve the energy estimates in the interior region, concluding the proof of Theorem 1.

3. The Linearized Equation


The purpose of this section is to write the energy inequality and the conformal energy
inequality for the linearized equation associated to the equation (1.5) in both the interior
11
and exterior regions D in and D ex . We will look at the following linear inhomogeneous
equation

(3.1) (−∂t2 + ∆x )W + (1 + u)∂y2 W = F, (t, x, y) ∈ R × R3 × S1 ,

where u is assumed to be a sufficiently small function, e.g. |u| ≤ 1/10.


We start by proving a weighted energy inequality on the exterior constant time slices Σex t
for general linear inhomogeneous equations of the form (3.1). In the following proposition
the lifespan T0 is arbitrary, DTex0 denotes the portion of exterior region in the time strip [2, T0 ]
and C[2,T0 ] is its null boundary

DTex0 = {(t, x, y) ∈ D ex : 2 ≤ t ≤ T0 } ,

C[2,T0 ] = {(t, x) : 2 ≤ t ≤ T0 , r = t − 1} × S1 .

Proposition 3.1. Let W be a solution to the equation (3.1), l ∈ L1 ([2, T0 ]) and suppose
that u is a function satisfying the following pointwise bounds in the exterior region DTex0

(3.2) kukL∞ (DTex ) ≤ 1/10


0
1
p
(3.3) (2 + r − t) 2 ∂y u(t, x, ·) ∞ 1 . ǫ l(t)

L (S )
(3.4) k(2 + r − t)∂u(t, x, ·)kL∞ (S1 ) . ǫ.

For any fixed α > 0 the following inequality holds true


ZZ
kWk2X ex,α (2 + r − t)α+1 |T W|2 + |∂y W|2 dσdydt

(3.5) +
T 0 C[2,T0 ]
α+1

. E ex,α (2, W) + (2 + r − t) 2 F kWkXTex,α .

L1t L2xy (DTex ) 0
0

where dσ is the area element of the sphere S2 .

We remark that the result of proposition 3.1 can be proved for any positive and increasing
weight ω = ω(r − t) only depending on the distance from the cone {t = r} if the hypothesis
on the function u are changed appropriately.

Proof. From the smallness assumption on u it will be enough to prove the inequality in
ex,α
the statement with E ex,α (t, W) and kWkXTex,α replaced by Equasi (t, W) and kWkXquasi,T
ex,α
0 0
respectively. A simple computation shows that for any positive weight ω = ω(r − t) one has
(3.6)
1 
ω(r − t)∂t W x,y W + u ∂y2 W = − ∂t ω(r − t) (∂t W)2 + |∇x W|2 + (1 + u)(∂y W)2
  
2
+ divx (ω(r − t)∂t W ∇x W) + ∂y (ω(r − t)(1 + u)∂t W∂y W)
 
1 ′ 2 2
 1 2
− ω (r − t) |T W| + (1 + u)(∂y W) − ω(r − t) ∂y u ∂t W∂y W − ∂t u(∂y W) .
2 2
12
We consider the case ω(z) = (2 + z)α+1 , integrate the above equality over the exterior region
DTex0 and use the Stokes’ theorem. We obtain that
(3.7) ZZ
2 ex,α
(2 + r − t)α+1 |T W|2 + (1 + u)|∂y W|2 dσdydt ≤ Equasi
 
kWkX ex,α + (2, W)
quasi,T0
C[2,T0 ]
 
1
ZZ
α+1 2
+2 (2 + r − t) F ∂t W + ∂y u ∂t W∂y W − ∂t u (∂y W) dxdydt.
DTex 2
0

The smallness of u assures that the integral in the above left hand side is equivalent to the one
in the left hand side of inequality (3.5). From the assumption (3.3) and the Cauchy-Schwarz
inequality we see that
ZZ
α+1
α+1
(2 + r − t) |∂y u∂t W∂y W|dxdydt . (2 + r − t) ∂y u · ∂y W 1 kWkXTex,α
2

ex Lt L2xy (DT ) 0
DTex 0
0
1

(2 + r − t) α2 ∂y W 2

. (2 + r − t) 2 ∂y u kWkXTex,α

L ex
L2t L∞ ex txy (DT0 )
xy (DT )
0
0

. ǫk lkL2t kWk2X ex,α . ǫkWk2X ex,α
T0 T0

and from (3.4)


ZZ ZZ
α+1 2
(2 + r − t) |∂t u(∂y W) |dxdydt . ǫ (2 + r − t)α (∂y W)2dxdydt . ǫkWk2X ex,α .
T0
DTex DTex
0 0

If ǫ ≪ 1 is sufficiently small the above two integrals can be hence absorbed in the left hand
side of (3.7). Finally, from the Cauchy-Schwarz inequality we also have that
ZZ
α+1
α+1
(2 + r − t) |F||∂t W|dxdydt ≤ (2 + r − t) F 1 kWkXTex,α .
2

ex Lt L2xy (DT ) 0
DTex 0
0


If we assume that the function u satisfies the pointwise bounds (3.2)-(3.4) in the whole
exterior region D ex we can also recover an energy inequality on the exterior truncated hy-
perboloids Hsex . In the following proposition C[2,s] denotes the lateral boundary of H[2,s]
ex

C[2,s] = (t, x) : 2 ≤ t ≤ (s2 + 1)/2, r = t − 1 × S1




and k · kX∞
ex,α is the stronger norm over the time interval [2, ∞)

ex,α := lim kWk ex,α .


kWkX∞ XT
T →∞

Proposition 3.2. Assume that W is solution to the linear inhomogeneous equation (3.1)
with u satisfying the decay properties (3.2)-(3.4) in the whole exterior region D ex . Then
ZZ
ex,h
E (s, W) + |T W|2 + |∂y W|2 dσdydt
C[2,s]

. E ex,0 (2, W) + ǫkWk2X ex,0 + kFkL1t L2xy (H[2,s]


ex ) kWk ex,0
X∞

for any s ≥ 2.
13
Proof. From the smallness assumption on u it will be enough to prove the the statement
ex,h ex,0
with E ex,h (t, W) and E ex,0 (2, W) replaced by Equasi (t, W) and Equasi (2, W) respectively.
ex
We integrate the relation (3.6) in the case where ω ≡ 1 over the region H[2,s] which we
s
foliate by the constant time slices Σt for t ≥ 2 (see picture 2), where
n √ o
s 3
Σt = x ∈ R : r ≥ max(t − 1, t − s ) × S1 .
2 2

We obtain that

Hsex
t

Hsin ex
H[2,s]

Σex
2

ex
Figure 2. Vertical section of the region H[2,s] and its foliation projected onto R1+3

ZZ
ex,h ex,0
Equasi (s, W) + |T W|2 + (1 + u)|∂y W|2 dσdydt ≤ Equasi (2, W)
C[2,s]
1
ZZ ZZ
+ ∂y u ∂t W∂y W − ∂t u (∂y W)2 dxdydt + F∂t W dxdydt.
ex
H[2,s] 2 ex
H[2,s]

From the assumptions on the derivatives of u, the Cauchy-Schwarz inequality and the defi-
ex,0
nition of the norm X∞ we immediately see that
ZZ Z ∞
∂u ∂y W ∂W dxdydt ≤ k∂ukL∞ s k∂y WkL2 (Σs ) k∂WkL2 (Σs ) dt
xy (Σt ) xy t xy t
ex
H[2,s] 2
Z ∞ p
. kWkX∞
ex,0 ǫ l(t)k∂y WkL2xy (Σst ) dt . ǫkWk2X ex,0 .

2

Furthermore ZZ
F∂t W dxdydt ≤ kFkL1t L2xy (H[2,s]
ex ) kWk ex,0 .
X∞
ex
H[2,s]


14
We now prove the interior energy inequality for (3.1). In the following proposition the
hyperbolic lifespan s0 is arbitrary and C[2,s] will denote the later boundary of the hyperbolic
in
region H[2,s]
C[2,s] = (t, x) : t = r + 1 and 3/2 ≤ t ≤ (s2 − 1)/2 × S1 .


Proposition 3.3. Let W be a solution to the equation (3.1) and suppose that u is a function
in
that satisfies the following bounds in the hyperbolic region H[2,s 0]

(3.8) kukL∞ (H[2,s


in ) ≤ 1/10
0]
1
(3.9) |∂t u0 (t, x)| . ǫt− 2 s−1
3
(3.10) k∂u(t, x, ·)kL∞ (S1 ) . ǫt− 2 +δ
R
for some small δ > 0, where u0 = S1 u(t, x, y)dy and u = u − u0. Then
(3.11) ZZ Z s
1
E in (s, W) . E in (2, W) + |T W|2 + |∂y W|2 dσdydt + kFkL2xy (Hτin ) E in (τ, W) 2 dτ
C[2,s] 2

for all s ∈ [2, s0 ], where dσ is the surface element on the sphere S2 .


in
Proof. We integrate the equality (3.6) in the case where ω ≡ 1 over the region H[2,s] (see
picture 1) for any s ∈ [2, s0 ] and use the Stokes’ theorem together with system (3.1). We
in
foliate H[2,s] by hyperboloids Hτin with τ ∈ [2, s0 ] and obtain the following
(3.12) ZZ
in in
Equasi (s, W) ≤ Equasi (2, W) + |T W|2 + (1 + u)|∂y W|2 dσdydt
C[2,s]
Z sZ Z sZ
1
+ (τ /t)∂y u ∂t W∂y W − (τ /t)∂t u (∂y W)2 dxdydτ + (τ /t)F∂t W dxdydτ.
2 Hτin 2 2 Hτin

The integral on the null boundary C[2,s] in the above right hand side is bounded by its
counterpart in the right hand side of (3.11) thanks to the smallness assumption (3.8). From
the assumption (3.10) and the fact that τ ≤ t we derive that
Z sZ
1 2

(τ /t)∂y u ∂t W∂y W − (τ /t)∂t u (∂y W) dxdydτ

2 Hτin 2
Z s Z s
3
in
. k∂ukL∞ (Hτin ) E (τ, W) dτ . ǫ τ − 2 +δ E in (τ, W) dτ
2 2

while from (3.9) we have


Z sZ Z s
1 2 3
τ − 2 E in (τ, W) dτ.

(τ /t)∂t u0 (∂y W) dxdydτ . ǫ

2 Hτin 2

2

Finally, the Cauchy-Schwarz inequality yields


Z sZ Z s0
1
kFkL2xy (Hτin ) E in (τ, W) 2 dτ.


(τ /t)F∂t W dxdydτ ≤
in2 Hτ 2


15
As detailed in Section 6, it will be important to distinguish between the two components
W0 and W of the solution W to (1.5), in particular to show that the energies associated
to the zero mode W0 are uniformly bounded in time. We will make use of the following
classical result on the energy on interior truncated hyperboloids for linear inhomogeneous
wave equations on the flat space R1+3 .
Proposition 3.4. Let W0 be a solution to the following linear inhomogeneous wave equation
(3.13) (−∂t2 + ∆x )W0 = F0 , (t, x) ∈ R × R3 .
For all s ∈ [2, s0 ] we have the following energy inequality
Z Z s
2
in in
E (s, W0 ) ≤ E (2, W0 ) + |T W0 | dσdt + kF0 kL2x (Hτin ) E in (τ, W0 )1/2 dτ.
C[2,s] 2

We also state below the interior and exterior conformal energy inequalities for linear
inhomogeneous wave equations on R1+3 . We will need to have a control on the higher order
conformal energies of the zero-mode W0 of our solution W in order to recover pointwise
bounds for W0 and ZW0 later in the paper.
Proposition 3.5. Let W0 be a solution to (3.13). We have the following inequality
Z s
1
c,in c,in
E (s, W0 ) ≤ E (2, W0) + ksF0 kL2 (Hτin ) E c,in (τ, W0 ) 2 dτ
2
Z
+ |(t + r)(∂t W0 + ∂r W0 ) + 2W0 |2 + (t − r)2 (|∇W0 |2 − (∂r W0 )2 ) dσdt,
C[2,s]

where dσ is the surface element on the sphere S2 .


Proof. Let K = (t2 + r 2 )∂t + 2rt∂r denote the Morawetz multiplier. Then
(3.14)
(KW0 + 2tW0 ) x W0
 
1 2 2 2 2
 2
= −∂t (t + r ) |∂t W0 | + |∇x W0 | + 2rt∂t W0 ∂r W0 + 2tW0 ∂t W0 − W0
2
+ divx (t + r 2 )∂t W0 ∇x W0 + 2rt∂r W0 ∇x W0 + tx |∂t W0 |2 − |∇x W0 |2 ) + 2tW0 ∇x W0
 2 

and the result of the statement follows from the integration of the above equality over the
in
interior hyperbolic region H[2,s] and from the Stokes’ theorem. 
Proposition 3.6. Let W0 be a solution to the equation (3.13). Then
Z
c,ex
E (T, W0 ) + |(t + r)(∂t W0 + ∂r W0 ) + 2W0 |2 + (t − r)2 (|∇W0 |2 − (∂r W0 )2 ) dσdt
C[2,T ]
Z T
1
c,ex
≤E (2, W0 ) + k(t + r)F0 kL2x (Σex
t )
E c,ex (t, W0 ) 2 dt,
2
where dσ is the surface element on the sphere S2 .
Proof. The result of the proposition follows from the integration of equality (3.14) over the
region DTex combined with Stokes’ theorem and the following equality
KW0 + 2tW0 = t (S W0 + 2W0 ) + rΩ0r W0 .

16
In the interior region we will recover pointwise bounds on ZW0 from the higher conformal
energies of W0 via Klainerman-Sobolev inequalities on hyperboloids (see lemma 5.1). This
will require a control for the conformal energy of the solution on a portion of the hyperboloid
Hs in the exterior region D ex , that in turn will be obtained from a control of the conformal
energy on the flat hypersurfaces Σst defined below. We hence state the following modification
of the conformal energy inequality.

Proposition 3.7. For any s ≥ 2, s < T1 < T2 and any t ∈ [T1 , T2 ] let

Σst := {x ∈ R3 : |x| ≥ t2 − s2 }

and
Z 3
X
E c,s
(t, W0 ) := |S W0 + 2W0 |2 + |Ω0i W0 |2 dx.
Σst i=0

Then
3
1 s2 X
Z
2
E c,s
(T2 , W0 ) + |KW0 + 2tW0 | + 2 |Ω0j W0 |2 dx
Hs ∩[T1 ,T2 ] t2 t j=0
Z T2
1
c,s
≤E (T1 , W0 ) + k(t + r)F0 kL2x (Σst ) E c,s (t, W0 ) 2 dt
T1

Proof. The result of the statement follows by integrating (3.14) over the region between ΣsT2 ,
ΣsT1 and Hs ∩ [T1 , T2 ], which can be foliated by the hypersurfaces Σst for t ∈ [T1 , T2 ]. 

4. Global existence in the exterior region


The main goal of this section is to prove the global existence of solutions (u, v) to system
(1.1), or equivalently of solutions W to (1.5), in the exterior region D ex under the a-priori
energy assumption (2.1). This result will follow immediately from the proof of proposition
2.1, which is organized in two steps. First we recover sharp pointwise bounds for the so-
lution W from the energy bounds (2.1) by means of weighted Sobolev inequalities. Then
we compute the equation satisfied by the differentiated variable Z γ W , compare it to the
inhomogeneous linear equation (3.1) and estimate perturbatively the source terms to finally
propagate (2.1).
As a result of proving global energy bounds in the exterior region we also obtain bounds
for the higher order conformal energy E5c,ex (t, W ) for all t ≥ 2 and a uniform-in-time control
of the higher order energy on exterior hyperboloids E5ex,h (s, W ) for all s ≥ 2.
This section is organized as follows: in subsections 4.1 and 4.2 we prove some weighted
Sobolev and Hardy inequalities; in subsection 4.3 we recover pointwise bounds for the solution
from the a-priori energy estimate (2.1); finally, subsection 4.4 is dedicated to the proof of
proposition 3.1.

4.1. Weighted Sobolev inequalities.

17
Lemma 4.1. Let β ∈ R. For any sufficiently smooth function w we have

(4.1) sup (2 + r − t)β r 2 |w(t, x, y)|2


Σex
t
ZZ
. (2 + r − t)β+1 (∂r Z ≤2 w)2 + (2 + r − t)β−1 (Z ≤2 w)2 dxdy.
Σex
t

Proof. Let (r, σ) be the spherical coordinates in R3 , r = |x| and σ = x/|x| ∈ S2 . We begin
by observing that the Sobolev embedding H 2 (S2 × S1 ) ⊂ L∞ (S2 × S1 ) implies
X Z
2
sup |w(t, r, σ, y)| ≤ |∇lσ ∂yk w(t, r, σ, y)|2dσdy.
S2 ×S1 0≤l+k≤2

We then remark that for any function v and (t, x, y) ∈ Σex


t
h i
∂r (2 + r − t)β r 2 v(t, x, y)2
= 2(2 + r − t)β r 2 v∂r v + β(2 + r − t)β−1 r 2 v 2 + 2(2 + r − t)β rv 2
≥ 2(2 + r − t)β r 2 v∂r v + β(2 + r − t)β−1 r 2 v 2 ,
so if v is compactly supported in x we can write
Z ∞
β 2 2
∂ρ (2 + ρ − t)β ρ2 v(t, x, y)2 dρ
 
(2 + r − t) r v(t, x, y) = −
r
Z ∞ Z ∞
β 2
(4.2) .β (2 + ρ − t) |v∂ρ v|ρ dρ + (2 + ρ − t)β−1 v 2 ρ2 dρ
Zr ∞ r
Z ∞
β+1 2 2
.β (2 + ρ − t) (∂ρ v) ρ dρ + (2 + ρ − t)β−1 v 2 ρ2 dρ.
r r

By replacing v with ∇lσ ∂yk w(t, r, σ, y) for k + l ≤ 2 we obtain (4.1) in the case where w is
compactly supported. In the general case where w is not compactly supported we consider
a cut-off function χ ∈ C0∞ (R) and apply the inequality (4.1) to χ(ǫr)w for any ǫ > 0
sup (2 + r − t)β r 2 |χ(ǫr)w|2
Σex
t
X ZZ
. (2 + r − t)β+1 (∂r ∇kσ ∂yl w)2 + (2 + r − t)β−1 (∇kσ ∂yl w)2 dxdy
k+l≤2 Σex
t

X ZZ
+ (2 + r − t)β+1 ǫ2 |χ′ (ǫr)|2 (∇kσ ∂yl w)2 dxdy.
k+l≤2 Σex
t

On the intersection of Σex ′ 2 2


t with the support of χ (ǫr) we have that (2 + r − t) ǫ . 1 so
ZZ ZZ
β+1 2 ′ 2 k l 2
(2 + r − t) ǫ |χ (ǫr)| (∇σ ∂y w) dxdy . (2 + r − t)β−1 (∇kσ ∂yl w)2 dxdy.
Σex
t Σex
t

By letting ǫ → 0 we derive (4.1) also in the case of non compactly supported w. 

Slight modifications of the above proof yield the following three results.
18
Lemma 4.2. Let β ∈ R. For a sufficiently regular function w we have
ZZ  
β 2 2
(4.3) sup (2 + r − t) r |w(t, x, y)| . (2 + r − t)β (∂r Z ≤2 w)2 + (Z ≤2 w)2 dxdy.
Σex
t Σex
t

Proof. It follows by estimating v and ∂ρ v in (4.2) with the same weight and using the fact
that 2 + r − t ≥ 1 on Σext . 
Lemma 4.3. Let β ∈ R. For any sufficiently regular function w we have
(4.4) ZZ
β 2 2
sup (2 + r − t) r kw(t, r, ·)kL2(S2 ×S1 ) . (2 + r − t)β+1 (∂r w)2 + (2 + r − t)β−1 w 2 dxdy.
Σex
t Σex
t

Proof. The inequality (4.4) follows by replacing v with the L2 (S2 × S1 ) norm of w in both
left and right hand sides of inequality (4.2). 
Lemma 4.4. Let β ∈ R. For any sufficiently regular function w we have
ZZ
β 2 2
(4.5) sup (2 + r − t) r kw(t, r, ·)kL4 (S2 ×S1 ) . (2 + r − t)β (∂r≤1 Z ≤1 w)2 dxdy.
Σex
t Σex
t

Proof. The inequality (4.5) follows by estimating v and ∂ρ v in (4.2) with the same weight,
then applying the inequality with v replaced by the L4 (S2 × S1 ) norm of w and finally using
the Sobolev injection H 1 (S2 × S1 ) ⊂ L4 (S2 × S1 ). 
4.2. Weighted Hardy inequalities.
Lemma 4.5. Let β > −1. For any compactly supported function w we have
ZZ ZZ
β 2
(2 + r − t) w dxdy . (2 + r − t)β+2 (∂r w)2dxdy.
Σex
t Σex
t

Proof. A simple computation shows that for any β ∈ R and (t, x, y) ∈ Σex t
 2 β+1 β+1
+ (β + 1)r (2 + r − t) ≥ (β + 1)r 2 (2 + r − t)β .
2 β

∂r r (2 + r − t) = 2r(2 + r − t)
Consequently, if β > −1 and w is a compactly supported function
Z Z ∞Z
β 2
(2 + r − t) w dx = (2 + r − t)β r 2 w 2 drdσ
r≥t−1 t−1 S 2
Z ∞Z
1
≤ ∂r (r 2 (2 + r − t)β+1 )w 2 drdσ
β + 1 t−1 S2
2 1
Z Z
β+1 2
=− (2 + r − t) w∂r w r drdσ − w 2 r 2 dσ

β + 1 r≥t−1 β + 1 S2 ×S1 r=t−1
Z 1
 2 Z 2 1

β 2 β+2 2
. (2 + r − t) w dx (2 + r − t) (∂r w) dx .
r≥t−1 r≥t−1

The result of the lemma follows then after further integration on S1 . 


Corollary 4.6. Let β > −1 and Z be any Klainerman vector field. For a sufficiently regular
function w we have
ZZ ZZ
β 2
(2 + r − t) (Zw) dxdy . (2 + r − t)β+2 (∂Z ≤1 w)2 dxdy.
Σex
t Σex
t
19
Proof. For any ǫ > 0 and any fixed cut-off function χ we apply lemma 4.5 to χ(ǫr)Zw and
use the fact that |Zw| . r|∂u| on Σex t . Since rǫ is bounded on the support of χ(ǫr) we have
ZZ
(2 + r − t)β (χ(ǫr)Zw)2 dxdy
Σ ex
ZtZ ZZ
β+2 2 2
. (2 + r − t) |χ(ǫr)| (∂r Zw) dxdy + (2 + r − t)β+2 ǫ2 |χ′ (ǫr)|2 (Zw)2 dxdy
Σex Σex
ZZ t ZZ t

. (2 + r − t)β+2 (∂r Zw)2 dxdy + (2 + r − t)β+2 (∂w)2 dxdy


Σex
t Σex
t

and obtain the result after letting ǫ → 0. 

4.3. Pointwise Bounds. Using the lemmas introduced in the previous sections we can
recover sharp pointwise bounds for the solution W from the a-priori energy bounds (2.1).
Proposition 4.7. Assume that the solution W = (u, v)T to (1.5) satisfies the a-priori energy
bounds (2.1) for some fixed time T0 and some fixed α > 0. There exists l ∈ L1 ([2, T0 ]) such
that we have the following pointwise estimates in DTex0
(4.6) sup |W | . ǫ
S1
α+1
(4.7) sup |∂Z ≤2 W | . C0 ǫr −1 (2 + r − t)− 2

S1
p α
(4.8) sup |T Z ≤2 W | + |∂y Z ≤2 W | + |Z ≤2 W| . C0 ǫr −1 l(t)(2 + r − t)− 2
S1
α
(4.9) sup |ZZ ≤2 W | . C0 ǫr −1 (2 + r − t)− 2 .
S1

Proof. The bounds (4.7) and (4.8) follow immediately from lemma 4.2 with β = α + 1 and
β = α respectively and from (2.2). The pointwise bound (4.9) follows instead applying
lemma 4.1 with β = α and corollary 4.6 to write that
ZZ ZZ
α−1 ≤4 2
(2 + r − t) (ZZ W ) dxdy . (2 + r − t)α+1 (∂Z ≤5 W )2 dxdy.
Σex
t Σex
t

Finally, the bound (4.6) on W is obtained from the integration of (4.7) along the direction
∂q = ∂r − ∂t until the initial time slice t0 = 2 and from the smallness assumption on the
initial data. 
A trivial consequence of the decomposition (1.7) and the bounds (4.8)-(4.9) that will be
useful later in section 5 is the following estimate for the zero-mode W0 of the solution
α
sup ZZ ≤1 W0 . C0 ǫr −1 (2 + r − t)− 2 .

(4.10)
S1

4.4. Propagation of the exterior energy bounds. In order to prove proposition 2.1 we
start by considering any multi-index γ with |γ| = n ≤ 5 and compare the system satisfied
by the differentiated function W γ = (uγ , v γ )T - which is a short hand notation for Z γ W =
(Z γ u, Z γ v) - to the inhomogeneous linear equation (3.1). Here the variable that plays the
role of the linear variable W is W γ . We remind the reader that all vector fields Z in the
family (1.13) are related to the geometry of the problem and are the generators of the Lorentz
20
transformations of the Minkowski space R1+3 . In particular, they preserve the structure of
system (1.1) and equivalently of the equation (1.5).
The equation satisfied by W γ is obtained by commuting Z γ to the equation (1.5)
(4.11) (−∂t2 + ∆x )W γ + (1 + u)∂y2 W γ = Fγ
where the inhomogeneous term Fγ is given by
X X
(4.12) Fγ = −δγ uγ1 ∂y2 W γ2 + N(W γ1 , W γ2 )
|γ1 |+|γ2 |=|γ| |γ1 |+|γ2 |≤|γ|
|γ2 |<|γ|

where δγ = 0 if |γ| = 0, 1 otherwise.


Lemma 4.8. The nonlinear term N(·, ·) in the right hand side of (4.12) is a two-vector of
new linear combinations of the quadratic null forms introduced in (1.3) that arise from the
commutation of the Klainerman vector fields with N1 and N2 .
Proof. The proof follows by computing the commutator terms and iterating such formulas.
For the vector fields Ωij , 1 ≤ i < j ≤ 3, we have

∂j ,
 if k = i
[Ωij , ∂0 ] = 0, [Ωij , ∂k ] = −∂i , if k = j

0, otherwise
which implies that
Ωij Q0 (φ, ψ) = Q0 (Ωij φ, ψ) + Q0 (φ, Ωij ψ)
(
Q0j (φ, ψ), if k = i
Ωij Q0k (φ, ψ) = Q0k (Ωij φ, ψ) + Q0k (φ, Ωij ψ) +
Q0i (φ, ψ), if k = j
and for 1 ≤ h < k ≤ 3


 0, if h = i, j = k




 Qjk (φ, ψ), if h = i, j < k
−Qkj (φ, ψ), if h = i, j > k



Ωij Qhk (φ, ψ) = Qhk (Ωij φ, ψ) + Qhk (φ, Ωij ψ) + Qhj (φ, ψ), if h < j, k = i

−Qik (φ, ψ),

 if h = j, k > i

Qih (φ, ψ), if k = j, h > i





−Qhi (φ, ψ), if k = j, h < i

For the vector fields Ω0i , i = 1, 3, we have


(
∂0 , if k = i,
[Ω0i , ∂0 ] = −∂i , [Ω0i , ∂j ] =
0, otherwise
which implies
Ω0i Q0 (φ, ψ) = Q0 (Ω0i φ, ψ) + Q0 (φ, Ω0i ψ)

−Qij (φ, ψ),
 if i < j
Ω0i Q0j (φ, ψ) = Q0j (Ω0i φ, ψ) + Q0j (φ, Ω0i ψ) + Qji (φ, ψ), if i > j

0, otherwise
21
and for 1 ≤ j < k ≤ 3

Q0k (φ, ψ),
 if j = i
Ω0i Qjk (φ, ψ) = Qjk (Ω0i φ, ψ) + Qjk (φ, Ω0i ψ) + −Q0j (φ, ψ), if k = i

0, otherwise

We have seen in proposition 4.9 that under the a-priori energy assumption (2.1) the solu-
tion W satisfies the pointwise bounds (4.6) and (4.7). The hypothesis of proposition 3.1 are
then fulfilled and we have that
ZZ
γ 2
(4.13) kW kX ex,α + |T W γ |2 + |∂y W γ |2 dσdydt
T0
C[2,T0 ]
α+1

. E ex,α (2, W γ ) + (2 + r − t) 2 Fγ kW γ kXTex,α .

L1t L2xy (DTex ) 0
0

Proof of proposition 2.1. It is enough for our purpose to estimate the weighted norm of the
source term Fγ and prove that for every |γ| ≤ 5
α+1

− α+1
γ 2 2
p
(4.14) (2 + r − t) F . C ǫ (t − 1) l(t),
2 2
ex 0
L2xy (Σt )

where l ∈ L1 ([2, T0 ]). In fact, if we plug (4.14) and a-priori energy bound (2.1) into (4.13)
we obtain that there exists some universal positive constant C so that
2
X ZZ
kW kX 5,α + |T W γ |2 + |∂y W γ |2 dσdydt ≤ CE5ex,α (2, W ) + 2CC03 ǫ3 .
T0
|γ|≤5 C[2,T0 ]

For any fixed constant K > 1 (e.g. K = 2) we can then choose C0 > 0 sufficiently large such
that
C 2 ǫ2
E5ex,α (2, W ) ≤ 0
CK
2
and ǫ0 > 0 sufficiently small so that 2CC0 ǫ < 1/K for ǫ ≤ ǫ0 to finally obtain
2
X ZZ 2C0 ǫ2
|T Z γ W |2 + |∂y Z γ W |2 dσdydt ≤

(4.15) kW kX 5,α + .
T0
C[2,T ] K
|γ|≤5 0

We estimate the different contributions to Fγ separately. In all the estimates that follow
we will use the a-priori energy bound (2.1) and the fact that r > t − 1 in the exterior region.
1. The null terms: we use here the null form representation via the formula (1.15)
N(W γ1 , W γ2 ) ∼ T W γ1 · ∂W γ2 + ∂W γ1 · T W γ2 .
The products in the above right hand side are equivalent given the range of γ1 and γ2 and
we only focus on the analysis of the first one. We distinguish between the different values γ1
and γ2 can take and remind the reader that |γ1 | + |γ2| ≤ |γ| = n ≤ 5.
a. The case |γ1| = 0: here γ2 = γ and we immediately obtain from (4.8) that
α+1
p
γ γ 2 2 −1
(2 + r − t) T W · ∂W ≤ kT W kL∞ kW k ex,α . C ǫ (t − 1) l(t).
2
xy
X T 0
L2xy 0

22
b. The case |γ2| = 0: here γ1 = γ and we obtain from (4.7) that
α+1
1
α
(2 + r − t) T W · ∂W 2 ≤ (2 + r − t) ∂W ∞ (2 + r − t) 2 T W γ L2
γ

2 2
Lxy Lxy xy
p
. C02 ǫ2 (t − 1)−1 l(t).
c. The case |γ1 |, |γ2| > 0: we use spherical polar coordinates and Cauchy-Schwarz inequal-
ity to bound the weighted L2xy (Σex t ) norm of T W
γ1
· ∂W γ2 as follows
2 Z ∞ ZZ
α+1
γ1 γ2
(2 + r − t) T W · ∂W 2 = (2 + r − t)α+1 |T W γ1 |2 · |∂W γ2 |2 r 2 drdσdy
2
Lxy 2 1
Z ∞t−1 S ×S
. (2 + r − t)α+1 kT W γ1 k2L4 (S2 ×S1 ) k∂W γ2 k2L4 (S2 ×S1 ) r 2 dr.
t−1

In the case where |γ1 | ≤ n − 2 we apply the inequality (4.5) to T W γ1 with β = α and the
Sobolev’s injection H 1 (S2 × S1 ) ⊂ L4 (S2 × S1 ) to ∂W γ2 . We derive that
Z ∞
(2 + r − t)α+1 kT W γ1 k2L4 (S2 ×S1 ) k∂W γ2 k2L4 (S2 ×S1 ) r 2 dr
t−1
Z ∞
α
≤n
2 2
(2 + r − t)r −2 ∂Z ≤n W L2 (S2 ×S1 ) r 2 dr

. (2 + r − t) T Z W L2 (Σex )
2
xy t
t−1
α 2 α+1
2
. (t − 1)−2 (2 + r − t) 2 T Z ≤n W L2 ≤n

(2 + r − t) ∂Z W 2 .
2

(Σex ) Lxy (Σex
t )
xy t

In the remaining case where |γ1| = n − 1 and |γ2 | = 1 we apply the inequality (4.5) to ∂W γ2
with β = α + 1 and the injection H 1 (S2 × S1 ) ⊂ L4 (S2 × S1 ) to T W γ1 . We get
Z ∞
(2 + r − t)α+1 kT W γ1 k2L4 (S2 ×S1 ) k∂W γ2 k2L4 (S2 ×S1 ) r 2 dr
t−1
2 ZZ
α+1
≤n
. (2 + r − t) 2 ∂Z W r −2 |T Z ≤n W |2 dxdy

2 ex
Lxy (Σt ) Σex
t
α 2
. (t − 1)−2 (2 + r − t) 2 T Z ≤n W L2 kW k2X n,α .

ex
xy (Σt ) T0

In both scenarios we obtain that


α+1
p
γ1 γ2
(2 + r − t) T W ∂W . C02 ǫ2 (t − 1)−1 l(t).
2

L2xy

2. The uγ1 · ∂y2 W γ2 terms: since |γ1 | ≥ 1 we can write uγ1 = Z uγ̃1 for some |γ̃1 | = |γ1 | − 1.
We can also write ∂y2 W γ2 = ∂y W γ̃2 for some other γ̃2 such that |γ̃2| = |γ2| + 1 and observe
that then |γ̃1| + |γ̃2 | = n. Depending on γ1 = (α1 , β1 ) we can distinguish two cases:
a. The case |α1 | > 0: here we choose γ̃1 so that Z uγ̃1 = ∂uγ̃1 . The products ∂uγ̃1 · ∂y W γ̃2
have the same behavior of the null terms treated in case 1.
b. The case |α1 | = 0: here Z γ1 = Z β1 is a pure product of Klainerman vector fields and
Z uγ̃1 = Zuγ̃1 . We choose the exponents (p1 , p2 ) as follows

(2, ∞)
 if |γ̃1 | = n − 1
(p1 , p2 ) = (∞, 2) if |γ̃2 | = n

(4, 4) otherwise
23
and will place the two factors in Lp1 (S2 × S1 ) and Lp2 (S2 × S1 ) respectively. We use the
Sobolev’s injections H 2 (S2 × S1 ) ⊂ L∞ (S2 × S1 ) and H 1 (S2 × S1 ) ⊂ L4 (S2 × S1 ) to derive
2 Z ∞
α+1
γ̃1 γ̃2
2 2
(2 + r − t)α+1 Zuγ̃1 Lp1 (S2 ×S1 ) ∂y W γ̃2 Lp2 (S2 ×S1 ) r 2 dr

(2 + r − t) Zu ∂y W 2 .
2
Lxy t−1
Z ∞
2 2
(2 + r − t)α+1 ZZ ≤4 u L2 (S2 ×S1 ) ∂y Z ≤5 W L2 (S2 ×S1 ) r 2 dr.

.
t−1

Applying the inequality (4.4) to ZZ ≤4 u with β = α and successively the weighted Hardy
inequality proved in corollary 4.6 with β = α − 1 we find that
2
sup (2 + r − t)α r 2 ZZ ≤4 u L2 (S2 ×S1 )

r≥t−1
α+1
2 α−1
2
. (2 + r − t) 2 ∂r Z ≤5 u + (2 + r − t) 2 ZZ
≤4
u

2 ex
L2xy (Σex
t ) Lxy (Σt )

2
α+1

. (2 + r − t) 2 ∂Z ≤5 u .

ex L2xy (Σt )
We can therefore continue the previous chain of inequalities
2 Z ∞
α+1
≤5 −2 ≤5
2
2 2 1 r 2 dr

. (2 + r − t) 2 ∂r Z u (2 + r − t)r ∂ Z W

ex
y L (S ×S )
L2xy (Σt ) t−1
α 2
. (t − 1)−1−α kW k2X 5,α (2 + r − t) 2 ∂y Z ≤5 W L2

ex
T0 xy (Σt )

and finally conclude that


X α+1
α+1 p
γ1 2 γ2
(2 + r − t) u · ∂y W . C02 ǫ2 (t − 1)− l(t).
2 2
L2xy
|γ1 |+|γ2 |=n
|γ1 |≥1


An immediate consequence of the global energy bounds (2.3) obtained from proposition
2.1 is the following estimate on the higher order energies of the solution W on the truncated
exterior hyperboloids Hsex .
Proposition 4.9. Let W = (u, v)T be a global solution to the Cauchy problem (1.5)-(1.6) in
the exterior region D ex . Then
X ZZ
ex,h
E5 (s, W ) + |T Z γ W |2 + |∂y Z γ W |2 dσdydt . C02 ǫ2 , s ∈ [2, ∞).
|γ|≤5 C[2,s]

Proof. The result follows by applying proposition 3.2 with W = W γ and F = Fγ for any
multi-index γ such that |γ| ≤ 5 and then using the global energy bound (2.3), the estimate
(4.14) of the source term Fγ and the fact that
Z ∞
γ
kF kL1t L2xy (H ex ) . kFγ kL2xy (Σex
t )
dt . C02 ǫ2 .
[2,s]
2

We conclude this section with the derivation of a bound for the higher order exterior
conformal energy of W0 as well as for the higher order conformal energy on portions of the
hyperboloid Hs in the exterior region D ex .
24
Proposition 4.10. Assume we have a solution W = (u, v)T to the Cauchy problem (1.5)-
(1.6) in the exterior region DTex0 that satisfies the a-priori exterior energy bounds (2.1). Then
(4.16)
XZ 2
sup E4c,ex (t, W0 ) + |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ | + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt
[2,T0 ] |γ|≤4 C[2,T0 ]

. C02 ǫ2 ln T0 ,
where the implicit constant only depends on C0 .
Proof. Let us fix |γ| ≤ 4. By integrating (4.11) over the sphere S1 we obtain that W0γ is
solution to the following linear inhomogeneous wave equation
(4.17) (−∂t2 + ∆x )W0γ = Fγ0
with source term
Z Z
Fγ0 = γ
F dy − u · ∂y2 W γ dy
S1 S1
(4.18) X Z X Z
γ1 γ2
= N(W , W ) dy + ∂y uγ1 · ∂y W γ2 dy.
|γ1 |+|γ2 |≤|γ| S1 |γ1 |+|γ2 |=|γ| S1

When applying proposition 3.6 with W0 = W0γ and F0 = Fγ0 we derive that for all T ∈ [2, T0 ]
(4.19) Z
E c,ex
(T, W0γ ) + |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ |2 + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt
C[2,T ]
Z T
1
≤E c,ex
(2, W0γ ) + k(t + r)Fγ0 kL2x (Σex
t )
E c,ex (t, W0γ ) 2 dt
2
where
X X
k(t + r)Fγ0 kL2x (Σex
t )
≤ k(t + r)N(W γ1 , W γ2 )kL2xy + k(t + r)∂y uγ1 · ∂y W γ2 kL2xy .
|γ1 |+|γ2 |≤|γ| |γ1 |+|γ2 |=|γ|

We estimate the different contributions to Fγ0


separately. We start by observing that since
|γ1 | + |γ2| ≤ 4, at least one of the two multi-indices has length less or equal than 2. We call
this index γj and will place the factor carrying this index in L∞ , the other one in L2 .
1. The ∂y uγ1 · ∂y W γ2 term: we use the pointwise bound (4.8) and the energy bound (2.1)
and obtain that there exists l ∈ L1 ([2, T0 ]) such that
p
k(t + r)∂y uγ1 · ∂y W γ2 k 2 ex . C0 ǫ l(t) ∂y Z ≤4 W 2 ex . C 2 ǫ2 l(t).

Lxy (Σt ) Lxy (Σt ) 0

2. The null terms: here we use the null form representation via the formula (1.14) and
the relation ∂ = t−1 Z to write
1 1 t−r
(4.20) N(W γ1 , W γ2 ) = ZW γ1 · ∂W γ2 + ∂W γ1 · ZW γ2 + ∂W γ1 · ∂W γ2 .
t t t
The first two products in the above right hand side are equivalent given the range of γ1
and γ2 so we will just analyze the first one. In the case where |γ1 | ≤ 2 we deduce from the
pointwise bound (4.9) that
(t + r)t−1 ZW γ1 · ∂W γ2 2 ex . C0 ǫ (t + r)r −1 t−1 ∂W γ2 2 ex . C0 ǫt−1 E ex,0 (t, W ) 21 .

L (Σ ) xy L (Σ ) 4 xy
t t
25
In the case where |γ1 | ≥ 3 we estimate ∂W γ2 with the pointwise bound (4.7) and decompose
W γ1 using (1.7). On the one hand we apply the Poincaré inequality to obtain
(t + r)t−1 ZWγ1 · ∂W γ2 2 ex . C0 ǫ (t + r)r −1t−1 ∂y ZWγ1 2 ex . C0 ǫt−1 E5ex,0 (t, W ) 21

L (Σ ) xy t L (Σ ) xy t

and on the other hand we use corollary 4.6 with β = α − 1 to get



(t + r)t−1 ZW γ1 · ∂W γ2 2 ex . C0 ǫt−1 −α+ (α−1) γ1

0 L (Σ )
(2 + r − t) 2 ZW0
xy t L2xy (Σex
t )
1
. C0 ǫt−1 E5ex,α (t, W ) 2 .
The last quadratic term in the right hand side of (4.20) is estimated using again (4.7)
1−α

k(t + r)(t − r)/t ∂W γ1 · ∂W γ2 kL2xy (Σex . C ǫt−1
(2 + r − t) ∂Z ≤4
W
2

t )
0 2 ex
Lxy (Σt )
1
. C0 ǫt−1 E4ex,0 (t, W ) , 2

which concludes that


1
kN(W γ1 , W γ2 )kL2xy (Σex
t )
. C0 ǫt−1 E5ex,α (t, W ) 2 .

The combination of step 1 and step 2 with the energy bound (2.1) yields
(4.21) k(t + r)Fγ0 kL2xy (Σex
t )
. C02 ǫ2 l(t) + C02 ǫ2 t−1 ,

which plugged in (4.19) for all |γ| ≤ 4 gives


c,ex
XZ 2
E4 (T, W0 ) + |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ | + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt
|γ|≤4 C[2,T ]
Z T
1
E4c,ex (2, W0 ) C02 ǫ2 l(t) + C02 ǫ2 t−1 E4c,ex (t, W0 ) 2 dt

. +
2
. E4c,ex (2, W0 ) + C02 ǫ2 sup E4c,ex (t, W0 ) + C02 ǫ2 ln T.
[2,T0 ]

Therefore, assuming ǫ ≪ 1 sufficiently small we get that


XZ 2
sup E4c,ex (t, W0 ) + |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ | + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt
[2,T0 ] |γ|≤4 C[2,T0 ]

. E4c,ex (2, W0 ) + C02 ǫ2 ln T0


so the result of the proposition follows from the smallness of the conformal energy at the initial
time, which in turn follows from the assumptions on the initial data. 
Lemma 4.11. Let s ≥ 2 and 2 < T1 < T2 be such that the portion of hyperboloid Hs in
the time strip [T1 , T2 ] is entirely contained in the exterior region D ex . Assume W = (u, v)T
is the solution to the Cauchy problem (1.5)-(1.6) in D ex satisfying the global energy bounds
(2.1). Then there exists constants C1 , C2 > 0 such that
3
s2 X
 
XZ 1 γ γ 2 γ 2 2 2 2 2 T2
2
|KW0 + 2tW0 | + 2 |Ω0j W0 | dx ≤ C1 C0 ǫ ln T1 + C2 C0 ǫ ln
|γ|≤4 Hs ∩[T1 ,T2 ] t t i=0 T1
26
Proof. The result follows from the application of proposition 3.7 with W0 = Z γ W with
|γ| ≤ 4, from the observation that Σ2t ⊂ Σex
t for all t ∈ [T1 , T2 ] and from the estimates (4.21)
and (4.16). 
4.5. A sharper pointwise bound for the non-zero modes. The scope of this section
is to show that the pointwise bound (4.8) for ∂y Z ≤2 W ≤2
p and Z W obtained by Sobolev’s
injections can be improved and that one can replace l(t) by the explicit decay in time
t−1/2 .
Proposition 4.12. We have the following pointwise estimate in the exterior region D ex
1 α
sup |∂y Z ≤2 W | + |Z ≤2 W| . C0 ǫr −1 t− 2 (2 + r − t)− 2 .
S1

Proof. Thanks to lemma 4.2 and the Poincaré inequality we have that
ZZ
α 2 ≤2 2 ≤2
(2 + r − t)α |∂ ≤1 ∂y Z ≤4 W |2 dxdy,

(2 + r − t) r sup |∂y Z W | + |Z W| ≤
S1 Σex
t

so it will be enough to prove that


ZZ
I(t) := (2 + r − t)α |∂y Z ≤5 W |2 dxdy . C02 ǫ2 t−1 , t ∈ [2, ∞).
Σex
t

We fix t ∈ [2, ∞) and let k ∈ Z be such that t ∈ [2k , 2k+1). Since I ∈ L1 ([2, ∞)) there exists
τk ∈ [2k , 2k+1 ] such that
I(τk ) . 2−k . t−1 .
Observe that in the above inequalities the implicit constants are independent of k. We can
assume τk < t (the case τk > t being similar) and use the fundamental theorem of calculus
to write that Z k
|I(t) − I(τk )| ≤ |∂t I(s)| ds.
τk
By integration over Σex
t of the equality (3.6) with W = ∂y Z ≤4 W and ω(z) = (1 + z)α and
the smallness of u given by (4.6) we derive that

1
ZZ
α
(2 + r − t) ∂y u∂t W∂y W − ∂t u(∂y W) + (2 + r − t)α F≤5 ∂t W dxdy
2

|∂t I(t)| .
Σex
t
2
where F≤5 = |γ|≤5 Fγ and Fγ is the nonlinear term appearing in the equation satisfied
P

by Z γ W given by (4.12). On the one hand, the pointwise bounds (4.7) and (4.8) and the
energy bound (2.1) imply that

1
ZZ
(2 + r − t) ∂y u∂t W∂y W − ∂t u(∂y W) dxdy . C03 ǫ3 (t − 1)−1 l(t).
α 2

Σt ex 2
On the other hand, from the proof of proposition 2.1 it follows that F≤5 = Fa + Fb , where
Fa contains the terms analyzed in case 1 and 2.a such that
α+1 p
k(2 + r − t) 2 Fa kL2xy . C02 ǫ2 (t − 1)−1 l(t)
and Fb contains the products treated in case 2.b that only satisfy
α+1
α+1 p
b
(2 + r − t) F . C02 ǫ2 (t − 1)− 2 l(t).
2
L2xy
27
α
The estimate for Fb with the weaker weight (2 + r − t) 2 can actually be improved to the
following one, obtained using Sobolev injections and the pointwise bound (4.8)
2 Z ∞
(2 + r − t) α2 Fb 2 2 . α+1
≤5 −2 ≤5
2
2 2 1 r 2 dr

(2 + r − t) 2 ∂Z u r ∂ Z W

L ex
y L (S ×S )
xy L2xy (Σt ) t−1
. C04 ǫ4 (t −2
− 1) l(t).

When W = ∂y Z ≤5 this yields


ZZ p α
(2 + r − t)α F≤5 ∂t W dxdy .C02 ǫ2 (t − 1)−1 l(t) (2 + r − t) 2 ∂y Z ≤5 W L2

x,y
Σex
t

.C03 ǫ3 (t − 1)−1 l(t),

which finally allows us to conclude that


Z t
|I(t) − I(τk )| . C03 ǫ3 (s − 1)−1 l(s) ds . C03 ǫ3 t−1 .
τk

5. Pointwise Estimates in the Interior Region


The goal of this section is to recover pointwise estimates for the solution W = (u, v)T to
in
(1.5) in the interior hyperbolic region H[2,s 0]
under the a-priori assumptions (2.4)-(2.6) and
to propagate the a-priori pointwise estimate (2.6) on ZW0 .
We remind the reader that we proved uniform-in-time energy bounds for the solution
on the exterior truncated hyperboloids Hsex (see proposition 4.9) as well as the exterior
pointwise bound (4.10) for ZW0 , therefore if we suppose that A, B ≫ C0 we can think of
the a-priori bounds (2.4)-(2.6) as being valid not only on Hsin but on the whole hyperboloid
Hs for any s ∈ [2, s0 ]. The pointwise estimates we will obtain in this section will then be
valid in the whole hyperbolic strip between H2 and Hs0 , that we denote by H[2,s0]

H[2,s0 ] := (t, x) : 4 ≤ t2 − r 2 ≤ s20 × S1 .




5.1. Pointwise estimates from Klainerman-Sobolev inequalities. A first subset of


pointwise estimates for W0 and W is immediately obtained from (2.4) and (2.5) via the
following Sobolev inequality on hyperboloids, whose proof can be found in [21].

Lemma 5.1. Let W = W(t, √ x) be a sufficiently regular function in the cone C = {t > r}.
For all (t, x) ∈ C , let s = t2 − r 2 and B(x, t/3) be the ball centered at x with radius t/3.
Then
XZ p 2
2 −3 γ
|W(t, x)| ≤ Ct Z W( s2 + |y|2, y) dy

|γ|≤2 B(x,t/3)

where C is a positive universal constant and Zj = xj ∂t + t∂j , j = 1, 3.

28
Lemma 5.2. Let In,k denote the set of multi-indices of type (n, k). Under the a-priori
energy bounds (2.4) and (2.5) we have the following pointwise estimates in H[2,s0 ]
1
(5.1) |∂Z ≤3 W0 (t, x)| . ǫt− 2 s−1 ,
3
(5.2) |∂Z ≤3 W0 (t, x)| . ǫt− 2 ,
3
X
kZ γ W(t, x, ·)kL∞ (S1 ) + ∂y≤1 Z γ W(t, x, ·) L2 (S1 ) . ǫt− 2 sδk+2 ,

(5.3) k = 0, 3
I3,k
1
X
(5.4) k∂tx Z γ W(t, x, ·)kL2 (S1 ) . ǫt− 2 s−1+δ5 ,
|γ|=3

and
¯ ≤1 Z γ W(t, x, ·) 2 1 . ǫt− 25 sδk+3 ,
X
(5.5) ∂∂ y L (S )
k = 0, 2
I2,k
7
∂¯2 ∂ ≤1 Z γ W(t, x, ·)
X
. ǫt− 2 sδk+4 ,

(5.6) y L2 (S1 )
k = 0, 1
I1,k
5
∂¯ ∂tx Z W(t, x, ·)
2
. ǫt− 2 s−1+δ5 .

(5.7) L2 (S1 )

Moreover
(5.8) sup |W | . ǫ.
H[2,s0 ]

Proof. The estimates (5.1) to (5.4) are immediate consequence of lemma 5.1 and (2.4)-(2.5).
The estimates on the ∂ derivatives of W are deduced
√ from (5.3) and (5.4) using the fact that
∂ = t−1 Z. Finally, if we set W̃0 (s, x) = W0 ( s2 + r 2 , x) we see that
Z s Z s
τ √
|W̃0 (s, x) − W̃0 (2, x)| ≤ 2 2
∂τ W̃0 (τ, x) dτ = ∂t W0 ( τ + x , x) dτ . ǫ

2 2 t

and hence derive from the smallness of the initial data that
|W0 (t, x)| . |W̃0 (s, x) − W̃0 (2, x)| + |W̃0 (2, x)| . ǫ.
The combination of the above estimates with (5.3) yields (5.8). 

5.2. Improved pointwise estimates on the non-zero modes. The bounds for the W
obtained in lemma 5.2 via Sobolev embeddings are affected by the small growth in s of
the energies and are not sharp but they can be improved if one studies more closely the
equation satisfied by W. Enhancing such bounds, and in particular (5.3), will be fundamental
to propagate the a-priori pointwise bound (2.6) later in proposition 5.6. We make use of
the following result, which is motivated by the work of Klainerman [15] and whose proof is
an adaptation of a similar estimate for Klein-Gordon equations initially proved in [20], later
revisited in [6] in the case of Klein-Gordon equations with variable mass.
Proposition 5.3. Assume W is a solution of the following equation
(5.9) x,y W + u∆y W = F, (t, x, y) ∈ R1+3 × S1
29
R √
such that S1 Wdy = 0. For every fixed (t, x) in the cone C = {t > r}, let s = t2 − r 2 and
Ytx , Atx , Btx be the functions defined as follows
2
3
Z
Ytx2 (λ) := λ Wλ + (S W)λ + λ3 (1 + uλ )|∂y Wλ |2 dy

(5.10)
S1 2

1
(5.11) Atx (λ) := sup (S u)λ + sup |∂y uλ|

S1 2λ S1
Z
(5.12) 2
Btx (λ) = λ−1 |(RW)λ |2 dy
S1
λt λr

where fλ (t, x, y) = f s
, s ,y and
3
RW(t, x, y) = s2 ∂¯i ∂¯i W + xi xj ∂¯i ∂¯j W + W + 3xi ∂¯i W − s2 F.
4
Then W satisfies the following inequality in the hyperbolic region H[2,∞)
 Z s  R
3 1 s
Btx (λ)dλ e 2 Atx (λ) dλ .

s 2 kWkL2 (S1 ) + k∂y WkL2 (S1 ) + s 2 kS WkL2 (S1 ) . Ytx (2) +
2
3
Proof. For every fixed (t, x, y) ∈ H[2,∞) we define ωtxy (λ) := λ W( λts , λx
2
s
, y) to be the evalu-
3/2
ation of W on the hyperboloid Hλ , then dilated by factor λ . We have that
 
1/2 3
ω̇txy (λ) = λ Wλ + (S W)λ
2
and
1
ω̈txy (λ) = 1 (P W)λ
λ2
where
3
P W = W + 3(t∂t W + xi ∂i W) + (t2 ∂t2 W + 2txi ∂i ∂t W + xi xj ∂i ∂j W).
4
Using the equation (5.9) we derive that ωtxy satisfies the following equation
 
3
− 1
2 ¯i ¯ i j ¯ ¯ 3 i ¯
ω̈txy − (1 + uλ ) ∆y ωtxy = −λ 2 Fλ + λ 2 s ∂ ∂i W + x x ∂i ∂j W + W + 3x ∂i W .
4 λ
We drop the lower indices in ωtxy (λ) in order to have a lighter notation and simply denote
it by ω(λ) in what follows. We multiply the above equation by ∂λ ω and integrate over S1
Z
∂λ ω ∂λ2 ω − (1 + uλ ) ∆y ω dy

S1
 Z  Z
d 1
Z
2
= |∂λ ω| dy + (1 + uλ )∂y ω ∂y ∂λ ω dy + ∂λ ω ∂y uλ ∂y ω dy
dλ 2 S1 S1 S1
 Z 
d 1 1
Z Z
2 2 2
= |∂λ ω| + (1 + uλ)|∂y ω| dy − ∂λ uλ |∂y ω| dy + ∂λ ω ∂y uλ ∂y ω dy.
dλ 2 S1 2 S1 S1
We obtain that
d 2
Ytx (λ) . Atx (λ)Ytx2 (λ) + Btx (λ)Ytx (λ)

30
with Atx , Btx , Ytx as in the statement and from the Gronwall lemma
 Z s  R
s
Ytx (s) . Ytx (2) + Btx (λ) dλ e 2 Atx (λ) dλ .
2
Finally, from the definition of Ytx , the Poincaré inequality and the fact that s ≥ 1 we get
3  1
s 2 kWkL2(S1 ) + k∂y WkL2 (S1 ) + s 2 kS WkL2 (S1 ) . Ytx (s).

Proposition 5.4. Under the a-priori assumptions (2.4)-(2.6) we have
3
  3
∂ W L2 (S1 ) + ∂y ∂ W L2 (S1 ) + sup t 2 S ∂ j W L2 (S1 ) . ǫ,
j j

(5.13) sup t 2 j = 0, 1
H[2,s0 ] H[2,s0 ]

3
3
  t2
(5.14) sup t 2 kZWkL2 (S1 ) + k∂y ZWkL2 (S1 ) + sup kS ZWkL2 (S1 ) . ǫsσ ,
H[2,s0 ] H[2,s0 ] s
and
1 1
sup st 2 k∂ 2 WkL2 (S1 ) + k∂y ∂ 2 WkL2 (S1 ) + sup t 2 kS ∂ 2 WkL2 (S1 ) . ǫ,

(5.15)
H[2,s0 ] H[2,s0 ]

1
  1
(5.16) sup st 2 k∂ZWkL2 (S1 ) + k∂y ∂ZWkL2 (S1 ) + t 2 kS ∂ZWkL2 (S1 ) . ǫsσ .
H[2,s0 ]

Proof. For any fixed j = 0, 2 and k = 0, 1 we compare the equation satisfied by the differen-
tiated functions ∂ j W and ∂ k ZW respectively with the equation (5.9) and apply the result of
proposition 5.3. A simple computation shows that
x,y ∂ j W + u∆y ∂ j W = F1j , j = 0, 2
x,y ∂ k ZW + u∆y ∂ k ZW = F2k , k = 0, 1
with source terms given by
Z Z
0
F1 = N(W, W ) − N(W, W )dy + ∂y u · ∂y W dy
S1 S1
X
F1j = ∂ j F10 − ∂ h u · ∂y2 ∂ j−h W, j = 1, 2
1≤h≤j

and
F20 = ZF10 − Zu · ∂y2 W
F21 = ∂ZF10 − ∂Zu · ∂y2 W − ∂u · ∂y2 ZW − Zu · ∂y2 ∂W.
From proposition 5.3 we have that
 Z s  R
3 1 s
Btx (λ)dλ e 2 Atx (λ) dλ

s kWkL2(S1 ) + k∂y WkL2(S1 ) + s kS WkL2(S1 ) . Ytx (2) +
2 2

with W = {∂ j W, ∂ k ZW : j = 0, 2, k = 0, 1} and corresponding source term F = {F1j , F2k :


j = 0, 2, k = 0, 1}. In order to obtain the bounds in the statement we need to estimate
the quantities Ytx (2), Atx (λ) and Btx (λ) defined in (5.10), (5.11), (5.12) for all the different
values of W and F.
31
1. The Atx (λ) term: this is the same for all values of W. Here we decompose u = u0 + u
and rewrite the scaling vector field as follows:
(5.17) S = (t − r)∂t + (r − t)∂r + (t∂r + r∂t ).
Using the pointwise bounds (5.1)-(5.3) and the fact that s/t ≤ 1 in the interior of the light
cone we derive that
s 3
Atx (λ) . sup |(∂u)λ | + λ−1 |(Zu)λ| + |(∂y u)λ | . ǫλ− 2 +δ5
y∈S1 t + r

and consequently
Z s
(5.18) Atx (λ)dλ . ǫ.
1
2. The Ytx (2) term: the functions appearing here are evaluated on the hyperboloid H2 .
From the bound (5.3) on W, the smallness of u given by (5.8) and the decomposition (5.17)
it follows that for all values of W under consideration
 s  32
(5.19) |Ytx (2)| . kW2 kL2 (S1 ) + k(S W)2 kL2 (S1 ) + k1 + u2 kL∞ (S1 ) k(∂y W)2 kL2 (S1 ) . ǫ .
t
The Btx (λ) terms are the only ones for which we need to distinguish between the different
values of W and hence of F.
3. The Btx (λ) term for W = Z ≤1 W: from the pointwise bounds (5.3), (5.5) and (5.6) we
immediately obtain the following estimate
(5.20)
 
3
s ∂¯ ∂¯i W + x x ∂¯i ∂¯j W + W + 3x ∂¯i W
−1/2 2 i i j i

λ
4
λ L2 (S1 )

r2  3
 
1 r
3
¯2 ¯ − 12 −2+δ5 s 2
. λ 1 + 2 k(∂ W)λ kL (S ) + λ k(∂W)λ kL (S ) + λ kWλ kL (S ) . ǫλ
2 2 1 2 2 1 2 1 .
s s t
As concerns the estimate of the L2 (S1 ) norm of the source term Fλ we distinguish betweeen
the different values {F10 , F11 , F20 } it takes in this case, after observing that
1 3
−2 2 
λ s F λ
= λ 2 kF k 2 1 .
λ L (S )
2 1L (S )

3.a. The case F = F10 : this corresponds to W = W. We decompose each occurrence of W


in the formula for F10 using (1.7)
N(W, W ) = N(W0 , W0 ) + 2N(W0, W) + N(W, W)
and observe that F10 does not make appear the zero-mode interactions W0 × W0 . Simply
using the pointwise bounds (5.1) and (5.3) we deduce that
kF10 kL2 (S1 ) . ǫ2 t−2 s−1+δ5 ,
and hence that
3 3
 s 2
(5.21) λ 2 kFλ kL2 (S1 ) . ǫ2 λ− 2 +2δ5 , F = F10
t
which summed up with (5.20) yields
Z s  s  32
(5.22) Btx (λ)dλ . ǫ .
2 t
32
The combination of (5.22) with (5.18) and (5.19) implies (5.13) for j = 0.
3.b The case F = F11 : this corresponds to W = ∂W with ∂ = {∂t , ∂x , ∂y }. When W = ∂y W
the argument used above yields (5.13). When W = ∂tx W the only additional contribution
that needs to be analyzed is ∂tx u0 · ∂y2 W, which appears in F11 after using the decomposition
(1.7). This product is bounded using (5.1) and (5.3)
k∂u0 · ∂y2 WkL2 (S1 ) . k∂u0 kL∞ (S1 ) k∂y2 WkL2 (S1 ) . ǫ2 t−2 s−1+δ5 .
We obtain (5.21) for F = F11 and consequently (5.22), which summed up with (5.18) and
(5.19) gives us (5.13) for j = 1.
3.c The case F = F20 : this corresponds to W = ZW. The term ZF10 is estimated using the
argument in case 3.a and the only contribution to analyze is Zu · ∂y2 W. We again decompose
u using (1.7) and obtain from (5.3) that
Zu · ∂ 2 W 2 1 . ǫ2 t−3+2δ5 ,

y L (S )

while from the a-priori bound (2.6) on ZW0 and the enhanced bound (5.13) on ∂y2 W we get
Zu0 · ∂y2 W 2 1 . ǫ2 t− 25 sσ .

L (S )

Therefore from (5.21) and the above bounds we derive that F20 satisfied the following estimate
 5
2 − 23 +2δ5 s 2 −1+σ s 2
3
 2  s 2
0
λ k(F2 )λ kL2 (S1 ) . ǫ λ
2 +ǫ λ . ǫ2 λ−1+σ
t t t
which together with (5.20) gives
Z s  s  32
Btx (λ)dλ . ǫsσ
2 t
and finally (5.14) when combined with (5.18) and (5.19).
4. The Btx (λ) term for W = {∂ 2 W, ∂ZW}: here the bounds (5.3), (5.5) and (5.7) give us
that
 
−1/2 2 i
λ 3
¯ ∂¯i W + W + 3x ∂¯i W
i

s ∂
4
λ L2 (S1 )
(5.23)
1 r
  3 −δ5
3
¯ 2 ¯ − 21 −2+δ5 s 2
. λ k(∂ W)λ kL2 (S1 ) + λ k(∂W)λ kL2 (S1 ) + λ kWλ kL2 (S1 ) . ǫλ
2 2
s t
and
1  s  12
−2 i j
(5.24) λ (x x ∂ i ∂ j W)λ 2 1 . ǫλ−2+δ5 .

L (S ) t
We then separately analyze the L2 (S1 ) norm of Fλ when F = {F12 , F21 }.
4.a The case F = F12 : this corresponds to W = ∂ 2 W with ∂ = {∂t , ∂x , ∂y }. The argument
to perform here is analogous to the one in 3.b in that one should first consider the case where
W = ∂y2 W, followed by W = ∂y ∂tx W and finally by W = ∂tx 2
W. It is a simple exercise to verify
that
3 3
 s  52 −δ5
λ 2 (F12 )λ L2 (S1 ) . ǫ2 λ− 2 +δ5

t
33
which summed up with (5.23) and (5.24) gives
Z s  s  12
Btx (λ)dλ . ǫ .
2 t
The combination of the above estimate with (5.18) and (5.19) yields then (5.15).
4.b The case F = F21 : this corresponds to W = ∂ZW. The term ∂ZF10 enjoys the same
estimate (5.21) as F10 in case 3.a. We decompose each occurrence of u in the remaining
contributions to F21 using (1.7) and focus on discussing the terms involving products u0 × W.
The products u × W are estimated using (5.3) and are bounded by ǫ2 t−3+2δ5 .
When W = ∂y ZW the only term that is left to control is the L2 (S1 ) norm of Zu0 · ∂y3 W,
which is estimated using the a-priori bound (2.6) and (5.15) so that
Zu0 · ∂ 2 ∂W 2 1 ≤ kZu0kL∞ (S1 ) k∂ 2 ∂WkL2 (S1 ) . ǫ2 t− 32 s−1+σ .

y L (S ) y

Therefore in this case


3 3
 s 2  s  32  s  23
λ 2 (F21 )λ L2 (S1 ) . ǫ2 λ− 2 +2δ5 + ǫ2 λ−1+σ . ǫ2 λ−1+σ

t t t
which together with (5.23) and (5.24) gives
Z s  1
σ s 2
(5.25) Btxy (λ)dλ . ǫs .
2 t
Summing this estimate up with (5.18) and (5.19) we deduce (5.16) with ∂ = ∂y .
When W = ∂tx ZW the remaining terms to control in the L2 (S1 ) norm are

∂Zu0 · ∂y2 W + ∂u0 · ∂y2 ZW + Zu0 · ∂y2 ∂W.

The first term is estimated using the bounds (5.1) and (5.13) as follows
∂Zu0 · ∂ 2 W 2 1 ≤ k∂Zu0 kL∞ (S1 ) k∂ 2 WkL2 (S1 ) . ǫ2 t−2 s−1 ,

y L (S ) y

the second one using (5.1) and (5.16) with ∂ = ∂y proved above
∂u0 · ∂y2 ZW 2 1 ≤ k∂u0 kL∞ (S1 ) k∂y2 ZWkL2 (S1 ) . ǫ2 t−1 s−2+σ ,

L (S )

the third one using the a-priori bound (2.6) and (5.15)
Zu0 · ∂y2 ∂W 2 1 ≤ kZu0kL∞ (S1 ) k∂y2 ∂WkL2 (S1 ) . ǫ2 t− 32 s−1+σ .

L (S )

We obtain in this case that


3
 s 2 s  s  32 s
2 − 32 +2δ5 2 − 32 +σ
λ 2 k(F21 )λ kL2 (S1 ) .ǫ λ +ǫ λ 2 −1+σ
+ǫ λ 2 −1+σ
.ǫ λ
t t t t
which together with (5.23) and (5.24) again gives (5.25). Summing this estimate up with
(5.18) and (5.19) finally yields (5.16) when ∂ = ∂t x. 
34
5.3. The propagation of the a-priori pointwise bound. In order to propagate the a-
priori bound (2.6) on the Z derivative of W0 in the interior region H[2,s0 ] we use L∞ − L∞
type estimates. For this we give a closer look to the wave equation satisfied by ZW0 and
use the enhanced pointwise bounds on the solution recovered in the previous subsection to
estimate the nonlinear terms. The argument that follows will in addition provide us with a
decay estimate for W0 without derivatives and hence improve on (5.8).
We will make use of the following lemma, due to Alinhac [1].
Lemma 5.5. Let W0 be the solution to tx W0 = F0 with zero initial data and suppose that
F0 is spatially compactly supported satisfying the following pointwise bound
|F0 (t, x)| ≤ Ct−2−ν (t − |x|)−1+µ
for some positive constant C. Then
C
|W0 (t, x)| . (t − |x|)µ−ν t−1 .
µν
Proposition 5.6. There exists a constant B > 0 sufficienly large and ǫ0 sufficiently small
such that, for any 0 < ǫ < ǫ0 if W = (u, v)T is solution to the Cauchy problem (1.5)-(1.6) and
satisfies the a-priori bounds (2.4)-(2.6) in the interior region H[2,s0 ] and the global energy
bounds (2.1) in the exterior region D ex , then in H[2,s0 ] it actually satisfies the enhanced
pointwise bound
|ZW0 (t, x)| ≤ Bǫt−1 sσ .
Proof. We are interested in propagating the a-priori pointwise bound (2.6) and to recover
improved estimates on ZW0 in the interior region H[2,s0] , which is contained in the cone
{t > r + 1}. We consider a cut-off function χ ∈ C0∞ (R) such that χ(z) = 1 for |z| ≤ 1/2 and
χ(z) = 0 for |z| > 1, and decompose W0 into the sum W01 + W02 , where W01 and W02 solve to
the following Cauchy problems
W01 = (1 − χ(t − r))F0 , (W01 , ∂t W01 )|t=2 = (0, 0)
x W02 = χ(t − r)F0 , (W02 , ∂t W02 )|t=2 = (W0 , ∂t W0 )|t=2 .
with F0 given by Z Z
F0 = N(W, W )dy + ∂y u · ∂y W dy.
S1 S1
The scope of such decomposition is to estimate ZW01
and ZW02 separately and in particular
1
to apply lemma 5.5 to ZW0 , which is solution to a wave equation with zero data and source
term supported in the interior of the cone {t = r + 1/2}
ZW01 = (1 − χ)(t − r)ZF0 − [Z, χ(t − r)]F0 , (ZW01 , ∂t ZW01 )|t=2 = (0, 0).
In order to do so, we need to estimate F0 and ZF0 in the region t > r + 1/2. We remark
that the commutator term is uniformly bounded and also supported in {t > r + 1/2}.
We decompose each occurrence of W in N(W, W ) by means of (1.7) and use the null
structure representation (1.14) of N for the quadratic interactions W0 ×W0 . From (5.1),(5.2)
and (5.5) we deduce the following
s2
Z
|Z ≤1 N(W0 , W0 )| dy . |∂Z ≤1 W0 ||∂W0 | + |∂Z ≤1 W0 ||∂W0 | + |∂Z ≤1 W0 ||∂W0 | . ǫ2 t−2 s−1
S1 t2
35
Z Z
3
|Z ≤1 N(W0 , W)| dy . |Z ≤1 (∂W0 · ∂W)| dy . ǫt− 2 ∂Z ≤1 W L2 (S1 )

S1 S1

Z Z
≤1
≤1
Z (∂y u · ∂y W) dy . k∂Z ≤1 WkL2 (S1 ) k∂WkL2 (S1 ) .

|Z N(W, W)| dy +
S1 S1
p
The improved bounds (5.13) and (5.16) on W coupled with the fact that s = (t + r)(t − r) imply
that in {t > r + 1/2}
5 1
|F0 (t, x)| . ǫ2 t− 2 ht − ri− 2
and
5 σ 1 σ
|ZF0 (t, x)| . ǫ2 t− 2 + 2 ht − ri− 2 + 2 ,
so from lemma 5.5 we obtain that

|ZW01 (t, x)| ≤ Cǫ2 t−1 (t − r)σ ,

for some constant C = C(A, B) that depends quadratically on A and B.


The remaining term ZW02 is estimated via the Klainerman-Sobolev inequality of lemma
5.1 in terms of the higher order conformal energy of W02 on hyperboloids. More precisely,
we have that for any (t, x) ∈ H[2,s0]
1 1
X
t 2 s|ZW02(t, x)| . E2c,in (s, W02 ) 2 + (s/t)ZZ γ W02 2

L (Hs ∩[T s ,T s ]) x 1 2
|γ|≤2

where s2 = t2 − r 2 ,pT1s = (s2 + 1)/2 is the time the hyperboloid Hs intersects the cone
t = r + 1, and T2s = s2 + |y|2 with |y − x| = t/3 (the second term in the above right hand
side should be omitted when T2s ≤ T1s ). Since the source term in the equation satisfied by
Z γ W02 is supported in the exterior region t < r + 1, we obtain from proposition 3.5 with
W0 = Z γ W02 , the bound (4.16) (which is valid also for W02 ) with T0 = (s2 + 1)/2, and the
smallness assumption on the initial data, that

E c,in (s, Z γ W02 ) ≤ E c,in (2, Z γ W02 ) + C02 ǫ2 ln s . C02 ǫ2 ln s.

Moreover, from lemma 4.11 with Tj = Tjs for j = 1, 2 and the observation that ln(T2s /T1s ) .
ln s we also derive that
(s/t)ZZ γ W 2 2 2
X
2 2

0 L (Hs ∩[T s ,T s ]) . C0 ǫ ln s.
x 1 2
|γ|≤2

This concludes that


1
|ZW02 (t, x)| ≤ C̃C0 ǫt− 2 s−1+σ
for a universal constant C̃ and therefore
1
|ZW0 (t, x)| ≤ |ZW01 (t, x)| + |ZW02(t, x)| ≤ C̃C0 ǫt− 2 s−1+σ + Cǫ2 t−1 (t − r)σ ≤ Bǫt−1 sσ

if we choose B sufficiently large so that B ≥ 2C̃C0 and ǫ0 sufficiently small so that 2Cǫ ≤ B.
36
6. Energy Estimates in the Interior Region
The goal of this section is to propagate the interior energy bounds (2.4) and (2.5) on
the two components W0 and W of the solution W to the Cauchy problem (1.5)-(1.6). We
remind the reader that for any multi-index γ the differentiated function W γ = (uγ , v γ ) is
solution to the equation (4.11) with source term (4.12), and that its zero-mode W0γ solves
the inhomogeneous wave equation (4.17) with source term (4.18).
We first start by recovering an energy bound for the higher order conformal energy of W0 .
Such bound follows from (2.4) and (2.5) as well as from the pointwise estimates obtained in
section 5. It will be necessary for the propagation of (2.5), and the computations that leads
to it will be useful in the proof of proposition 6.2.
Proposition 6.1. Assume the solution W = (u, v)T to (1.5)-(1.6) satisfies the a-priori
estimates (2.4)-(2.6) in the interior region H[2,s0] as well as the global exterior energy bounds
(2.1) in the exterior region D ex . Then
(6.1) sup Ekc,in (s, W0 ) ≤ Cǫs1+2µk , s ∈ [2, s0 ], k = 0, 4
[2,s]

where µk = δk if k ≤ 3 and µ4 = δ2 + δ4 .
Proof. We consider here only multi-indices γ of type (k, k), i.e. γ = (0, β) with |β| = k and
Z γ = Z β . Applying proposition 3.5 with W0 = W0γ and F0 = Fγ0 we derive that
(6.2) Z s
γ γ 1
c,in c,in
sup E (τ, W0 ) ≤ E (2, W0 ) + kτ Fγ0 kL2 (Hτin ) E c,in (τ, W0γ ) 2 dτ
[2,s] 2
Z
2
+ |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ | + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt.
C[2,s]

The integral over the boundary C[2,s] corresponds to the integral in the left hand side of
(4.16) for T0 = (s2 + 1)/2 hence
Z
(6.3) |(t + r)(∂t W0γ + ∂r W0γ ) + 2W0γ |2 + (t − r)2 (|∇W0γ |2 − (∂r W0γ )2 ) dσdt . ǫ2 ln s.
C[2,s]

The different contributions to the inhomogeneous term Fγ0 are estimated separately below:
1. The ∂y uγ1 · ∂y W γ2 terms: we will place the factor whose index has length smaller than
k/2 in L2y L∞ 2
x and the remaining one in Lxy . Using the enhanced pointwise bounds (5.13)
and (5.14), as well as the pointwise bound (5.3) in the case where |γ1 | = |γ2 | = 2 (which
appears only if k = 4) we see that
h i
− 23 1
X
γ1 γ2 in σ in δ4 in
k∂y u · ∂y W kL1y L2x (Hτin ) . ǫτ Ek (τ, W ) + τ Ek−1 (W ) + ν4 τ E2 (W )
2

|γ1 |+|γ2 |=k≤4

where ν4 = 1 if k = 4 and 0 otherwise.


2. The null terms: we apply the decomposition (1.7) to both factors W γ1 and W γ2 so that
N(W γ1 , W γ2 ) = N(W0γ1 , W0γ2 ) + N(Wγ1 , W0γ2 ) + N(W0γ1 , Wγ2 ) + N(Wγ1 , Wγ2 ).
We express N(W0γ1 , W0γ2 ) using the representation formula (1.14)
t−r
N(W0γ1 , W0γ2 ) = ∂W0γ1 · ∂W0γ2 + ∂W0γ1 · ∂W0γ2 + ∂W0γ1 · ∂W0γ2
t
37
and estimate all terms in the above right hand side using (5.1) and (5.2) as follows
Z
N(W0γ1 , W0γ2 ) dy (t/τ )∂Z ≤3 W0 ∞ in (τ /t)∂Z ≤k W0 2 in

1 2 in . Lx (Hτ ) Lx (Hτ )
S Lx (Hτ )

+ ∂Z ≤k W0 L2 (H in ) ∂Z ≤3 W0 L∞ (H in )

x τ x τ
− 32 1
. ǫτ Ekin (τ, W ) 2 .
We neglect the null structure for all the remaining quadratic forms. The products involving
W0 × W are equivalent given the range of γ1 and γ2 and are estimated using (5.1) whenever
|γ1 | ≤ 3 and (5.13) otherwise
Z
γ
N(W 1 , Wγ2 ) dy ≤3 ≤k

1 0 2 in ≤ k(t/τ )∂Z W0 k L ∞ (H in ) (τ /t)∂Z
x τ
W L2xy (Hτin )
S Lx (Hτ )

+ (τ /t)∂Z ≤4 W0 L2 (H in ) k(t/τ )∂WkL∞



in
x τ xy (Hτ )

− 32 1
. ǫτ Ekin (τ, W ) 2 .
The quadratic term involving products W × W is estimated as in step 1.
Summing up the estimates in step 1 and 2, using the a-priori energy bounds (2.4)-(2.5),
and choosing σ, δk−1 small so that σ + δk−1 ≤ δk we obtain that
(
3 δk if k ≤ 3
(6.4) kFγ0 kL2x (Hτin ) . ǫ2 τ − 2 +µk with µk =
δ4 + δ2 if k = 4.
Plugging the above estimate together with (6.3) into (6.2) and using the smallness assump-
tions on the initial data we finally obtain
Z s
c,in c,in 1 1
2 2Cǫ
sup Ek (τ, W0 ) . Ek (2, W0) + ǫ s ln s + ǫ2 τ − 2 +µk Ekc,in (τ, W0 ) 2 dτ
[2,s] 2

.ǫ s2 1+2µk
+ǫ 2
sup Ekc,in (τ, W0 )
[2,s]

and the result of the proposition follows choosing ǫ sufficiently small. 


We observe that from lemma p 5.1, the conformal energy bound (6.1) and lemma 4.11 with
2
T1 = (s + 1)/2 and T2 = s2 + |y|2 for |y − x| = t/3, we have the following additional
pointwise bound for the Z derivatives of W0 in H[2,s0]
(
1 1 δk if k ≤ 3
(6.5) |ZZ j W0 (t, x)| . ǫt− 2 s− 2 +µj+2 for j = 0, 2, where µk =
δ4 + δ2 if k = 4.
We now have all the ingredients to propagate the a-priori energy bounds (2.4) and (2.5).
Proposition 6.2. There exists a constant A > 0 sufficiently large, some small parameters
σ ≪ δk ≪ δk+1 for k = 1, 4 and ǫ0 > 0 sufficiently small such that, if for any 0 < ǫ < ǫ0
the solution W = (u, v)T to the Cauchy problem (1.5)-(1.6) satisfies the a-priori bounds
in
(2.4)-(2.6) in the hyperbolic region H[2,s 0]
and the global exterior energy bounds (2.1) in the
ex
exterior region D , then it also satisfies the enhanced energy bound
(6.6) E5in (s, W0 ) ≤ A2 ǫ2 , s ∈ [2, s0 ].
38
Proof. For any |γ| ≤ 5, the equation (4.17) satisfied by W0γ has the same structure as the
inhomogeneous wave equation (3.13), therefore proposition 3.4 with W0 = W0γ and F0 = Fγ0
implies that
Z Z s
γ γ γ 2 1
in in
E (s, W0 ) . E (2, W0 ) + |T W0 | dσdt + kFγ0 kL2x (Hτin ) E in (τ, W0γ ) 2 dτ
C[2,s] 2

for all s ∈ [2, s0 ]. The implicit constant in the above right hand side is independent of s0 .
The integral over the boundary C[2,s] has already been proved to be finite and small in
proposition 4.9. The source term Fγ0 has also already been estimated in the case |γ| ≤ 4 (see
(6.4) in the previous proposition). When |γ| = 5 we simply use the estimate (5.3) to deduce
3 1
X
k∂y uγ1 · ∂y W γ2 kL1y L2x (Hτin ) + kN(Wγ1 , Wγ2 )kL1y L2x (Hτin ) . ǫτ − 2 +δ4 E5in (t, W ) 2
|γ1 |+|γ2 |=5
|γ1 |≥1

and (5.1), (5.2) and (5.13)


3 1
X
kN(W0γ1 , W0γ2 )kL1y L2x (Hτin ) + kN(W0γ1 , Wγ2 )kL1y L2x (Hτin ) . ǫτ − 2 E5in (s, W ) 2 .
|γ1 |+|γ2 |≤5

Therefore
3 1
X
(6.7) kFγ0 kL2x (Hτin ) . ǫτ − 2 +δ4 E5in (τ, W ) 2
|γ|≤5

and from the a-priori energy bound (2.5) and the exterior energy bound (4.15) we obtain
that, for some fixed K > 1 and some universal constant C > 0
Z s
in in C02 ǫ2 3
E5 (s, W0 ) ≤ E5 (2, W0 ) + + A2 Cǫ3 τ − 2 +δ4 +δ5 dτ
K 2
2 2
C ǫ
≤ E5in (2, W0 ) + 0 + A2 Cǫ3 .
K
The wished improved energy bound then follows choosing K = 3, ǫ0 small such that 3Cǫ20 < 1
and A ≥ C0 sufficiently large so that
A2 ǫ20
E5in (2, W0 ) ≤ .
3

Proposition 6.3. There exists a constant A > 0 sufficiently large, some small parameters
σ ≪ δk ≪ δk+1 for k = 1, 4 and ǫ0 > 0 sufficiently small such that, if for any 0 < ǫ < ǫ0
the solution W = (u, v)T to the Cauchy problem (1.5)-(1.6) satisfies the a-priori bounds
in
(2.4)-(2.6) in the hyperbolic region H[2,s 0]
and the global exterior energy bounds (2.1) in the
ex
exterior region D , then it also satisfies the enhanced energy bound
(6.8) in
E5,k (s, W) ≤ A2 ǫ2 s2δk , s ∈ [2, s0 ].
Proof. We start by considering a multi-index γ of type (n, k) with k ≤ n ≤ 5 and compare
the equation satisfied by Wγ = (uγ , vγ )T with the linear inhomogeneous equation (3.1). We
have that
x,y Wγ + u ∂y2 Wγ = Fγ − Fγ0 , (t, x, y) ∈ R1+3 × S1
39
so applying proposition 3.3 with W = Wγ and F replaced by Fγ − Fγ0 we derive the following
inequality
ZZ
in γ in γ
E (s, W ) . E (2, W ) + |T Wγ |2 + |∂y Wγ |2 dσdydt
C[2,s]
(6.9) Z s
1
+ kFγ − Fγ0 kL2xy (Hτin ) E in (τ, Wγ ) 2 dτ.
2

All quadratic terms in F can be estimated similarly to those in Fγ0 – which satisfy the bound
γ

(6.7) – except for the products uγ01 · ∂y2 W γ2 when Z γ1 = Z β1 is a pure product of Klainerman
vector fields, obtained after using the decomposition (1.7) on uγ1 . The reason is that the
L2x norm of Z β1 u0 is not controlled by the higher order energies of W0 but rather by the
higher order conformal energy of W0 (up to multiplication with a factor (s/t)). Those are
the only terms that need to be analyzed here since the argument already used in proposition
6.2 coupled with the Poincaré’s inequality shows that
3 1
X X
uγ1 · ∂y2 Wγ2 2 kN(W γ1 , W γ2 )kL2xy (Hτin ) . ǫτ − 2 +δ4 E5in (τ, W ) 2 .

L (H in )
+
xy τ
|γ1 |+|γ2 |=|γ| |γ1 |+|γ2 |≤|γ|
|γ2 |<|γ|

Depending on the starting multi-index γ1 = (α1 , β1 ) we distinguish between two cases:


1. The case |α1 | > 0: here we can write uγ01 = ∂uγ̃01 for some other index γ̃1 with
|γ̃1 | = |γ1| − 1 and estimate the product ∂uγ̃01 · ∂y2 Wγ2 as done for N(W0γ1 , Wγ2 ) in the proof of
proposition 6.2.
2. The case |α1 | = 0: in this case γ1 is of type (k1 , k1 ) with k1 ≤ k ≤ n and γ2 is of
type (n − k1 , k2 ) with k1 + k2 = k. We write uγ01 = Zu0γ̃1 for some other index γ̃1 of type
(k1 − 1, k1 − 1) and distinguish between the different values k1 can take.
If k1 = n then |γ2 | = 0 and we get from the pointwise bound (5.13) and the conformal
energy bound (6.1) that
1
γ̃1 2 c,in
. ∂y2 W L2 L∞ (H in ) En−1 (τ, W0 ) 2 . ǫ2 τ −1+µn−1 .

Zu0 · ∂y W
L2xy (Hτin ) y x τ

If k1 = n − 1 we have |γ2 | = 1, so from the pointwise bounds (5.15), (5.16) and the
conformal energy bound (6.1)
1
γ̃1 2 γ2 c,in
. ∂y2 Wγ2 L2 L∞ (H in ) En−2 (τ, W0 ) 2 . ǫ2 τ −1+µn−2 +σ .

Zu0 · ∂y W
L2xy (Hτin ) y x τ

If k1 = n − 2 then |γ2| = 2 and n = 3, 5. We distinguish the following cases depending on


if Z γ2 is of the form ∂ 2 , ∂Z or Z 2 .
a. The case Z γ2 = ∂ 2 : here we use (5.3) and (6.1) to derive that
1
γ̃1 2 γ2 c,in
. ∂y2 ∂ 2 W L2 L∞ (H in ) En−3 (τ, W0 ) 2 . ǫ2 τ −1+µn−3 +δ2 .

Zu0 · ∂y W 2 in
Lxy (Hτ ) y x τ

b. The case Z γ2 = ∂Z: when n = 3 we have k1 = 1 and hence |γ̃1 | = 0. From the
pointwise bound (2.6) and the a-priori energy bound (2.5) we find
1
Zu0 · ∂y2 Wγ2 2 in 2 −1+σ+δ1

in
L (H )
. kZu 0 k L ∞ (H in ) E
τ 3,1 (τ, W ) 2 . ǫ τ .
xy τ
40
When n = 4, 5 we use (5.3) and (6.1) to derive that
1
γ̃1 2 γ2 c,in
. ∂y2 ∂ZW L2 L∞ (H in ) En−3 (τ, W0 ) 2 . ǫ2 τ −1+µn−3 +δ3 .

Zu0 · ∂ W
L2xy (Hτin ) y x τ

c. The case Z γ2 = Z 2 : when n = 3 we estimate Zu0γ̃1 using (2.6) while for n = 4, 5 we


use (6.5). Together with the a-priori bound (2.5) these give that
1
γ˜1 2 γ2
Zu0 · ∂y W 2 . kZu0γ˜1 kL∞ (Hτin ) E3,2
in
(τ, W ) 2 . ǫ2 τ −1+νn +δ2 .
in Lxy (Hτ )

where ν3 = σ, ν4 = µ3 and ν5 = µ4 .
The final cases to treat correspond to k1 = n − 3 and |γ2 | = 3, which appears for n = 4, 5,
and k = n − 4 and |γ2 | = 4 which only appears when n = 5. If k1 = 1 then |γ2| = n − 1 and
n = 4, 5, so using (2.5) and (2.6) we obtain
1
Zu0 · ∂y2 Wγ2 2 in
(τ, W ) 2 . ǫ2 τ −1+σ+δn−1 , n = 4, 5.

L (H in )
. kZu0kL∞ (Hτin ) En,n−1
xy τ

If k1 = 2 then |γ2 | = 3 and n = 4, so the bounds (2.5) and (6.5) yield


2 1
Z u0 · ∂y2 Wγ2 2 . kZ 2 u0 kL∞ (Hτin ) E4,3
in
(τ, W ) 2 . ǫ2 τ −1+2δ3 .

L (H in )
xy τ

Choosing appropriately σ ≪ δk ≪ δk+1 for k = 1, 4 so that δk+1 is bigger than some linear
combination of σ and δj with j ≤ k, we finally obtain that
γ1 2 γ 2 −1+δk
u0 · ∂ W 2 2 in . ǫ τ
y Lxy (Hτ )

with an implicit constant that depends on A and B. The same bound holds then true for
Fγ so plugging it into (6.9) together with the estimate for Fγ0 , the exterior energy estimate
(4.15), and the a-priori energy bounds (2.5), gives us that
Z s
in in CC02 ǫ2
E5,k (s, W) ≤ CE5,k (2, W) + + A2 Cǫ3 τ −1+2δk dτ.
2 2
The end of the proof follows finally by choosing appropriately the constants A and ǫ0 . 
References
[1] S. Alinhac. Semilinear hyperbolic systems with blowup at infinity. Indiana Univ. Math. J., 55(3):1209–
1232, 2006.
[2] L. Andersson, P. Blue, Z. Wyatt, and S.-T. Yau. Global stability of spacetimes with supersymmetric
compactifications. Preprint, arXiv:2006.00824v1, 2020.
[3] A. Bachelot. Problème de Cauchy global pour des systèmes de Dirac-Klein-Gordon. Ann. Inst. H.
Poincaré Phys. Théor., 48(4):387–422, 1988.
[4] D. Christodoulou and S. Klainerman. The global nonlinear stability of the Minkowski space, volume 41
of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993.
[5] S. Dong, P. G. LeFloch, and Z. Wyatt. Global evolution of the U(1) Higgs Boson: nonlinear stability
and uniform energy bounds. Ann. Henri Poincaré, 22(3):677–713, 2021.
[6] S. Dong and Z. Wyatt. Stability of a coupled wave-Klein-Gordon system with quadratic nonlinearities.
J. Differ. Equations, 269(9):7470–7497, 2020.
[7] S. Dong and Z. Wyatt. Two dimensional wave–Klein-Gordon equations with semilinear nonlinearities.
Preprint, 2020.
[8] B. Ettinger. Well-posedness of the equation for the three-form field in eleven-dimensional supergravity.
Trans. Amer. Math. Soc., 367(2):887–910, 2015.
[9] V. Georgiev. Global solution of the system of wave and Klein-Gordon equations. Math. Z., 203(4):683–
698, 1990.
41
[10] M. Ifrim and A. Stingo. Almost global well-posedness for quasilinear strongly coupled wav-Klein-Gordon
systems in two space dimensions. Preprint, arXiv:1910.12673v1, 2019.
[11] A. Ionescu and B. Pausader. The Einstein-Klein-Gordon coupled system: global stability of the
Minkowski solution. To appear in Annals of Math Studies, 2020.
[12] A. D. Ionescu and B. Pausader. On the global regularity for a wave-Klein-Gordon coupled system. Acta
Math. Sin. (Engl. Ser.), 35(6):933–986, 2019.
[13] T. Kaluza. Zum Unitätsproblem der Physik. Int. J. Mod. Phys. D, 27(14):1870001, 2018.
[14] S. Katayama, T. Ozawa, and H. Sunagawa. A note on the null condition for quadratic nonlinear Klein-
Gordon systems in two space dimensions. Comm. Pure Appl. Math., 65(9):1285–1302, 2012.
[15] S. Klainerman. Global existence of small amplitude solutions to nonlinear Klein-Gordon equations in
four space-time dimensions. Comm. Pure Appl. Math., 38(5):631–641, 1985.
[16] S. Klainerman. The null condition and global existence to nonlinear wave equations. In Nonlinear sys-
tems of partial differential equations in applied mathematics, Part 1 (Santa Fe, N.M., 1984), volume 23
of Lectures in Appl. Math., pages 293–326. Amer. Math. Soc., Providence, RI, 1986.
[17] S. Klainerman, Q. Wang, and S. Yang. Global solution for massive Maxwell-Klein-Gordon equations.
Comm. Pure Appl. Math., 73(1):63–109, 2020.
[18] O. Klein. Quantum Theory and Five-Dimensional Theory of Relativity. (In German and English). Z.
Phys., 37:895–906, 1926.
[19] P. G. LeFloch and Y. Ma. The hyperboloidal foliation method, volume 2 of Series in Applied and Com-
putational Mathematics. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2014.
[20] P. G. LeFloch and Y. Ma. The global nonlinear stability of Minkowski space for self-gravitating massive
fields. Comm. Math. Phys., 346(2):603–665, 2016.
[21] P. G. LeFloch and Y. Ma. The global nonlinear stability of Minkowski space for self-gravitating massive
fields, volume 3 of Series in Applied and Computational Mathematics. World Scientific Publishing Co.
Pte. Ltd., Hackensack, NJ, 2018.
[22] P. H. Lesky and R. Racke. Nonlinear wave equations in infinite waveguides. Comm. Partial Differential
Equations, 28(7-8):1265–1301, 2003.
[23] H. Lindblad and I. Rodnianski. The global stability of Minkowski space-time in harmonic gauge. Ann.
of Math. (2), 171(3):1401–1477, 2010.
[24] Y. Ma. Global solutions of non-linear wave-Klein-Gordon system in two space dimension: semi-linear
interactions. Preprint, arXiv:1712.05315, 2017.
[25] Y. Ma. Global solutions of quasilinear wave-Klein-Gordon system in two-space dimension: completion
of the proof. J. Hyperbolic Differ. Equ., 14(4):627–670, 2017.
[26] Y. Ma. Global solutions of quasilinear wave-Klein-Gordon system in two-space dimension: technical
tools. J. Hyperbolic Differ. Equ., 14(4):591–625, 2017.
[27] Y. Ma. Global solutions of nonlinear wave-Klein-Gordon system in one space dimension. Nonlinear
Anal., 191:111641, 57, 2020.
[28] Y. Ma. Global solutions of nonlinear wave-Klein-Gordon system in two spatial dimensions: a prototype
of strong coupling case. J. Differential Equations, 287:236–294, 2021.
[29] J. Metcalfe, C. D. Sogge, and A. Stewart. Nonlinear hyperbolic equations in infinite homogeneous
waveguides. Comm. Partial Differential Equations, 30(4-6):643–661, 2005.
[30] J. Metcalfe and A. Stewart. Almost global existence for quasilinear wave equations in waveguides with
Neumann boundary conditions. Trans. Amer. Math. Soc., 360(1):171–188, 2008.
[31] T. Ozawa, K. Tsutaya, and Y. Tsutsumi. Normal form and global solutions for the Klein-Gordon-
Zakharov equations. Ann. Inst. H. Poincaré Anal. Non Linéaire, 12(4):459–503, 1995.
[32] A. Stingo. Global existence of small amplitude solutions for a model quadratic quasi-linear coupled
wave-Klein-Gordon system in two space dimension, with mildly decaying Cauchy data. To appear in
Memoirs of the AMS.
[33] D. Tataru. Strichartz estimates for second order hyperbolic operators with nonsmooth coefficients. III.
J. Amer. Math. Soc., 15(2):419–442, 2002.
[34] K. Tsutaya. Global existence of small amplitude solutions for the Klein-Gordon-Zakharov equations.
Nonlinear Anal., 27(12):1373–1380, 1996.

42
[35] Y. Tsutsumi. Global solutions for the Dirac-Proca equations with small initial data in 3 + 1 space time
dimensions. J. Math. Anal. Appl., 278(2):485–499, 2003.
[36] Y. Tsutsumi. Stability of constant equilibrium for the Maxwell-Higgs equations. Funkcial. Ekvac.,
46(1):41–62, 2003.
[37] Q. Wang. Global existence for the Einstein equations with massive scalar fields. 2015. Lecture at the
workshop Mathematical Problems in General Relativity.
[38] Q. Wang. An intrinsic hyperboloid approach for Einstein Klein-Gordon equations. J. Differential Geom.,
115(1):27–109, 2020.
[39] E. Witten. Instability of the Kaluza-Klein vacuum. Nuclear Physics B, 195(3):481 – 492, 1982.
[40] Z. Wyatt. The weak null condition and Kaluza-Klein spacetimes. J. Hyperbolic Differ. Equ., 15(2):219–
258, 2018.

École Polytechnique & CNRS


Email address: cecile.huneau@polytechnique.edu

École Polytechnique
Email address: annalaura.stingo@gmail.com, annalaura.stingo@polytechnique.edu

43

You might also like